26
arXiv:2011.12173v1 [quant-ph] 24 Nov 2020 A game of quantum advantage: linking verification and simulation Daniel Stilck Fran¸ ca 1 and Raul Garcia-Patron 2 1 QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2 School of Informatics, University of Edinburgh, Edinburgh EH8 9AB, UK Abstract We present a formalism that captures the process of proving quantum superiority to skeptics as an interactive game between two agents, supervised by a referee. Bob, is sam- pling from a classical distribution on a quantum device that is supposed to demonstrate a quantum advantage. The other player, the skeptical Alice, is then allowed to propose mock distributions supposed to reproduce Bob’s device’s statistics. He then needs to pro- vide witness functions to prove that Alice’s proposed mock distributions cannot properly approximate his device. Within this framework, we establish three results. First, for random quantum circuits, Bob being able to efficiently distinguish his distribution from Alice’s implies efficient approximate simulation of the distribution. Secondly, finding a polynomial time function to distinguish the output of random circuits from the uniform distribution can also spoof the heavy output generation problem in polynomial time. This pinpoints that exponential resources may be unavoidable for even the most basic verifica- tion tasks in the setting of random quantum circuits. Beyond this setting, by employing strong data processing inequalities, our framework allows us to analyse the effect of noise on classical simulability and verification of more general near-term quantum advantage proposals. 1 Introduction The transition from the reign of classical computers to quantum superiority is expected not to be a singular event but rather a process of accumulation of evidence. It will most probably happen through an iterative process of claims of proofs and refutations until a consensus is reached among the scientific community. The series of claims a few years back of the advantage of quantum annealers followed by refutations and an intense debate inside the quantum computation community [SSSV14, SS14, BRI + 14, WRB + 13] can be seen as an example of that. Similarly, the recent claim of quantum advantage [AAB + 19] was followed by a growing interest in its potential simulation by a classical device [HZN + 20, PGN + 19]. Ideally, one would like to demonstrate the advantage of quantum computers solving a well-established hard computational problem, such as factoring large numbers or sim- ulating large-sized molecules. Such demonstration will most likely need a fault-tolerant quantum computer, which will not be available in the near-future. Thus, a lot of attention has been focused in the last years to quantum advantage proposals based on sampling from the outcomes of random quantum circuits, a task considered achievable. This effort culminated in the recent landmark experiment of [AAB + 19]. 1

simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

arX

iv:2

011.

1217

3v1

[qu

ant-

ph]

24

Nov

202

0

A game of quantum advantage: linking verification and

simulation

Daniel Stilck Franca1 and Raul Garcia-Patron2

1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark2School of Informatics, University of Edinburgh, Edinburgh EH8 9AB, UK

Abstract

We present a formalism that captures the process of proving quantum superiority toskeptics as an interactive game between two agents, supervised by a referee. Bob, is sam-pling from a classical distribution on a quantum device that is supposed to demonstratea quantum advantage. The other player, the skeptical Alice, is then allowed to proposemock distributions supposed to reproduce Bob’s device’s statistics. He then needs to pro-vide witness functions to prove that Alice’s proposed mock distributions cannot properlyapproximate his device. Within this framework, we establish three results. First, forrandom quantum circuits, Bob being able to efficiently distinguish his distribution fromAlice’s implies efficient approximate simulation of the distribution. Secondly, finding apolynomial time function to distinguish the output of random circuits from the uniformdistribution can also spoof the heavy output generation problem in polynomial time. Thispinpoints that exponential resources may be unavoidable for even the most basic verifica-tion tasks in the setting of random quantum circuits. Beyond this setting, by employingstrong data processing inequalities, our framework allows us to analyse the effect of noiseon classical simulability and verification of more general near-term quantum advantageproposals.

1 Introduction

The transition from the reign of classical computers to quantum superiority is expectednot to be a singular event but rather a process of accumulation of evidence. It willmost probably happen through an iterative process of claims of proofs and refutationsuntil a consensus is reached among the scientific community. The series of claims a fewyears back of the advantage of quantum annealers followed by refutations and an intensedebate inside the quantum computation community [SSSV14, SS14, BRI+14, WRB+13]can be seen as an example of that. Similarly, the recent claim of quantum advantage[AAB+19] was followed by a growing interest in its potential simulation by a classicaldevice [HZN+20, PGN+19].

Ideally, one would like to demonstrate the advantage of quantum computers solvinga well-established hard computational problem, such as factoring large numbers or sim-ulating large-sized molecules. Such demonstration will most likely need a fault-tolerantquantum computer, which will not be available in the near-future. Thus, a lot of attentionhas been focused in the last years to quantum advantage proposals based on samplingfrom the outcomes of random quantum circuits, a task considered achievable. This effortculminated in the recent landmark experiment of [AAB+19].

1

Page 2: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

The classical hardness of computing the outcome probability of random circuits hasbeen reduced to standard complexity theoretic assumptions in various settings [HM17,BFNV19, LBR17, Mov18, BJS11, AC17, BMS17, HHB+19]. For instance, in [BFNV19]the authors prove that it is ♯P hard to compute the exact probability of the outputsof random quantum circuits unless the polynomial hierarchy collapses. This result canbe extended to devices with very low noise by assuming a couple of widely-acceptedconjectures. Despite this significant progress on putting the classical hardness of samplingfrom a distribution close to the outputs of the random circuit on solid grounds, equivalenthardness statements are not known to hold for the levels of noise that affect currentquantum computing architectures. Moreover, certifying closeness to the ideal distributionin total variation distance requires an exponential number of samples [HKEG19]. Thus,it is both not feasible to verify closeness in total variation, and the actual distance isunlikely to be in the regime where a quantum advantage has been proven.

These shortcomings have shifted the interest to benchmarks of advantage that arethought to be more robust against noise and that are known to be verifiable with afeasible number of samples. Prominent examples are the heavy output generation problem(XHOG) [AC17, AG19] and the related linear cross-entropy benchmarking (linear XEB)fidelity [NRK+18, BIS+18, AAB+19]. The recent quantum advantage experiment usedlinear XEB as a benchmark for the quantum state generated by the 53 qubit device.However, these approaches have two main drawbacks. First, they require the computationof the probability of sampled strings under the circuit’s ideal distribution, which consumesa running time growing exponentially with the system’s size. Secondly, the number ofrequired samples grows exponentially with the size of the system for a constant noiserate [AAB+19]. Thus, the linear XEB verification approach requires us to be in the“sweet spot” where both the number of samples needed given the noise level and the sizeof the circuit are not too large to render the verification impossible. This approach is notscalable to larger system sizes with current levels of noise. Besides, it is still unknownhow to reduce the hardness of the heavy output generation problem (XHOG) [AC17,AG19, Kre20] to standard complexity-theoretic assumptions.

This difficulty of finding efficient certification protocols and benchmarks for randomcircuit sampling extends to other proposals of near-term quantum advantage, such asboson sampling [AA14b]. This has sparked interest in simpler sanity checks, such asefficiently distinguishing the output distribution from the uniform and other “easy” dis-tributions [AA14a, CMS+14, PWR+19, SVB+14]. As mentioned in [HKEG19], theseverification forms are manifestly weaker than certifying the total variation distance anddo not preclude the possibility of the device being classically simulable. However, in thesetting of random circuit sampling, not even an efficient verification test that allows usto distinguish the outputs from the uniform distribution is known.

Independent of the recent quantum advantage experiment, the development of moreefficient certification techniques of quantum advantage that can be scaled with the in-creasing size of the quantum computer is an area of relevance for near-term quantumcomputation [BKVV20]. In parallel, it is important to develop a better understandingof how noise reduces the power of quantum computers and how noise affects quantumadvantage proposals or near term applications of quantum devices.

To these ends, in this work, we envision this certification process as an interactive gamebetween two agents, Alice that uses classical computing resources, and Bob that holds a(noisy) quantum computer and wants to convince Alice of its computational advantage.They are both supervised by the referee Robert. To win, Bob has to find functionsthat allow him to efficiently distinguish the output of his device from every alternativedistribution Alice proposes. In turn, Alice needs to propose alternative distributions thatapproximate the expectation values obtained by Bob’s quantum computer, otherwise she

2

Page 3: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

loses the game.Central to our result is the connection between the mirror descent algorithm [TRW05,

Bub15] and the proposed framework of the certification game. This allows us to connectdistinguishing probability distributions from a target quantum distribution and learningan approximate classical description of the latter. Our framework is inspired by a recentresult of the authors [FGP20]. There, we show how to use mirror descent, strong dataprocessing inequalities and related concepts to analyse the performance of noisy quantumdevices performing optimization. In contrast, this article’s main result also hold fornoiseless circuits and our overarching goal is to formally link the hardness of verificationof quantum advantage proposals and their approximate classical simulation.

2 Summary of results

We envision a quantum advantage demonstration as a game played between Bob, who issampling from a (noisy) quantum circuit and claiming that it demonstrates a quantumadvantage, and Alice, who is skeptical of Bob’s claim and defends that her classicalcomputer can mimic Bob’s behaviour efficiently.

