kibble-zurek mechanism and beyond - arxiv3department of physics, tohoku university, sendai 980-8578,...

22
Probing the Universality of Topological Defect Formation in a Quantum Annealer: Kibble-Zurek Mechanism and Beyond Yuki Bando, 1, * Yuki Susa, 2, Hiroki Oshiyama, 3 Naokazu Shibata, 3 Masayuki Ohzeki, 4, 1, 5 Fernando Javier omez-Ruiz, 6 Daniel A. Lidar, 7, 8 Adolfo del Campo, 6, 9, 10 Sei Suzuki, 11 and Hidetoshi Nishimori 1, 4, 12 1 Institute of Innovative Research, Tokyo Institute of Technology, Nagatsuta-cho, Midori-ku, Yokohama 226-8503, Japan 2 Institute of Innovative Research, Tokyo Institute of Technology, Oh-okayama, Meguro-ku, Tokyo 152-8550, Japan 3 Department of Physics, Tohoku University, Sendai 980-8578, Japan 4 Graduate School of Information Sciences, Tohoku University, Sendai 980-8579, Japan 5 Sigma-i Co. Ltd., Konan, Minato-ku, Tokyo 108-0075, Japan 6 Donostia International Physics Center, E-20018 San Sebasti´ an, Spain 7 Departments of Electrical and Computer Engineering, Chemistry, and Physics, University of Southern California, Los Angeles, CA 90089, USA 8 Center for Quantum Information Science & Technology, University of Southern California, Los Angeles, CA 90089, USA 9 IKERBASQUE, Basque Foundation for Science, E-48013 Bilbao, Spain 10 Department of Physics, University of Massachusetts Boston, 100 Morrissey Boulevard, Boston, MA 02125, USA 11 Department of Liberal Arts, Saitama Medical University, Moroyama, Saitama 350-0495, Japan 12 RIKEN, Interdisciplinary Theoretical and Mathematical Sciences (iTHEMS), Wako, Saitama 351-0198, Japan (Dated: May 28, 2020) The number of topological defects created in a system driven through a quantum phase transition exhibits a power-law scaling with the driving time. This universal scaling law is the key prediction of the Kibble-Zurek mechanism (KZM), and testing it using a hardware-based quantum simulator is a coveted goal of quantum information science. Here we provide such a test using quantum annealing. Specifically, we report on extensive experimental tests of topological defect formation via the one-dimensional transverse-field Ising model on two different D-Wave quantum annealing devices. We find that the quantum simulator results can indeed be explained by the KZM for open-system quantum dynamics with phase-flip errors, with certain quantitative deviations from the theory likely caused by factors such as random control errors and transient effects. In addition, we probe physics beyond the KZM by identifying signatures of universality in the distribution and cumulants of the number of kinks and their decay, and again find agreement with the quantum simulator results. This implies that the theoretical predictions of the generalized KZM theory, which assumes isolation from the environment, applies beyond its original scope to an open system. We support this result by extensive numerical computations. To check whether an alternative, classical interpretation of these results is possible, we used the spin-vector Monte Carlo model, a candidate classical description of the D-Wave device. We find that the degree of agreement with the experimental data from the D-Wave annealing devices is better for the KZM, a quantum theory, than for the classical spin-vector Monte Carlo model, thus favoring a quantum description of the device. Our work provides an experimental test of quantum critical dynamics in an open quantum system, and paves the way to new directions in quantum simulation experiments. I. INTRODUCTION Quantum simulations are emerging to be one of the im- portant applications of quantum annealing [14], quite different, and arguably more natural, than the original intent of using such devices for optimization, the sub- ject of many recent studies [515]. Prominent examples include the simulation of the Kosterlitz-Thouless topo- logical phase transition [16, 17] and three-dimensional spin glasses [18] using the D-Wave quantum annealing * [email protected] Present address: System Platform Research Laboratories, NEC Corporation, Kawasaki 211-8666, Japan devices, that have successfully reproduced the behav- ior of various physical quantities and the structure of the phase diagram, as predicted by classical simulations. Quantum simulation has also been pursued using other systems such as trapped ions [1921]. Here we use D-Wave quantum annealers to perform quantum simulations of the Kibble-Zurek mechanism (KZM) [22, 23], which predicts the kink (or defect [24]) formation when a system crosses a phase transition point at a finite rate. While the theory of the KZM was originally formulated for classical phase transitions, it has been extended to describe quantum critical dynam- ics [2528]. As the dominant paradigm to describe the universal dynamics of a quantum phase transition, it has motivated a wide variety of experimental and theoretical studies [2932]. Laboratory tests of the KZM in quan- arXiv:2001.11637v3 [quant-ph] 26 May 2020