As Bob is the person looking to convince the others that his quantum computer hasan advantage over Alice’s classical device, we place the burden of the demonstration onhim. In addition, Bob discloses publicly a description of the hardware and the quantumalgorithm his device is implementing. The game consists of different rounds at which Al-ice can propose an alternative hypothesis to the claim that Bob has achieved a quantumadvantage. At the beginning of the game, they both agree on a distinguishability param-eter ǫ > 0, which captures how close Alice needs to match Bob’s result, and confidenceprobability δ, which captures the probability of the outcome of the game being correct.In what follows, we denote the probability distribution Bob is sampling from by ν.

2.1 The quantum advantage game

At the beginning of the first round, Alice discloses an alternative hypothesis to quan-tum advantage in the form of a randomized classical algorithm sampling from a givendistribution µ0(x), for example the uniform distribution. It is then Bob’s role to refuteAlice’s proposal. He does so by providing a function f1 : 0, 1n → [0, 1] such that withprobability at least 1− δ,

Eµ0(f1)− Eν(f1) ≥ ǫ. (1)

Alice is then allowed to update her hypothesis to µ1(x) given the information gainedfrom the first round of the game, which Bob needs to refute, providing a new witness f2.Alice’s new distribution then needs to pass both tests for the functions f1 and f2. Thegame continues with Bob proposing new distinguishing functions ft+1 and Alice mockdistributions µt. The game ends if one of two players is declared defeated following a setof previously established rules.

For example, Alice could concede defeat if she takes too much time to propose a newcandidate or sample from it. Similarly, Bob could be forced to acknowledge his defeat ifhe takes too long to offer a new witness that challenges Alice. In some sense, the rulesshould be consistent with the process of building community consensus on the validity ofa quantum advantage result.

It is important to remark that the condition in Eq. (1) must be checked by a refereeRobert, to whom Alice and Bob provide samples at each round. That is, they give Robertenough samples of their distributions to estimate the expectation values by computing theempirical average for the functions fi on the samples. The required number of samples can

3

Page 4: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

be estimated by e.g. an application of Hoffding’s inequality, as we show later. Note thatthis guarantees that Alice cannot cheat by using Bob’s samples to find better distributions.For instance, if Bob’s quantum computer is solving an NP problem, then knowing thesamples themselves would give Alice an efficient classical strategy. In order to suppressstatistical anomalies, we also make the number of samples depend on the current round.More specifically, the number of samples for round t should be O(tǫ−2 log(tδ−1)). Thischoice ensures that the overall probability of an error occurring remains O(δ).

In order for the verification game to work, we may further request that sampling fromAlice’s distributions, evaluating Bob’s test functions fi and the size of the messages sentto Robert to be tasks that need be achieved efficiently with the resources at hand. Inwhat follows, we define efficient as a consumption of resources that scales polynomiallywith the size of the problem, i.e., the number of qubits of the quantum computer. But amore at hands definition, where we impose constraints on their size and time of compu-tation justified by the state of the art of classical computing hardware, is also compatiblewith this framework and most probably be the definition used in any real experimentaldemonstration.

This framework is very general, and it can be applied to any quantum algorithmsupposed to sample from a classical distribution, beyond random quantum circuits. Asan example, let us consider the optimization version of an NP-complete problem, such asMAXCUT on a ∆-regular graph with ∆ ≥ 3:

Example 2.1 (MAXCUT). suppose that Bob claims his quantum computer can achievea better value for an NP optimization problem, say MAXCUT, than Alice’s classicalcomputer. Recall that for a graph G = (V,E) with n vertices and maximum degree ∆,MAXCUT of the graph can be cast as finding the maximum over 0, 1n of the function

fG(x) =1

n∆

(i,j)∈E

(1 − δ(xi, xj)). (2)

Here, the normalization n∆ ensures that 0 ≤ f(x) ≤ 1. Thus, in this case, Bob canpropose the function fG to distinguish his distribution from Alice’s classical computer. Ifthe value for MAXCUT he can achieve is indeed at least ǫ better than what Alice canachieve, he wins the game. On the other hand, if classical methods can yield a better cutthan Bob’s quantum computer, then fG will not suffice and Bob will have to find anotherfunction to justify his claim of quantum advantage. Note that our choice of normalizationin Eq. (2) implies that an additive error ǫ on approximating the expectation value of fGimplies a multiplicative error of order ǫ for the cut’s value.

As exemplified above, for NP optimization problems, there is a clear choice for whichfunction Bob should propose and it can be computed in polynomial time.

At first sight, the possibility of the game always requiring an exponential number ofrounds seems a plausible outcome. However, there is an update rule for Alice’s distribu-tion that will lead to the game having at most O(nǫ−2) rounds, where n is the number ofqubits of Bob’s device, as we explain below. Note that we do not claim that this updaterule will always lead to Alice winning, only that it will converge after a finite number ofrounds. However, we will argue that this update rule is an appropriate choice againstrandom circuits and when competing with noisy quantum devices.

This update rule uses the connection between our certification game and mirror de-scent [Bub15], a method to learn probability distributions by approximating them by asequence of Gibbs distributions. In a nutshell, Alice can exploit each test function ftthat Bob provides to improve her guess of the distribution ν. The updates are of the

form µt+1 ∝ elog(µt)−ǫ4 ft . She can use her method of choice to sample from the Gibbs

distribution, such as rejection sampling. One can then show that at every round of the

4

Page 5: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

game, the guess of Alice gets closer to the ideal distribution by at least a finite amount,converging to µt being ǫ-close to the ideal quantum distribution ν in a finite number ofrounds. In fact, regardless of the distribution Bob is sampling from, if Alice chooses touse mirror descent to update her distribution, then it follows from standard propertiesof mirror descent that the game will end after at most 16nǫ−2 rounds. We explain thisin more detail in Section 4. The caveat is that knowledge from which Gibbs distributionAlice needs to sample from does not guarantee that she can do it efficiently.

Let us exemplify this again with MAXCUT:

Example 2.2 (MAXCUT continued-mirror descent and simulated annealing). In thesame setting as in Example 2.1, one can show that if Alice uses mirror descent to updateher probability distributions, her sequence of probability distributions µt is given by:

µt(x) =e

t

4ǫfG(x)

Zt

, Zt =∑

x∈0,1n

et

4ǫfG(x).

That is, her strategy will be akin to performing simulated annealing to try to obtain abetter value of MAXCUT. This is one of the most widely used methods for combinatorialoptimization [KGV83]. If one picks t large enough, µt is guaranteed to be sharply con-centrated around the maximum of fG, but the complexity of sampling from µt increasesaccordingly and at some point Alice won’t be able to sample from it anymore. We referto e.g. [GLS12, Chapter 28] and references therein for a discussion of the application ofsimulated annealing to combinatorial optimization. We can interpret the parameter t/ǫas the inverse temperature β. Thus, if Alice picks the mirror descent strategy, she wouldwin if a classical Monte Carlo algorithm has a performance comparable to that of Bob’squantum computer.

2.2 Main results

As we anticipated, this framework of quantum advantage certification allows us to provethree main results. Our first result is that for random circuits, we are only required toplay a number of rounds that is independent of the number of qubits n and Bob neverwins if Alice plays mirror descent and ǫ = Ω(log(n)−1).

Theorem 2.1 (Alice approximately learns ν after ǫ−2 rounds for random circuits). Letν be the probability distribution of the outcome of a random quantum circuit on n qubitsstemming from a 2−2n−1-approximate two design and ǫ = Ω(log(n)−1) be given. Supposethat Bob succeeds in providing functions f1, . . . , ft that can be computed in polynomialtime and distinguish ν from a sequence µ1, . . . , µt of distributions that Alice discloses,with t = O(ǫ−2). Then there is an algorithm that allows Alice to learn a distribution µt+1

exploiting the revealed information on ft that can be sampled from in polynomial time.Moreover, µt is ǫ close in total variation distance to ν.

The above result is connected to the entropy of the outcome distributions of randomcircuits, as it reads S(ν) = n−O(1) with high probability. Thus, up to a small constant,its value matches that of n random coins. This implies that mirror descent converges fast.While in the case of MAXCUT we needed, in general, n iterations to lower the tempera-ture and reach the ground state, because of the entropy bound for random circuits, theoutputs of a random circuit are well-approximated by a high-temperature Gibbs state.Thus, one way of rephrasing Theorem 2.1 is that if Bob can efficiently find distinguishingfunctions and they can be computed efficiently, then Alice can also find and sample froma high-temperature Gibbs state that is close to the ideal distribution.

An important corollary of this is that either the hardness conjectures of randomquantum circuits are not valid for distances ǫ = Ω(log(n)−1) or a complete certification

5

Page 6: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

strategy for Bob, providing a discrimination function for every guess of Alice, is impossiblewith polynomial resources.