Upload: others

Post on 21-Jan-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

Probing the Universality of Topological Defect Formation in a Quantum Annealer:Kibble-Zurek Mechanism and Beyond

Yuki Bando,1, ∗ Yuki Susa,2, † Hiroki Oshiyama,3 Naokazu Shibata,3 Masayuki Ohzeki,4, 1, 5 Fernando Javier

Gomez-Ruiz,6 Daniel A. Lidar,7, 8 Adolfo del Campo,6, 9, 10 Sei Suzuki,11 and Hidetoshi Nishimori1, 4, 12

1Institute of Innovative Research, Tokyo Institute of Technology,Nagatsuta-cho, Midori-ku, Yokohama 226-8503, Japan

2Institute of Innovative Research, Tokyo Institute of Technology,Oh-okayama, Meguro-ku, Tokyo 152-8550, Japan

3Department of Physics, Tohoku University, Sendai 980-8578, Japan4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579, Japan

5Sigma-i Co. Ltd., Konan, Minato-ku, Tokyo 108-0075, Japan6Donostia International Physics Center, E-20018 San Sebastian, Spain

7Departments of Electrical and Computer Engineering, Chemistry,and Physics, University of Southern California, Los Angeles, CA 90089, USA

8Center for Quantum Information Science & Technology,University of Southern California, Los Angeles, CA 90089, USA

9IKERBASQUE, Basque Foundation for Science, E-48013 Bilbao, Spain10Department of Physics, University of Massachusetts Boston,

100 Morrissey Boulevard, Boston, MA 02125, USA11Department of Liberal Arts, Saitama Medical University, Moroyama, Saitama 350-0495, Japan

12RIKEN, Interdisciplinary Theoretical and Mathematical Sciences (iTHEMS), Wako, Saitama 351-0198, Japan(Dated: May 28, 2020)

The number of topological defects created in a system driven through a quantum phase transitionexhibits a power-law scaling with the driving time. This universal scaling law is the key predictionof the Kibble-Zurek mechanism (KZM), and testing it using a hardware-based quantum simulatoris a coveted goal of quantum information science. Here we provide such a test using quantumannealing. Specifically, we report on extensive experimental tests of topological defect formationvia the one-dimensional transverse-field Ising model on two different D-Wave quantum annealingdevices. We find that the quantum simulator results can indeed be explained by the KZM foropen-system quantum dynamics with phase-flip errors, with certain quantitative deviations fromthe theory likely caused by factors such as random control errors and transient effects. In addition,we probe physics beyond the KZM by identifying signatures of universality in the distribution andcumulants of the number of kinks and their decay, and again find agreement with the quantumsimulator results. This implies that the theoretical predictions of the generalized KZM theory,which assumes isolation from the environment, applies beyond its original scope to an open system.We support this result by extensive numerical computations. To check whether an alternative,classical interpretation of these results is possible, we used the spin-vector Monte Carlo model, acandidate classical description of the D-Wave device. We find that the degree of agreement with theexperimental data from the D-Wave annealing devices is better for the KZM, a quantum theory,than for the classical spin-vector Monte Carlo model, thus favoring a quantum description of thedevice. Our work provides an experimental test of quantum critical dynamics in an open quantumsystem, and paves the way to new directions in quantum simulation experiments.