Our second result concerns a connection between the hardness of fooling the XHOGproblem and distinguishing the output of a random circuit from the uniform. We referto Sec. 5 for a precise definition of the XHOG problem. There we also show that ifthe conjectures related to the hardness of fooling the XHOG problem are true, not evendistinguishing from the uniform distribution in polynomial time should be possible. Inthis case, Bob will have to resort to verification strategies that take super-polynomialtime to demonstrate a quantum advantage. More precisely, in Section 5 we prove thefollowing result.

Theorem 2.2 (Distingushing from uniform and HOG). Let f : 0, 1n → 0, 1 bea function that for some ǫ > 0 satisfies: Eν(f) − EU (f) ≥ ǫ, where ν is the outcomedistribution of a random quantum circuit stemming from a 2−2n−1-approximate two designin n qubits. Then there is an algorithm that samples from a distribution that fools XHOGup to ǫ using O(ǫ−4) evaluations of f in expectation.

One possible criticism of the above framework is that it might be in general hard todistinguish the outcome of a circuit stemming from a design from the uniform distribution.However, we show in Appendix C that in case Bob is sampling from a stabilizer circuit,then Alice can easily fool XHOG.

All the results above concern the outcome distribution of the ideal circuit. In Sec. 6we extend our results to the approximate simulability of the outcome distribution ofnoisy devices. We show that, under doubly-stochastic noise, the number of rounds of theverification game when Alice uses mirror descent decreases exponentially with the depthof the circuit Bob is implementing. As we believe these results are interesting beyondquantum advantage proposals and apply to the broader topic of classical simulability ofnoisy circuits, we state them in more general terms. Below we state a specialized versionof our main result regarding the complexity of approximating the statistics of outcomesof noisy circuits, Theorem 6.1:

Theorem 2.3. Let ν be the outcome distribution of a noisy quantum circuit on n qubitsof depth D affected by local depolarizing noise with rate p after each gate, measured inthe computational basis. Given functions f1, . . . , fk : 0, 1n → [0, 1], mirror descent willconverge to a distribution µt satisfying:

|Eν(fi)− Eµt(fi)| ≤ ǫ

for all 1 ≤ i ≤ k in at most T = O(ǫ−2(1−p)2D+2n) iterations. Moreover, we can samplefrom µt by evaluating f1, . . . , fk at most

exp

(

4(1− p)2D+2n

ǫ

)

times.

As discussed in the recent [FGP20], which we discuss in more detail shortly, thisrestrains the power of noisy quantum computers to demonstrate a significant advantageversus classical methods for more structured problems. Let us exemplify this with thenoisy MAXCUT example:

Example 2.3 (MAXCUT continued- noisy circuits). Let us exemplify the consequencesof Theorem 2.3 to approximating MAXCUT on a noisy quantum device. Suppose thatBob’s device suffers from local depolarizing noise with rate p and consists of a circuit of

6

Page 7: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

depth D. In this scenario, Alice will be able to obtain an expected value of MAXCUT thatis ǫ close to Bob’s by sampling from the distribution µt given by:

µt(x) =eβfG(x)

Z, Z =

x∈0,1n

eβfG(x)

with β = ǫ−1(1 − p)2D+2n. That is, the noise decreases the inverse temperature β wehave to go when performing classical simulated annealing to obtain comparable results.

As is discussed in more detail in [FGP20], results like the one above can be used torigorously establish maximal depths before noisy quantum devices are not outperformedby polynomial time classical algorithms. But the main message of the example abovein our context of verification is that if there are clear candidate functions to distinguishthe output of the noisy quantum circuit, then the noise will make it easier for Alice tosimulate the output of the device, as one would expect. We refer to Section 6 for aderivation of this bound and a more detailed discussion of its consequences.

3 Preliminaries

We will now introduce some basic definitions and notation together with the concepts ofmirror descent and rejection sampling, which are relevant to our work.

3.1 Notation

Probability distributions on binary strings: we define F = (0, 1n)[0,1]

tobe the set of functions f : 0, 1n → [0, 1].

Given two probability distributions µ, ν on 0, 1n, we define their total variationdistance ‖ν − µ‖TV as:

‖ν − µ‖TV =1

2

x∈0,1n

|µ(x) − ν(x)| .

The total variation is widely accepted as one of the most natural and operationally rel-evant measures of distinguishability for two probability distributions. One of the reasonsfor that is its dual formulation. One can show that:

‖ν − µ‖TV = 2 supf∈F,‖f‖∞≤1

(Eµ(f)− Eν(f)), (3)

where

‖f‖∞ = supx∈0,1n

f(x).

Moreover, defining S = x ∈ 0, 1n : µ(x) ≥ ν(x) and letting χS be the indicatorfunction of S, it is easy to see that:

‖ν − µ‖TV = 2(Eµ(χS)− Eν(χS)).

That is, we can restrict to indicator functions in Eq. (3). This indicator function providesthe optimal guessing strategy between distributions µ, ν connecting it with the totalvariation distance [CT12].

The characterization given in Eq. (3) can also be used to certify that the total variationbetween two probability distributions is at least ǫ. To see why, given some function f

7

Page 8: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

with ‖f‖∞ ≤ 1 and samples X1, . . . , Xs from µ and Y1, . . . , Ys from ν, it follows fromHoeffding’s inequality that:

s−1s∑

i=1

f(Xi)− Eµ(f)

≤ǫ

2,

s−1s∑

i=1

f(Yi)− Eν(f)

≤ǫ

2(4)

with probability at least 1− δ as long as

s = O(ǫ−2 log(δ−1)). (5)

Thus, if

s−1s∑

i=1

(f(Yi)− f(Xi))

≥ 2ǫ,

then a simple application of the triangle inequality yields together with Eq. (4) that

|Eµ(f)− Eν(f))| ≥ ǫ,

with probability at least 1− δ. This, in turn, implies that ‖µ−ν‖TV ≥ ǫ with probabilityat least 1 − δ. Thus, we conclude from Eq. (5) and the discussion above that as long asǫ−1 log(δ−1) = O(poly(n)), polynomially many samples and evaluations of the function fare sufficient to certify that two distributions are at least at a certain distance ǫ in totalvariation. Of course, this in no sense discards the possibility that finding the distinguish-ing function f itself or evaluating it may not be possible in polynomial time. This allowsus to estimate the number of samples we need to provide at each round of the game toensure that the probability of a deviation greater than ǫ from the target is upper boundedby 1−O(δ) for some δ. As at each round t we have to estimate t expectation values upto ǫ, obtaining O(ǫ−2 log(tδ−t)) samples for each round ensures that the probability oneof them deviates by more than ǫ is at most δt. By a union bound, the probability thatthere was a deviation after T rounds is at most

T∑

t=1

δt ≤δ

1− δ= O(δ)

for δ ≤ 13 . Thus, letting the number of samples per round grow like t log(tδ−1) is enough

to ensure that the probability of an error occurring at some point remains of order δ.

Distinguishability measures for quantum states and unitary designs:we are also going to need other distinguishability measures for distributions and quantumstates. We will introduce them only for quantum states and note that the correspondingclassical definition is obtained by considering the classical probability distribution as adiagonal quantum state. For two quantum states ρ, σ ∈ M2n we define their relativeentropy to be:

S(ρ||σ) = tr (ρ (log ρ− log σ))

if kern ρ ⊂ kern σ and +∞ otherwise. Moreover, we define the α−Renyi entropies Sα forα > 1 to be given by:

Sα(ρ) = −α

α− 1log (tr (ρα)) .

8

Page 9: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

and the von Neumann entropy to be S1(ρ) = S(ρ) = −tr (ρ log(ρ)). Note that we have:

n ≥ S(ρ) ≥ Sα(ρ).

Let us also set our notation and terminology for random quantum circuits. Given

a distribution τ on the unitary group on n qubits, U(2n), and t ∈ N, we define G(t)τ :

M2n → M2n to be the quantum channel:

G(t)τ (X) =

U(2n)

U⊗tX(

U †)⊗t

dτ.

G(t)τ is then said to be an (ǫ)-approximate t-design [AE07] if

‖G(t)τ − G(t)

µG‖⋄ ≤ ǫ,

where ‖ · ‖⋄ is the diamond norm and µG is the Haar measure on the unitary group.Moreover, given C distributed according to τ , we will always denote by ν the probabilitymeasure we obtain by measuring C |0〉 in the computational basis, i.e.

ν(x) = tr(

|x〉〈x|C|0〉〈0|⊗nC†)

for x ∈ 0, 1n.

3.2 Mirror descent

Mirror descent [TRW05, Bub15] is an optimization and learning algorithm to approximateprobability distributions efficiently and in a structured way through a series of Gibbsprobability measures. It allows us to formally connect the problem of distinguishingprobability distributions and learning them. That is, given some target distribution νon 0, 1n we wish to learn, say the output distribution of a given random quantumcircuit, mirror descent is an iterative procedure that proposes a sequence of µ0, . . . , µT

of guesses for ν. Furthermore, the initial distribution µ0 is the uniform distribution U .The algorithm requires us to find functions f1, . . . , fT : 0, 1n → [0, 1] that allow us todistinguish µt from ν, i.e.