I. INTRODUCTION

Quantum simulations are emerging to be one of the im-portant applications of quantum annealing [1–4], quitedifferent, and arguably more natural, than the originalintent of using such devices for optimization, the sub-ject of many recent studies [5–15]. Prominent examplesinclude the simulation of the Kosterlitz-Thouless topo-logical phase transition [16, 17] and three-dimensionalspin glasses [18] using the D-Wave quantum annealing

[email protected]† Present address: System Platform Research Laboratories, NEC

Corporation, Kawasaki 211-8666, Japan

devices, that have successfully reproduced the behav-ior of various physical quantities and the structure ofthe phase diagram, as predicted by classical simulations.Quantum simulation has also been pursued using othersystems such as trapped ions [19–21].

Here we use D-Wave quantum annealers to performquantum simulations of the Kibble-Zurek mechanism(KZM) [22, 23], which predicts the kink (or defect [24])formation when a system crosses a phase transition pointat a finite rate. While the theory of the KZM wasoriginally formulated for classical phase transitions, ithas been extended to describe quantum critical dynam-ics [25–28]. As the dominant paradigm to describe theuniversal dynamics of a quantum phase transition, it hasmotivated a wide variety of experimental and theoreticalstudies [29–32]. Laboratory tests of the KZM in quan-

arX

iv:2

001.

1163

7v3

[qu

ant-

ph]

26

May

202

0

Page 2: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

2

tum platforms have been carried out pursuing differentquantum simulation approaches, e.g., using qubits to em-ulate free fermion models [33–36] and via a fully digitalapproach using Rydberg atoms [37].

Tests of the KZM can also be used to quantitativelyassess the performance of a quantum device. Indeed,the KZM scaling is sensitive to, e.g., nonlinear drivingschemes [38, 39], disorder [40, 41], inhomogeneities in thesystem [42–45] and decoherence [46–49]. The use of theKZM to assess the performance of a quantum annealerwas studied by Gardas et al. [50], focusing on the one-dimensional case on two previous generation D-Wave 2Xdevices, and by Weinberg et al. [51] on a current gener-ation D-Wave 2000Q (DW2KQ) device, focusing on theIsing Hamiltonian on the two-dimensional square lattice.

Here we report on extensive DW2KQ experiments forthe one-dimensional transverse-field Ising model, usingtwo separate realizations of the device to perform quan-tum simulations of the predictions of the KZM for thekink density. We also test a recent theory for the kinkdensity distribution developed by one of us [52], thusprobing physics beyond the original KZM prediction ofthe average number of kinks [25–28]. Unlike Gardas etal. [50], our work finds a universal power-law scalingbehavior of the average number of kinks. We choosethe one-dimensional problem because departures fromthe ideal theoretical setting due to noise and other rea-sons would easily destroy ordering in one dimension, andtherefore it is easy to detect the effects of imperfectionsin one dimension, implying that the data would clearlyreveal open system effects. It is also an advantage of theone-dimensional problem that we can avoid the problemof embedding of the system on the Chimera graph of theD-Wave device [53]. Moreover, previous studies of bothantiferromagnetic [54, 55] and ferromagnetic chains [56]using previous generations of D-Wave devices obtainedgood agreement with open quantum systems theory [57].

Our work establishes the power-law scaling behaviorof the average number of kinks, variance and third-ordercumulant with the timescale in which the transition iscrossed. In doing so, we provide a strategy to assess thebehavior of quantum annealers, and find it to be welldescribed within the framework of open system quan-tum dynamics. Our work thus provides an experimentaltest of quantum critical dynamics in an open quantumsystem. The universal power law scaling found in thecumulants of the kink-number distribution shows thatsignatures of universality beyond the KZM recently pre-dicted in isolated quantum critical systems continue tohold in the presence of coupling to an environment, towhich we provide support by numerical computations.

This paper is organized as follows. Background on theKZM and its generalization, the problem we study, andthe experimental methods are described in Sec. II. Theempirical results on the kink density are presented andcompared with the generalized KZM theory in Sec. III,and Sec. IV similarly presents the kink distribution re-sults. In Sec. V we address the question of whether clas-

sical models suffice to explain our empirical results. Wedo this by modeling the kink distribution using the clas-sical Boltzmann distribution of the Ising spin chain, andby comparing the empirical results to the predictions ofthe classical spin-vector Monte Carlo model. We closethe paper with a discussion in Sec. VI, including a com-parison with Refs. [50, 51], and conclude in Sec. VII.Additional materials are presented in the Appendixes.