Eµt(ft)− Eν(ft) ≥ ǫ (6)

for some given distinguishability parameter ǫ > 0. One can now appreciate the directconnection between mirror descent and our verification game. Of course, it is a priorinot clear how to find such functions in a traditional mirror descent application. In thecertification game, this problem is overcome by having the responsibility to provide f onBob’s side. Also note that if no function ft exists that satisfies (6), then ‖ν − µt‖TV ≤ ǫby the dual formulation of the total variation distance in eq. (3).

As outlined in 1, mirror descent works by updating the probability measure as:

µt+1 = exp

(

−ǫ

4

t∑

i=0

fi

)

/Zt+1, (7)

where

Zt+1 =∑

x∈0,1n

exp

(

−ǫ

4

t∑

i=0

fi(x)

)

is the partition function. As we update the distributions, the candidate distributionsµt become closer and closer to the target distribution, as made precise by the followinglemma:

9

Page 10: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

Algorithm 1 Mirror descent for learning probability distributions.

1: function Mirror descent(T, ǫ)2: Set µ0 = U ⊲ initialize to the uniform distribution3: for t = 1, . . . , T = ⌈16S (ν‖U) ǫ−2⌉ do

4: Demand function ft such that Eµt(ft)− Eν(ft) ≥ ǫ

5: if Given ft then

6: Set µt+1(x) = exp(− ǫ4

t∑

i=0fi(x))/Zt+1. ⊲ Update the guess.

7: end if

8: if no such function exists then9: Return µt

10: break loop

11: end if

12: end for

13: Return µT and exit function ⊲ Current guess is ǫ indistinguishable from ν14: end function

Lemma 3.1. The distributions µt of the algorithm 1 satisfy:

S (ν‖µt) ≤ −tǫ2

16+ S (ν‖U) , (8)

where U is the uniform distribution.

Eq. (8) is a standard property of mirror descent [Bub15].Exploiting the direct connection between the verification protocol and mirror descent,

we can directly use lemma 3.1 to bound the number of rounds of the game of in terms ofS (ν‖U). We then immediately obtain:

Theorem 3.1. The output of algorithm 1 satisfies:

‖µt − ν‖TV ≤ ǫ (9)

after at most T ≤ 16ǫ−2S (ν‖U) iterations.

Proof. Using Pinsker’s inequality and Eq. (8) it is easy to prove the relation

‖µt − ν‖TV ≤√

2S (ν‖µt) ≤

2

(

S (ν‖U)− tǫ2

16

)

, (10)

This together with the positivity of the total variation distance provides the claim.

Note that Eq. (10) ensures that we make constant progress in total variation at eachiteration. Theorem 3.1 implies that if Bob provides a sequence of functions f1, . . . , fT+1

that allow distinguishing ν from the sequence µ0, . . . , µT of at most O(ǫ−2S (ν‖U)) dis-tributions up to an error ǫ, then we can also find a distribution that is ǫ close to it intotal variation distance. Furthermore, as we will show later, it is possible to use thisconnection to the relative entropy to quantify the effect of noise on the complexity oflearning the distribution.

10

Page 11: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

3.3 Rejection sampling

Let us now show how to generate samples from µt and the underlying complexity. Wewill use the standard technique of rejection sampling described in Algorithm 3.3. Werefer to [Appendix B.5][LP17] for a brief review of its properties. In rejection sampling wesample indirectly from a target distribution µt by first generating a sample x with an easyto sample distribution γ(x) and accepting the sample with probability µt(x)/(Mγ(x)),where M is a constant such that the ratio is bounded by 1. It is a standard fact thatrejection sampling will output a sample after M runs in expectation, as the probabilityof rejection follows a geometric distribution with parameter M−1. In the case of Gibbsdistributions µt = exp(−Ht)/Zt, where

Ht(x) =ǫ

4

t∑

i=1

fi(x),

a common choice for γ is the uniform distribution and Mt =2n

Zt, where Zt is once again

the partition function. Note that for this choice of Mt, we have that:

µt(x)

MtU(x)= e−Ht(x) ≤ 1,

as Ht(x) ≥ 0 for all x ∈ 0, 1n by our choice of ft. In particular, note that with thischoice, we never have to compute the partition function Zt to run rejection sampling,only Ht(x). Thus, we conclude that the complexity of running one round of rejectionsampling is the same as computing Ht(x). Let us now estimate how many rounds arerequired in expectation before we accept a sample:

Lemma 3.2 (Sampling from µt). Let µt be the guess at iteration t of Algorithm 1.Running rejection sampling with the uniform distribution as γ and Mt = 2n

Ztreturns a

sample from µt after at most eǫ

4t trials and evaluations of Ht, in expectation.

Proof. Note that by our previous discussion the expected number of trials isMt =2n

Zt. By

construction, ft are functions with image [0, 1]. Thus, it follows from a triangle inequalitythat:

‖Ht‖∞ ≤ǫ

4

t∑

i=1

‖fi‖∞ ≤tǫ

4.

This implies that

Zt =∑

x∈0,1n

exp(−Ht(x)) ≥ 2ne−ǫ

4t.

Algorithm 2 Rejection sampling.

Require: ability to generate samples from distribution γ on 0, 1n, distribution µt on

0, 1n, constant M such that µt(x)Mγ(x) ≤ 1, ability to compute µt(x)

Mγ(x) and samples from

U([0, 1]).1: function Rejection sampling(M)2: Sample u distributed according to U([0, 1]) and x distributed acoording to γ.

3: if u ≤ µt(x)Mγ(x) then

4: Output x5: end if

6: end function

11

Page 12: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

We conclude from the last inequality that

Mt =2n

Zt

≤ eǫ

4t

which yields the claim.

We see that as long as ǫt = O(log(n)), then we can sample from µt in a polynomialexpected number of trials and evaluations of Ht.

In practice, rejection sampling is not necessarily the most efficient way of simulatingprobability distributions and other techniques to sample from a Gibbs distribution suchas Glauber dynamics or simulated annealing [LP17] perform better. However, rejectionsampling allows for a simple analytical analysis, which is more challenging for more refinedtechniques.

4 Random quantum circuits

Let us now discuss the implications to the verification of quantum advantage propos-als based on sampling the output distribution of random circuits. The key technicalassumption behind various state of the art classical hardness proofs for quantum ad-vantage proposals is the property that the underlying ensemble is an approximate twodesign [HBVSE18]. Thus, we will also depart from this assumption. We then have:

Lemma 4.1. Let ν be the probability distribution of the outcome of a random quan-tum circuit on n qubits stemming from a 2−2n−1-approximate two design. Then, withprobability at least 1− δ, we have:

S(ν) ≥ n+ log(δ)− log(3).

Proof. We refer to Prop. A.1 in Appendix A for a proof.

Similar statements were shown in [HKEG19, AA14a]. From this, we have:

Theorem 4.1 (Disitinguishing output distributions and classical simulability). Let νbe the probability distribution of the outcome of a random quantum circuit on n qubitsstemming from a 2−2n−1-approximate two design and ǫ = Ω(log(n)−1) be given. Supposethat we can distinguish ν from a sequence µ1, . . . , µt of distributions that can be sampledfrom in polynomial time. Moreover, we can distinguish them by functions f1, . . . , fT thatcan be evaluated in polynomial time. That is:

∀1 ≤ t ≤ T : Eµt(ft)− Eν(ft) ≥ ǫ. (11)

with ft computable in polynomial time. Then, with probability at least 1− δ, we can finda distribution µt satisfying:

‖ν − µt‖TV =

2

(

log(3)− log(δ)− tǫ2

16

)

(12)

and sample from it in time O(ec

ǫ poly(n)) = O(poly(n)).

Proof. As we show in Prop. A.1, we have that with probability at least 1− δ

S(ν) ≥ n− log(2 + ǫ) + log(δ).

12

Page 13: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

Conditioned on the event above, it follows from Cor. 3.1 that mirror descent outputs adistribution satisfying Eq. (12) after at most

16ǫ−2 (log(3)− log(δ))

iterations, or equivalently after that many game rounds. Now, at each iteration t ofmirror descent, we need a function satisfying Eq. (11). Moreover, we have that µt ∝

exp

(

−ǫ/4∑

i

fi

)

. Thus, if all the fi can be computed in polynomial time, then it follows

from Lemma 3.2 that we can also sample from µt using rejection sampling in polynomialtime. This is because Lemma 3.2 implies that we need at most a polynomial number

exp

(

4(log(3)− log(δ))

ǫ

)

of rejection sampling rounds, in expectation. As each round of rejection sampling re-quires us to evaluate the functions fi once and we suppose that they can be evaluated inpolynomial time, this gives the claim.

Therefore, if Bob provides for every proposed distribution µt of Alice a polytimecomputable function ft that distinguishes it from ν, after at most a constant number ofrounds Alice will be sampling efficiently from an approximate distribution. It is interestingto point out that the certification game and the sampling of Alice remains efficient, evenif we relax the condition of Lemma 4.1 to S(ν) ≥ n−O(log(n)) and request ǫ to decreasewith the size n of the quantum device with scaling ǫ = O(log(n)−1).

A direct consequence of our result is that if the hardness conjecture of random quan-tum circuits is true, then Bob must fail to provide a certification function that is efficientlycomputable at some stage of the certification game. A natural question would then beto ask at which stage Bob will fail to provide such a function. In the following section,we will prove that if the XHOG conjecture [AG19] is correct, Bob must fail at the firstround of the game, i.e., even distinguishing the output distributions from the uniformdistribution cannot be done in polynomial time.

We note that these results have important differences from the results in [AA14a,Appendix 11]. There the authors show the existence of a high min-entropy distributionthat can be sampled from classically and is indistinguishable from the random quantumcircuit by classical circuits of polynomial size. This is because in our case we have theguarantee of a good approximation in total variation distance, i.e. the distributions areindistinguishable under any function after a couple of iterations. Moreover, we have anefficient way of finding the distribution and sampling from it if the distinguishing functionscan be computed efficiently, while, to the best of our knowledge, the aforementioned resultshould be taken as an existence result. Finally, our framework allows us to work withthe Shannon entropy instead of the min-entropy. The min-entropy is notoriously moredifficult to bound and always smaller than the Shannon entropy.

5 Distinguishing from the uniform distribution

To the best of our knowledge, the state of the art approach for the verification of quan-tum advantage proposals based on random circuit sampling is the linear cross-entropyheavy output generation problem (XHOG) [AG19], which is closely related to the linearcross-entropy benchmarking (linear XEB) fidelity FXEB [NRK+18, BIS+18, AAB+19].The XHOG refers to the problem of, given some circuit C, generating distinct samples

13

Page 14: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

z1, . . . , zk such that:

Ei

[

|〈zi|C|0n〉|

2]

≥ b/2n (13)

for some b > 1 with probability at least 12 + 1

2b and k satisfying:

k ≥1

((2s− 1)b− 1)(b− 1). (14)

Note that, given the samples z1, . . . , zk, verifying that they indeed satisfy eq. (13) requiresus to compute the probability of the outcomes under the ideal circuit. In turn, the linearcross-entropy fidelity for a distribution µ, FXEB(µ), as defined in [NRK+18, BIS+18,AAB+19], is given by:

FXEB(µ) = 2nEν (µ)− 1.

Here we see the probability distribution µ as a function µ : 0, 1n → R and, as usual, νis the output probability distribution of the ideal random circuit. A simple manipulationthen shows that samples zi from µ satisfy

Ei

[

|〈zi|C|0n〉|2

]

≥1 + FXEB(µ)

2n.

In [AG19], the authors relate the complexity of computing outcome probabilities of ran-dom quantum circuits to the XHOG problem. More precisely, the authors start by assum-ing that there is no polynomial-time classical algorithm that takes as input a (random)quantum circuit C and produces an estimate p of p0 = P[C outputs 0] such that

E

[

(

p0 − 2−n)2]

= E

[

(p0 − p)2]

+Ω(

2−3n)

.

where the expectations are taken over circuits C as well as the algorithm’s internal ran-domness. They then show that this conjecture implies that there is no polynomial timealgorithm that solves the XHOG problem. As the verification of XHOG requires us toestimate the outcome probabilities, this path to proving and verifying the hardness of thesampling task also implies that XHOG is not verifiable in polynomial time.

But the hardness of XHOG imposes barriers to even more basic verification tasks. Aswe will see now, the hardness of XHOG would imply that it is impossible to efficientlydistinguish the output from the circuit from the uniform distributions. In turn, efficientdistinguishability implies a polynomial time algorithm to spoof XHOG. Therefore, Bobwill have to resort to verification strategies that take superpolynomial time to demonstratea quantum advantage in our game. The proof of the statement will once again rely onthe fact that the output distribution is essentially flat. We will start by showing thatany distribution with large 2−Renyi entropy cannot have small sets of large mass in thefollowing sense:

Lemma 5.1. Let ν be a probability distribution on 0, 1n such that S2(ν) ≥ n− log(c)for some constant c > 0. Then, for any ǫ > 0 and subset L ⊂ 0, 1n we have that

ν(L) =∑

x∈L

ν(x) ≥ ǫ

implies that

|L| ≥ ǫ2c−12n.

14

Page 15: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

Proof. Note that the condition S2(ν) ≥ n− log(c) is equivalent to

x∈0,1n

ν(x)2 ≤ c2−n. (15)

Moroever, we have

x∈0,1n

ν(x)2 ≥∑

x∈L

ν(x)2 ≥ǫ2

|L|. (16)

To see the last inequality, note that, by the concavity of the function x 7→ x2,

1

|L|

x∈L

ν(x)2 ≥

(

1

|L|

x∈L

ν(x)

)2

≥ǫ2

|L|2.

Combining (16) with (15) we conclude that:

ǫ2

|L|≤ c2−n,

which yields the claim after rearranging the terms.

It then immediately follows that:

Lemma 5.2. Let f : 0, 1n → 0, 1 be a function that for some ǫ > 0 satisfies:

Eν(f)− EU (f) ≥ ǫ, (17)

where ν is the outcome distribution of a random quantum circuit stemming from a 2−2n−1-approximate two design on n qubits. Let

L = x ∈ 0, 1n : f(x) = 1.

Then, with probability at least 1− δ,

|L| = Ω(

ǫ2δ2n)

. (18)

and

1

|L|ν(L) ≥

1

2n+

ǫ

|L|≥

1 + ǫ

2n. (19)

Proof. First note that Eq. (17) is equivalent to

ν(L) ≥|L|

2n+ ǫ

and, in particular, ν(L) ≥ ǫ. Moreover, Eq. (19) readily follows by dividing the equationabove by |L|. As we show in Prop. A.1 in Eq. (24) that with probability at least 1− δ

S2(ν) ≥ n− log(3) + log(δ). (20)

Conditioned on Eq. (20), it follows from Lemma 5.1 that

|L| ≥ cǫ2δ2n (21)

for some constant c > 0, which yields Eq. (18).

15

Page 16: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

We restricted the result above to functions with binary outputs to simplify the argu-ments, but we show in Appendix B that this can be done without loss of generality. Thatis, given a function f that distinguished the distributions for ǫ and range [0, 1], therealways exists some f ′ that is binary and has the same properties and distinguished the

distribution up to ǫ2

4 .If the function f in Lemma 5.2 can be computed in polynomial time, then we can use

it to fool XHOG in polynomial time:

Theorem 5.1 (From distinguishing to fooling XHOG). Let f as in Lemma 5.2. More-over, let U(L) be the uniform distribution on L. Then we can sample from U(L) byevaluating f O(ǫ−4) many times, in expectation. Moreover, samples from U(L) violateHOG up to ǫ.

Proof. Let us start with the statement that samples from U(L) violate XHOG up to atleast ǫ. To see this, note that:

Eν (U(L)) =∑

x∈L

ν(x)

|L|≥

1 + ǫ

2n

by Eq. (19). To sample from U(L), we can once again resort to a variation of rejectionsampling. We sample a point x1 ∈ 0, 1n from the uniform distribution and computef(x1). If f(x1) = 1, we output x1. If not, we rerun this procedure with a new samplex2. It is easy to see that when we accept, xi is uniformly distributed on L. Moreover,the probability of accepting is L

2n . By Eq. (21), it then follows that the probability ofaccepting is at least Ω(ǫ2), from which we obtain that the expected number of rounds isO(ǫ−4). As for each round we have to evaluate f once, the claim follows.

It follows that if we can efficiently distinguish the output from the uniform distribu-tion, then we can fool XHOG. In particular, it would follow from the conjectures of [AG19]that it is not possible to distinguish the output of random circuits from the uniform dis-tribution for ǫ at least inverse polynomial in n in polynomial time if XHOG cannot besolved in polynomial time.

Note that the results of this section only required the property that the underlyingrandom circuit ensemble is a two design. However, it is well-known that random Cliffordunitaries also are two designs and can still be simulated efficiently classically. It is thennatural to ask if finding a function distinguishing the output of a random Clifford fromthe uniform distribution can be found in polynomial time. As we show in Appendix C,for random Cliffords, the function can be found and be computed in polynomial time.

6 Noisy devices

Most of the current implementations of quantum circuits have high levels of noise. Inprinciple, highly-correlated noise can make simulating the quantum device classicallyeven more complex [KBG+11]. However, in other contexts, as demonstrated for bosonsampling [QBQGP20], noise can render the simulation significantly less complex. Here, weshow that in the scenario of doubly stochastic Markovian noise, i.e., quantum channelsthat map the maximally mixed state to itself, the noise diminishes the complexity ofapproximating the underlying distribution being sampled from. To see what we mean, letus go back the game described in Section 2. There we considered Bob to have a quantumadvantage and win the game if he could find functions f1, . . . , ft whose expectation valuesunder his distribution Alice cannot reproduce classically. However, suppose now thatBob’s device is affected by noise, say a global depolarizing channel with parameter 0 ≤

16

Page 17: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

p ≤ 1. One would then expect that as p → 1, it should be easier for Alice to reproducethe statistics of Bob’s device and win the game. As we show below, if Alice once againuses mirror descent to find her candidate distributions, then the algorithm will convergefaster to a distribution that reproduces the statistics of Bob’s distribution as the levelof noise increases. For example, for MAXCUT, this will also make it easier for her tosample from the distribution mirror descent proposes, as a smaller number of iterationsimplies not having to go down to lower temperatures.