II. THEORETICAL AND EXPERIMENTALBACKGROUND

We first describe the problem to be studied, and thenexplain the predictions of the KZM, followed by our ex-perimental methods for testing the theoretical predic-tions.

A. The problem studied

The target system of our investigation is the one-dimensional transverse-field Ising model defined and pa-rameterized by the Hamiltonian,

H(s) =A(s)

2

L∑i=1

σxi +B(s)

2

L−1∑i=1

Jσzi σzi+1, (1)

where L is the chain length (system size), s = t/ta, withthe time t ∈ [0, ta] and the final time t = ta being theannealing time. The absolute value of the interactionstrength is chosen to be |J | = 1, and as explained belowwe consider both the ferromagnetic (J < 0) and anti-ferromagnetic (J > 0) cases. We adopt a free bound-ary condition as indicated by the upper bound L − 1in Eq. (1). Effects of the choice of a specific boundarycondition are of the order of 1/L, which is much smallerthan the statistical fluctuations in the data shown below.The functional forms of the annealing schedules A(s) andB(s) are shown in Fig. 1.

0

2

4

6

8

10

12

14

0 0.2 0.4 0.6 0.8 1

Ener

gy

[GH

z]

s

A(s)

B(s)

NASA

A(s)

B(s)

0 0.2 0.4 0.6 0.8 1s

Burnaby

(a) (b)

FIG. 1. The annealing schedules on the DW2KQ quantumannealers at (a) NASA Ames Research Center and (b) Burn-aby. For the actual quantum annealing processes, A(s)/2 andB(s)/2 are used as in Eq. (1). The energy scale is convertedinto frequency, i.e., the vertical axis is E/h, where h is thePlanck constant.

Page 3: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

3

This one-dimensional model has a second-order quan-tum phase transition at A(s) = B(s) for a time-independent system, i.e., when s is regarded as a fixedparameter [58]. The system is in the ferromagnetic phasewhen A(s) < B(s) and is paramagnetic for A(s) > B(s).The system is initially in the paramagnetic phase sinceA(0) > 0 and B(0) ≈ 0 as seen in Fig. 1. The rateof change of the annealing schedules A(s) and B(s) isfinite, and the system does not necessarily follow the in-stantaneous ground state even when the initial conditionis chosen to be the ground state of the initial HamiltonianH(0). Thus, at the end of the annealing schedule, whenA(1) = 0 and B(1) > 0, the system is generally in anexcited state with a number of kinks (defects) separatingferromagnetic domains (regions with aligned neighbor-ing spins) when J < 0, or kinks separating antiferro-magnetic domains (regions with anti-aligned neighboringspins) when J > 0.

B. Kibble-Zurek mechanism and its extension

The Kibble-Zurek mechanism [28] describes the pro-cess of kink formation assuming that the ratio of the pa-rameters A(s) and B(s) changes linearly as a function oftime near the critical point A(s)/B(s) = 1. This assump-tion of linear time dependence in the vicinity of the crit-ical point is reasonable because any analytical functioncan be expanded to linear order and we are interested inthe system behavior near the critical point. Non-analyticdriving schedules can also be accounted for within theKZM framework [38, 39, 45].

Let us state the main theoretical predictions of theKZM and its generalization to be tested on the D-Wavedevice.

1. The kink density ρkink, the number of kinks dividedby the system size, follows the formula

ρkink ∝ ta−dν

1+zν , (2)

where d is the spatial dimension, ν is the criticalexponent for the correlation function, and z is thedynamical critical exponent. In our case d = ν =z = 1, and thus:

ρkink ∝ ta−12 . (3)

2. The qth cumulant κq of the distribution function ofthe number of kinks P (n), divided by the averageκ1 = 〈n〉, is independent of the annealing time. Inparticular, for the one-dimensional transverse-fieldIsing model the second and third cumulants satisfy

κ2κ1

= 2−√

2 ≈ 0.586,

κ3κ1

= 4− 12√2

+8√3≈ 0.134.

(4)

3. Since the third and higher-order cumulants aresmall relative to the first and second order ones,the distribution function can be well approximatedby a Gaussian distribution

P (n) =1√

2π(2−√

2)〈n〉exp

[− (n− 〈n〉)2

2(2−√

2)〈n〉

]. (5)

Below we briefly describe how these formulas are de-rived based on Refs. [22, 28] for item 1, and on Ref. [36](its Supplementary Note 2 in particular) as well as onRef. [52] for items 2 and 3, to provide pertinent physicalbackground for our study. Readers interested only in theresults can skip to Sec. II C.

Second-order continuous phase transitions are charac-terized by the divergence of the correlation length ξ andthe relaxation time τ at the critical point. Specifically, asa function of the difference between the value of the con-trol parameter λ and its critical value λc, both quantitiesξ and τ exhibit a power-law behavior

ξ = ξ0|ε|−ν , τ = τ0|ε|−zν , (6)

where ε = (λ−λc)/λc, and ξ0 and τ0 are constants. Thedivergence of the relaxation time introduces a separationof time scales and allows one to describe the crossingof the phase transition as a sequence of stages: In thefirst stage, far from criticality where |ε| is not very small,the relaxation time is not large and system follows theinstantaneous equilibrium state, the ground state in thecontext of quantum phase transition at zero temperature.The system evolves adiabatically. Then, in the secondstage, as |ε| becomes small, the relaxation time growsrapidly and the state of the system has no time to relaxto the ground state, and the system becomes effectivelyfrozen. As the parameter further changes, the systementers the final third stage, and |ε| again becomes large,thus the dynamics becomes adiabatic again. This is theso-called adiabatic-impulse approximation [28, 59].

The key testable prediction of the KZM is that, af-ter crossing the phase transition in the second stage, theaverage length scale in which the order-parameter is uni-form is set by the equilibrium value of the correlationlength when the system unfreezes at the point where thethird stage is reached.

To formulate this idea quantitatively, consider a driv-ing scheme such that the distance to the critical pointvaries linearly in time according to ε = (t − tc)/ta ona timescale ta, where tc denotes the instant when thesystem parameters cross the critical point. Equatingthe instantaneous equilibrium relaxation time τ(t) tot − tc, the time elapsed after crossing the critical point,yields the freeze-out time scale t − tc = (τ0t

zνa )1/(1+zν),

which yields the average correlation length as ξ ≡ ξ(t) =ξ0(τ0/ta)ν/(1+zν). A kink may form at the interface be-

tween different domains of size ξ. Then the kink density

is given by the inverse of the volume ξd and scales as a

Page 4: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

4

universal power-law [22, 23]

ρkink =1

ξd∝(τ0ta

) dν1+zν

, (7)

for a system in d spatial dimensions. This is Eq. (2).When the system size is Ld, the average number of kinksis thus 〈n〉 = ρkinkL

d.This picture applies both to classical and quantum

phase transitions. In the quantum case, the relaxationtime is identified with the inverse of the energy gap be-tween the ground state and the first excited state thatcloses at the critical point, and the KZM describes thecritical dynamics as well [25–28].

The above physical picture is quite generic and theresult is valid independent of the details of the systemHamiltonian. If we restrict ourselves to the quantumphase transition of the one-dimensional transverse-fieldIsing model, more detailed information can be extractedon the distribution of kink numbers as follows [36, 52].

The one-dimensional transverse-field Ising model witha periodic boundary can be solved (diagonalized) un-der periodic boundary condition by the Jordan-Wignertransformation [58], which rewrites the spin operatorsin terms of spinless fermion operators. Kinks appear inpairs under periodic boundary, and we therefore considerthe number of kink pairs, which is described by the op-erator (for the ferromagnetic case)

N =1

4

L∑j=1

(1− σzjσzj+1

)=∑k≥0

γ†

kγk, (8)

where L is the total number of sites and γ†k and γk arecreation and annihilation operators of fermions. The dis-tribution function of kink pairs is defined by

P (n) = Tr(ρ δ(N − n)

), (9)

where ρ is the density matrix for the state after anneal-ing. It helps to use the characteristic function P (θ), theFourier transform,

P (n) =1

∫ π

−πdθ P (θ) e−inθ. (10)

Since kink pairs with different wave numbers are inde-pendent, the characteristic function is decomposed intoa product

P (θ) =∏k≥0

[1 + (eiθ − 1)pk

], (11)

where

pk = 〈γ†kγk〉 = e−2πJtak2/~. (12)

We have used the Landau-Zener formula for the creationof a kink pair. Equation (11) indicates that the numberof kink pairs follows the Poisson binomial distribution.