6.1 Strong data processing inequalities and mirror descent

Let us first recall the following definition:

Definition 6.1 (Strong data processing inequality). A doubly stochastic quantum channelT : Md → Md (i.e. a CPTP map satisfying T (I) = T ∗(I) = I) is said to satisfy a strongdata processing inequality with contraction α > 0 w.r.t. I

dif for all states ρ we have:

S

(

T (ρ)‖I

d

)

≤ (1− α)S

(

ρ‖I

d

)

.

Strong data processing inequalities for doubly stochastic can be derived using theframework of hypercontractivity and logarithmic Sobolev inequalities [KT13, MHFW16b,MHFW16a, HRF20, BDR20] and are known explicitly in some cases. See also [BSW19] foranother approach. As shown by other works [ABOIN96, BOGH13, MHRW15, FGP20],strong data processing can be used to quantify how useful a noisy quantum device is andfor how long it can sustain interesting computations.

Let us illustrate the strong data processing inequality with an example. Let Φp :M2 → M2 be the depolarizing channel on one qubit with depolarizing parameter p, i.e.

Φp(ρ) = (1− p)ρ+ pI

2.

The authors of [MHFW16b] show that:

S

(

Φ⊗np (ρ)‖

I

2n

)

≤(

1− 2p+ p2)

S

(

ρ‖I

2n

)

and similar results are available for other relevant noise models. In particular, optimalinequalities have been derived in [KT13, MHFW16a] for tensor products of single qubit,doubly stochastic channels.

Suppose now that the noisy circuit of interest consists of n qubits initialized to |0〉⊗n

,D layers of unitaries U1, U2, . . . , UD and a measurement in the computational basis. How-ever, due to imperfections in the implementation, the state initialization, the measure-ments and the unitaries are noisy. We will model this by assuming that every layer ofunitaries is proceeded by a layer of doubly stochastic quantum channel Φ that satisfies astrong data processing inequality α. Moreover, we will assume that the measurement isalso affected by an extra noisy channel Φ. We only make these assumptions to simplifythe notation and argument, but it is easy to adapt the argument to different channels atdifferent times and different noise rates. Under the assumptions above, the probabilitydistribution ν describing the outcomes of the noisy device is given by:

ν(i) = tr(

|i〉〈i|T (|0〉〈0|⊗n))

= tr(

|i〉〈i|(M Φ UD Φ . . . U1 Φ(|0〉〈0|⊗n))

, (22)

where Ui is the channel given by conjugations with Ui and M is the q.c. channel

M(ρ) =

2n−1∑

i=0

tr (ρ|i〉〈i|) |i〉〈i|

17

Page 18: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

and

T (|0〉〈0|⊗n) = (M Φ UD Φ . . . U1 Φ(|0〉〈0|⊗n).

We then have:

Theorem 6.1. Let ν be the distribution defined in Eq. (22) and assume Φ satisfies astrong data processing inequality with parameter α. Given functions f1, . . . , fk : 0, 1n →[0, 1], mirror descent will converge to a distribution µt satisfying:

|Eν(fi)− Eµt(fi)| ≤ ǫ

for all 1 ≤ i ≤ k in at most T = O(ǫ−2(1−α)D+1n) iterations. Moreover, we can samplefrom µt by evaluating f1, . . . , fk at most

exp

(

4(1− α)D+1n

ǫ

)

times.

Proof. Mirror descent will converge to a distribution with the desired properties after16ǫ−2S (ν‖U) iterations. Moreover, by Lemma 3.2, the complexity of sampling from µ isbounded by

exp

(

4S (ν‖U)

ǫ

)

evaluations of the functions fi. Thus, the statement follows if we can bound the relativeentropy of the outcome. Note that:

S (ν‖U) = S

(

T (|0〉〈0|⊗n)‖I

2n

)

,

as the maximally mixed state gives rise to the uniform distribution when measured in thecomputational basis. By the data processing inequality:

S

(

T (|0〉〈0|⊗n)‖I

2n

)

≤ S

(

Φ UD Φ . . . U1 Φ(|0〉〈0|⊗n)‖

I

2n

)

and by our assumption on Φ:

S

(

Φ UD Φ . . . U1 Φ(|0〉〈0|⊗n)‖

I

2n

)

≤ (1 − α)S

(

UD Φ . . . U1 Φ(|0〉〈0|⊗n)‖

I

2n

)

.

The relative entropy is unitarily invariant, thus:

(1 − α)S

(

UD Φ . . . U1 Φ(|0〉〈0|⊗n)‖

I

2n

)

= (1− α)S

(

Φ . . . U1 Φ(|0〉〈0|⊗n)‖

I

2n

)

.

Applying the chain of arguments above another D times we conclude that:

S

(

T (|0〉〈0|⊗n)‖I

2n

)

≤ (1− α)D+1S

(

|0〉〈0|⊗n‖I

2n

)

= (1 − α)D+1n,

which yields the claim.

18

Page 19: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

Thus, we see that the complexity of approximate sampling from the output of noisycircuits decreases exponentially with the noise level and depth. For instance, for circuitswith local depolarizing noise with parameter p and depth D, the theorem above gives acomplexity of:

exp

(

4(1− p)2D+2n

ǫ

)

. (23)

evaluations of the distinguishing functions. Thus, we see that our framework has thedesirable feature that it becomes easier for Alice to mock Bob’s device as the noiseincreases. Unfortunately, the scaling with n in the bound above is undesirable for high-entropy distributions. Thus, we plan to derive more specialized bounds in upcomingwork.

7 Acknowledgments

DSF was supported by VILLUM FONDEN via the QMATH Centre of Excellence underGrant No. 10059. RGP was supported by the Quantum Computing and SimulationHub, an EPSRC-funded project, part of the UK National Quantum Technologies Pro-gramme. We thank Anthony Leverrier and Juani Bermejo-Vega for helpful commentsand discussions.

References

[AA14a] Scott Aaronson and Alex Arkhipov. Boson sampling is far from uniform.Quantum Info. Comput., 14(15–16):1383–1423, November 2014.

[AA14b] Scott Aaronson and Alex Arkhipov. The computational complexity of linearoptics. In Research in Optical Sciences. OSA, 2014.

[AAB+19] Frank Arute, Kunal Arya, Ryan Babbush, Dave Bacon, Joseph C Bardin,Rami Barends, Rupak Biswas, Sergio Boixo, Fernando G S L Brandao,David A Buell, Brian Burkett, Yu Chen, Zijun Chen, Ben Chiaro, RobertoCollins, William Courtney, Andrew Dunsworth, Edward Farhi, BrooksFoxen, Austin Fowler, Craig Gidney, Marissa Giustina, Rob Graff, KeithGuerin, Steve Habegger, Matthew P Harrigan, Michael J Hartmann, AlanHo, Markus Hoffmann, Trent Huang, Travis S Humble, Sergei V Isakov,Evan Jeffrey, Zhang Jiang, Dvir Kafri, Kostyantyn Kechedzhi, Julian Kelly,Paul V Klimov, Sergey Knysh, Alexander Korotkov, Fedor Kostritsa, DavidLandhuis, Mike Lindmark, Erik Lucero, Dmitry Lyakh, Salvatore Mandra,Jarrod R McClean, Matthew McEwen, Anthony Megrant, Xiao Mi, KristelMichielsen, Masoud Mohseni, Josh Mutus, Ofer Naaman, Matthew Neeley,Charles Neill, Murphy Yuezhen Niu, Eric Ostby, Andre Petukhov, John CPlatt, Chris Quintana, Eleanor G Rieffel, Pedram Roushan, Nicholas CRubin, Daniel Sank, Kevin J Satzinger, Vadim Smelyanskiy, Kevin J Sung,Matthew D Trevithick, Amit Vainsencher, Benjamin Villalonga, TheodoreWhite, Z Jamie Yao, Ping Yeh, Adam Zalcman, Hartmut Neven, andJohn M Martinis. Quantum supremacy using a programmable supercon-ducting processor. Nature, 574(7779):505–510, 2019.

[ABOIN96] Dorit Aharonov, Michael Ben-Or, Russell Impagliazzo, and Noam Nisan.Limitations of noisy reversible computation. arXiv preprint quant-ph/9611028, 1996.

19

Page 20: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

[AC17] Scott Aaronson and Lijie Chen. Complexity-theoretic foundations of quan-tum supremacy experiments. In Proceedings of the 32nd ComputationalComplexity Conference, 2017.

[AE07] Andris Ambainis and Joseph Emerson. Quantum t-designs: t-wise indepen-dence in the quantum world. In Twenty-Second Annual IEEE Conferenceon Computational Complexity 07). IEEE, jun 2007.