Then the cumulants are easily evaluated, the first threeof which are

κ1 = 〈n〉 =∑k≥0

pk, (13a)

κ2 =∑k≥0

pk(1− pk), (13b)

κ3 =∑k≥0

pk(1− pk)(1− 2pk). (13c)

In the long time scale limit ta � ~/(2π3J), the firstcumulant (the average), reduces to

κ1 =∑k≥0

pk =L

∫ π

0

dk e−2πJtak2/~

→ L

∫ ∞0

dk e−2πJtak2/~ =

L

√~

2Jta. (14)

This is consistent with Eq. (3) of the KZM. Similarly, thesecond and third cumulants are evaluated to yield

κ2 =

(1− 1√

2

)κ1, (15a)

κ3 =

(1− 3√

2+

2√3

)κ1. (15b)

The cumulants κq for the number of kinks can be derivedfrom the above cumulants for the number of kink pairsas κq = 2qκq,

κ1 = 2κ1 =L

√~

2Jta, (16a)

κ2 = 4κ2 = (2−√

2)κ1, (16b)

κ3 = 8κ3 =

(4− 12√

2+

8√3

)κ1. (16c)

These give Eq. (4).The Gaussian distribution Eq. (5) follows from setting

to zero all cumulants κq with q ≥ 3, a reasonable approx-imation as their value is much smaller than κ1 and κ2;see [36, 52, 60].

C. Experimental methods

We used two different DW2KQ devices, one locatedat the NASA Ames Research Center and the other atD-Wave Systems, Inc. in Burnaby. The latter is alower-noise version of the former (for documentation seeRef. [61]). The D-Wave Chimera graph comprises ` × `unit cells of sparsely connected K4,4 bipartite graphs, fora total of 8`2 qubits, each coupled to up to 6 other qubits.We chose four chain lengths: L = 50, 200, 500, and 800.For each size we generated 200 instances of configura-tions of the one-dimensional chain with a free boundaryby self-avoiding random walks starting from a randomly

Page 5: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

5

selected qubit on the Chimera graph. For each of these200 instances, we carried out 1, 000 annealing cycles ata given annealing time ta. Thus, we generated 200, 000samples for each ta and L, and recorded the distribution(histogram) of the kink density. The annealing time taranges from 1µs to 2ms, for a total of 33 values.

We tested three cases of the coupling parameter J :ferromagnetic (J = −1), antiferromagnetic (J = 1), andrandomly chosen gauges. The latter starts from ferro-magnetic interactions, then half of the qubits are chosenrandomly and the signs of their interactions are flipped.This prescription is meant to cancel (unintended, device-specific) local biases toward a specific direction at eachqubit. As shown in Appendix A, the antiferromagneticand random-gauge cases give almost identical results,while the ferromagnetic case tends to exhibit unstablebehavior. We therefore show results for the antiferro-magnetic case in the main text.

III. AVERAGE KINK DENSITY

The average kink density as a function of the anneal-ing time ta and for different sizes L is shown in Fig. 2(a)for the NASA device and Fig. 2(b) for the Burnaby de-vice. We analyze the data for the time range ta ≤ 100 µsbecause the data beyond 100 µs show different, less sta-ble, behavior. Likely reasons include the effect of 1/fnoise, which becomes apparent at long annealing times,and a significant increase in the persistent current forta > 100 µs [61], which reduces qubit coherence. SeeAppendix A for data beyond 100 µs and a more detaileddiscussion.

The KZM assumes that the number of kinks is at least1 on average, which means that the inequality ρkink >1/L should hold. We therefore also exclude the data forL = 50 from the analysis because the kink density is toolow: we find empirically (see Fig. 2) that ρkink < 1/L =0.02, which implies that the KZM does not apply.