[AG04] Scott Aaronson and Daniel Gottesman. Improved simulation of stabilizercircuits. Physical Review A, 70(5), Nov 2004.

[AG19] Scott Aaronson and Sam Gunn. On the classical hardness of spoofing linearcross-entropy benchmarking, 2019. arXiv:1910.12085.

[BDR20] Salman Beigi, Nilanjana Datta, and Cambyse Rouze. Quantum reversehypercontractivity: Its tensorization and application to strong converses.Communications in Mathematical Physics, 376(2):753–794, may 2020.

[BFNV19] Adam Bouland, Bill Fefferman, Chinmay Nirkhe, and Umesh Vazirani. Onthe complexity and verification of quantum random circuit sampling. NaturePhysics, 15(2):159, 2019.

[BIS+18] Sergio Boixo, Sergei V. Isakov, Vadim N. Smelyanskiy, Ryan Babbush, NanDing, Zhang Jiang, Michael J. Bremner, John M. Martinis, and HartmutNeven. Characterizing quantum supremacy in near-term devices. NaturePhysics, 14(6):595–600, apr 2018.

[BJS11] Michael J Bremner, Richard Jozsa, and Dan J Shepherd. Classical simu-lation of commuting quantum computations implies collapse of the poly-nomial hierarchy. In Proceedings of the Royal Society of London A: Math-ematical, Physical and Engineering Sciences, volume 467, pages 459–472.The Royal Society, 2011.

[BKVV20] Zvika Brakerski, Venkata Koppula, Umesh Vazirani, and Thomas Vidick.Simpler Proofs of Quantumness, 2020. arXiv:2005.04826v1.

[BMS17] Michael J. Bremner, Ashley Montanaro, and Dan J. Shepherd. Achievingquantum supremacy with sparse and noisy commuting quantum computa-tions. Quantum, 1:8, apr 2017.

[BOGH13] Michael Ben-Or, Daniel Gottesman, and Avinatan Hassidim. Quantumrefrigerator. arXiv preprint arXiv:1301.1995, 2013.

[BRI+14] Sergio Boixo, Troels F. Rønnow, Sergei V. Isakov, Zhihui Wang, DavidWecker, Daniel A. Lidar, John M. Martinis, and Matthias Troyer. Evidencefor quantum annealing with more than one hundred qubits. Nature Physics,10(3):218–224, feb 2014.

[BSW19] Mario Berta, David Sutter, and Michael Walter. Quantum Brascamp-LiebDualities, 2019. arXiv:1909.02383v2.

[Bub15] Sebastien Bubeck. Convex Optimization: Algorithms and Complexity.Foundations and Trends® in Machine Learning, 8(3-4):231–357, 2015.

[CMS+14] Jacques Carolan, Jasmin D. A. Meinecke, Peter J. Shadbolt, Nicholas J.Russell, Nur Ismail, Kerstin Worhoff, Terry Rudolph, Mark G. Thompson,Jeremy L. Brien, Jonathan C. F. Matthews, and Anthony Laing. On theexperimental verification of quantum complexity in linear optics. NaturePhotonics, 8(8):621–626, jul 2014.

[CT12] Thomas M Cover and Joy A Thomas. Elements of information theory. JohnWiley & Sons, 2012.

20

Page 21: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

[DCEL09] Christoph Dankert, Richard Cleve, Joseph Emerson, and Etera Livine. Ex-act and approximate unitary 2-designs and their application to fidelity es-timation. Physical Review A, 80(1), jul 2009.

[DLT02] D.P. DiVincenzo, D.W. Leung, and B.M. Terhal. Quantum data hiding.IEEE Transactions on Information Theory, 48(3):580–598, Mar 2002.

[FGP20] Daniel Stilck Franca and Raul Garcia-Patron. Limitations of optimizationalgorithms on noisy quantum devices, 2020. arXiv:2009.05532v1.

[GLS12] Martin Grotschel, Laszlo Lovasz, and Alexander Schrijver. Geometric al-gorithms and combinatorial optimization, volume 2. Springer Science &Business Media, 2012.

[Got98] Daniel Gottesman. The heisenberg representation of quantum computers,1998. arXiv:quant-ph/9807006.

[HBVSE18] Dominik Hangleiter, Juani Bermejo-Vega, Martin Schwarz, and Jens Eis-ert. Anticoncentration theorems for schemes showing a quantum speedup.Quantum, 2:65, may 2018.

[HHB+19] Jonas Haferkamp, Dominik Hangleiter, Adam Bouland, Bill Fefferman, JensEisert, and Juani Bermejo-Vega. Closing gaps of a quantum advantage withshort-time hamiltonian dynamics, 2019. arXiv 1908.08069.

[HKEG19] Dominik Hangleiter, Martin Kliesch, Jens Eisert, and Christian Gogolin.Sample complexity of device-independently certified “quantum supremacy”.Phys. Rev. Lett., 122:210502, May 2019.

[HM17] Aram W Harrow and Ashley Montanaro. Quantum computationalsupremacy. Nature, 549(7671):203, 2017.

[HRF20] Christoph Hirche, Cambyse Rouze, and Daniel Stilck Franca. On contrac-tion coefficients, partial orders and approximation of capacities for quantumchannels, 2020. arXiv:2011.05949v1.

[HZN+20] Cupjin Huang, Fang Zhang, Michael Newman, Junjie Cai, Xun Gao,Zhengxiong Tian, Junyin Wu, Haihong Xu, Huanjun Yu, Bo Yuan, MarioSzegedy, Yaoyun Shi, and Jianxin Chen. Classical simulation of quantumsupremacy circuits, 2020. arXiv:2005.06787.

[KBG+11] M. Kliesch, T. Barthel, C. Gogolin, M. Kastoryano, and J. Eisert. Dissipa-tive quantum church-turing theorem. Physical Review Letters, 107(12), sep2011.

[KGV83] S. Kirkpatrick, C. D. Gelatt, and M. P. Vecchi. Optimization by simulatedannealing. Science, 220(4598):671–680, may 1983.

[Kre20] William Kretschmer. The Quantum Supremacy Tsirelson Inequality, 2020.arXiv:2008.08721v1.

[KT13] Michael J. Kastoryano and Kristan Temme. Quantum logarithmicsobolev inequalities and rapid mixing. Journal of Mathematical Physics,54(5):052202, may 2013.

[LBR17] AP Lund, Michael J Bremner, and TC Ralph. Quantum sampling prob-lems, Boson sampling and quantum supremacy. npj Quantum Information,3(1):15, 2017.

[LP17] David A Levin and Yuval Peres. Markov chains and mixing times, volume107. American Mathematical Soc., 2017.

21

Page 22: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

[MHFW16a] Alexander Muller-Hermes, Daniel Stilck Franca, and Michael M. Wolf. En-tropy production of doubly stochastic quantum channels. Journal of Math-ematical Physics, 57(2):022203, feb 2016.

[MHFW16b] Alexander Muller-Hermes, Daniel Stilck Franca, andMichael M. Wolf. Rela-tive entropy convergence for depolarizing channels. Journal of MathematicalPhysics, 57(2):022202, feb 2016.

[MHRW15] Alexander Muller-Hermes, David Reeb, and Michael M. Wolf. Quantumsubdivision capacities and continuous-time quantum coding. IEEE Trans-actions on Information Theory, 61(1):565–581, jan 2015.

[Mov18] Ramis Movassagh. Efficient unitary paths and quantum computationalsupremacy: A proof of average-case hardness of random circuit sampling.arXiv preprint arXiv:1810.04681, 2018.

[NRK+18] C. Neill, P. Roushan, K. Kechedzhi, S. Boixo, S. V. Isakov, V. Smelyanskiy,A. Megrant, B. Chiaro, A. Dunsworth, K. Arya, R. Barends, B. Burkett,Y. Chen, Z. Chen, A. Fowler, B. Foxen, M. Giustina, R. Graff, E. Jeffrey,T. Huang, J. Kelly, P. Klimov, E. Lucero, J. Mutus, M. Neeley, C. Quin-tana, D. Sank, A. Vainsencher, J. Wenner, T. C. White, H. Neven, andJ. M. Martinis. A blueprint for demonstrating quantum supremacy withsuperconducting qubits. Science, 360(6385):195–199, apr 2018.

[PGN+19] Edwin Pednault, John A. Gunnels, Giacomo Nannicini, Lior Horesh, andRobert Wisnieff. Leveraging secondary storage to simulate deep 54-qubitsycamore circuits, 2019. arXiv 1910.09534.

[PWR+19] D. S. Phillips, M. Walschaers, J. J. Renema, I. A. Walmsley, N. Treps,and J. Sperling. Benchmarking of gaussian boson sampling using two-pointcorrelators. Physical Review A, 99(2), feb 2019.

[QBQGP20] Haoyu Qi, Daniel J. Brod, Nicolas Quesada, and Raul Garcia-Patron.Regimes of classical simulability for noisy gaussian boson sampling. PhysicalReview Letters, 124(10), mar 2020.

[SS14] John A. Smolin and Graeme Smith. Classical signature of quantum anneal-ing. Frontiers in Physics, 2, sep 2014.

[SSSV14] Seung Woo Shin, Graeme Smith, John A. Smolin, and Umesh Vazirani.How ”quantum” is the d-wave machine?, 2014. arXiv 1401.7087.

[SVB+14] Nicolo Spagnolo, Chiara Vitelli, Marco Bentivegna, Daniel J. Brod, AndreaCrespi, Fulvio Flamini, Sandro Giacomini, Giorgio Milani, Roberta Ram-poni, Paolo Mataloni, Roberto Osellame, Ernesto F. Galvao, and FabioSciarrino. Experimental validation of photonic boson sampling. NaturePhotonics, 8(8):615–620, jun 2014.

[TRW05] Koji Tsuda, Gunnar Ratsch, and Manfred K Warmuth. Matrix exponen-tiated gradient updates for on-line learning and bregman projection. J.Mach. Learn. Res., 6(Jun):995–1018, 2005.

[WRB+13] Lei Wang, Troels F. Rønnow, Sergio Boixo, Sergei V. Isakov, Zhihui Wang,David Wecker, Daniel A. Lidar, John M. Martinis, and Matthias Troyer.Comment on: ”classical signature of quantum annealing”, 2013. arXiv1305.5837.

22

Page 23: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

A Bound on the Shannon entropy from design prop-

erty

We will now show that the Shannon entropy of the output distributions of approximatetwo designs when measured in the computational basis is essentially maximal. This resultis similar in spirit to those of [AA14a, HKEG19].

Proposition A.1. Let U be a (2, ǫ2−2n−1) approximate unitary design and define ν asbefore. Then, with probability at least 1− δ:

S(ν) ≥ n− log(2− ǫ) + log(δ)

Proof. For a Haar random unitary and x ∈ 0, 1n we have that:

E

(

|〈x|U |0〉|2)

=1

2n, E

(

|〈x|U |0〉|4)

=2

2n(2n + 1).

Thus, for an approximate two design as above, we have that:

E

(

|〈x|U |0〉|4)

≤2

2n(2n + 1)+

ǫ

2n(2n + 1).

Recall that the 2-Renyi entropy S2 is defined as:

S2(ν) = − log

(

x

ν(x)2

)

and that S(ν) ≥ S2(ν). Moreover, the function − log is convex. Thus, it follows fromJensen’s inequality that:

E

(

− log

(

x

ν(x)2

))

≥ − log

(

E

[

x

ν(x)2

])

.

From the computations above, we have that:

2 + ǫ

(2n + 1)≥ E

[

x

ν(x)2

]

,

from which we readily obtain that:

E

(

− log

(

x

ν(x)2

))

≥ n− log(2 + ǫ).

It follows from Markov’s inequality that

P

(

x

ν(x)2 ≥2 + ǫ

δ2n

)

≤ δ, (24)

from which the claim follows.

23

Page 24: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

B Auxiliary results for Section 5

In Section 5 we showed that being able to efficiently distinguish the probability distribu-tions arising from sampling from random circuits from the uniform distribution impliesthe ability to fool the XHOG problem associated to the same class of circuits to a certainlevel.

But we assumed that the function that distinguished the distributions has binaryoutputs, although our framework for distinguishability allows for functions with outputs in[0, 1]. We now show that it is always possible to obtain an efficiently computable functionwith binary outputs from an efficiently function with image [0, 1] that distinguishes thetwo probability distributions, at the expense of a smaller distinguishability power.

Lemma B.1. Let f : 0, 1n → [0, 1] be a function that can be computed in polynomialtime such that for some probability measure ν and ǫ > 0 we have that:

Eν(f)− EU (f) ≥ ǫ.

Then there exists a function f ′ : 0, 1n → 0, 1 that can be computed in polynomialtime such that:

Eν(f′)− EU (f

′) ≥ǫ2

4.

Proof. Assume w.l.o.g. that ǫ = m−1 for some integer m and consider the discretizationof f1 of f given by:

f1(x) = (2m)−1⌈2mf(x)⌉.

Note that ‖f − f0‖∞ ≤ (4m)−1 and, thus,

Eν(f0)− EU (f0) ≥ǫ

2.

Moreover, note that f1 only takes the 2m possible values (2m)−1, . . . , 1. Combiningthis fact and that for every real valued random variable X we have:

E(X) =

P(X ≥ x)dx,

we see that

Eν(f0) =2m∑

k=1

ν

(

f0(x) ≥k

2m

)

≥ǫ

4+

2m∑

k=1

∣f0(x) ≥k2m

2n.

It then follows that for at least one 1 ≤ k0 ≤ 2m we have that:

ν

(

f0(x) ≥k

2m

)

∣f0(x) ≥k2m

2n+

ǫ

8m. (25)

Thus, by setting f ′(x) = 1 if f0(x) ≥k2m and 0 else and recalling that m−1 = ǫ, we see

that Eq. (25) immediately implies

Eν(f′)− EU (f

′) ≥ǫ2

8,

which yields the claim.

24

Page 25: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

C Distinguishing the output distribution of stabilizer

states

Note that we only assumed in the proofs of Sec. 5 that the underlying circuit ensemble isan approximate two design. It is well-known that circuits being approximate two designsdoes not imply that one cannot sample from their output distribution efficiently, as isprominently exemplified by Clifford circuits. Random Cliffords are two-designs [DCEL09,DLT02] and it is possible to simulate measurements in the computational basis efficientlyfor them [Got98, AG04]. We now show that in this case, it is also possible to easilydistinguish the outcome distribution from the uniform one, if this is possible at all.

It is not difficult to see that for a Pauli string P = ⊗ni=1σi we have for a Clifford C

that

tr(

PC|0〉〈0|⊗nC†)

∈ −1, 0, 1.

This is because, as Cliffords stabilize the Pauli group, we have that P = C†PC is again,

up to a global sign, a Pauli string. And for a Pauli string tr(

P |0〉〈0|⊗n)

∈ 0, 1. Now

assume that there exists a Pauli string P consisting only of I and Z Pauli matrices thatdiffers from the identity and such that

tr(

PC|0〉〈0|⊗nC†)

6= 0.

Then, interpreting diagonal operators as functions, we have that the function f = P+I2

is a binary function that satisfies:

|Eν(f)− EU (f)| =1

2.

Thus, in this case, we have found a function that efficiently distinguishes the outcomefrom uniform. But note that in some cases such a Pauli string does not exist, such as ifthe Clifford is H⊗n, as then the outcome distribution is uniform.

Let us now discuss how to efficiently find the appropriate distinguishing Pauli string.We refer to [AG04] for a review of the basics of the stabilizer formalism. First, we recallthat a stabilizer state |ψ〉 on n qubits can always be described by n generators g1, . . . , gnof its stabilizer group SG(|ψ〉). Moreover, note that for a Pauli string

tr(

PC|0〉〈0|⊗nC†)

∈ −1, 1.

is equivalent to P ∈ SG(|ψ〉) or −P ∈ SG(|ψ〉). Thus, by our previous discussion, theproblem of finding a function to distinguish the output of the Clifford circuit from theuniform distribution is equivalent to finding a stabilizer of the state consisting solely ofZ and I Pauli operators.

Let us now discuss how to achieve this. First, decompose each generator gi as theproduct of a string of Pauli X and Pauli Z matrices plus a global ±1 phase and representeach one of these as vectors (xi, zi, si) in F

2n+12 . In order to simplify the presentation,

we are not going to keep track of the global phase of the elements of the stabilizer groupfor now. Thus, we restrict to the vectors (xi, zi) corresponding to the first 2n entries.It is easy to see that if we do not keep track of the global phase, then multiplying twogenerators is equivalent to adding the corresponding vectors in F

2n2 . Now define the

n× (2n+ 1) binary matrix A with the vectors xi in its rows. We then have:

Proposition C.1. Let |ψ〉 be a stabilizer state with generators g1, . . . , gn and correspond-ing vectors (xi, zi) ∈ Z

2n. Then there exists a Z string P ∈ SG(|ψ〉) if and only if:

span(x1, y1), . . . , (xn, yn) ∩ (0 × Zn) 6= 0 (26)

25

Page 26: simulation · 2020. 11. 25. · simulation Daniel Stilck Franc¸a1 and Raul Garcia-Patron2 1QMATH, Department of Mathematical Sciences, University of Copenhagen, Denmark 2School of

Proof. Note that

span(x1, y1), . . . , (xn, yn)

corresponds to the elements of the stabilizer group of the state, up to a global phase.This is because, as discussed before, multiplication in the Pauli group just correspondsto a sum of the vectors, up to the global phase. Thus, if we find a string of Z in thestabilizer group, then it is also in the intersection in eq. (26).

Thus, we can find the distinguishing Pauli operator by a nonzero element of thesubspace

span(x1, y1), . . . , (xn, yn) ∩ (0 × Zn) ,

which can be done by Gaussian elimination.

26