A first qualitative observation from Fig. 2 is that thekink density obeys a power law. It is also clearly seenthat the kink density is lower on the Burnaby device inFig. 2(b) than on the NASA device in Fig. 2(a) for thesame parameter values L and ta. This is in accordancewith the ‘low-noise’ characteristics of the Burnaby de-vice [61].

Delving deeper into quantitative aspects, we fit thedata to the power law

ρkink ∝ ta−α (17)

and evaluate the exponent α. The result is given in Ta-ble I.

Apart from the case of L = 50, the exponent α is al-most independent of the chain length L. The NASA de-vice has α ≈ 0.20 (L = 800) and the Burnaby device hasα ≈ 0.34 (L = 800). These values are far from the KZMprediction of 0.5 in Eq. (3). Preliminary numerical sim-ulations under unitary dynamics suggest that the value

TABLE I. Results from D-Wave device runs for the exponentα of the power-law scaling describing the decay rate of thekink density as shown in Fig. 2.

L NASA Burnaby50 0.347±0.008 0.587±0.016200 0.216±0.003 0.363±0.003500 0.201±0.003 0.320±0.005800 0.204±0.002 0.335±0.003

of α may not be attributed to the nonlinear functionalform of A(s) and B(s). As the schedules can be effec-tively linearized, corrections to KZM behavior resultingfrom nonlinear passage across the critical point [38, 39]may be ruled out. It is thus reasonable to suspect thatthe difference originates from deviations from unitary dy-namics that are not accounted for in the theory.

A natural first step is therefore to incorporate the cou-pling of qubits to the environment, for which we use thestandard spin-boson model with the following Hamilto-nian [62]:

Htotal = H(s) +∑i,k

{Ck(a†i,k + ai,k)σzi + ~ωi,ka†i,kai,k

},

(18)

where H(s) is the original Hamiltonian of Eq. (1). In-dependent bosons (harmonic oscillators) with frequencyωi,k couple to the z component of the ith Pauli matrix.The coefficient Ck is assumed to have an Ohmic spec-trum,

J(ω) =4π

~2∑i

C2kδ(ω − ωk) = 2πηω (ω < ωc) (19)

with the sharp cutoff frequency ωc and the coupling con-stant η.

The Ohmic spin-boson approach has been successfullyused many times in modeling the dynamics of open sys-tem quantum annealing [63–72]. In particular, Ref. [56]reported a closely related open quantum systems study oftransverse field Ising spin chains with alternating sectorsof strong/weak ferromagnetic coupling, but this studydid not include a comparison to KZM theory.

Ground-state (time-independent) properties of theabove model have already been studied by quantumMonte Carlo simulations [73] and renormalization groupmethods [74, 75]. The conclusion of these papers is thatthe quantum phase transition persists under a bosonic en-vironment and the values of the critical exponents changefrom ν = z = 1 for the isolated system to ν = 0.64and z = 1.99 for the system coupled to a zero tempera-ture bosonic environment. Note the sharp contrast withother models of decoherence which do not alter the barecritical exponents and lead to environmentally-inducedheating [46–48].

The modified values of the critical exponents ν andz are independent of the coupling constant η(> 0) in

Page 6: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

6

(a) (b)

10-3

10-2

100

101

102

rkink

ta

[μs]

L = 50

L = 200

L = 500

L = 800

10-3

10-2

100

101

102

rkink

ta

[μs]

L = 50

L = 200

L = 500

L = 800

FIG. 2. Kink density as a function of annealing time (log-log scale). Error bars indicate the 68% confidence interval. (a)Data from the C16 solver on the DW2KQ at NASA. (b) Data from the D 2000Q 5 solver on the DW2KQ in Burnaby. Data isaveraged over 200, 000 samples at each value of ta.

100 101 102 103

10-1

100

0.2

0.4

0.6

1

25

100 101 102

10-1

100

0.470.40.350.32

0.25

1.280.980.720.5

0.320.180 (closed) 0.5

(a) (b)

FIG. 3. Numerically computed kink density as a function of annealing time (log-log scale), using iTEBD with QUAPI. Alinear annealing schedule, Alin(s)/2 = 1 − s and Blin(s)/2 = s, is employed here, where the unit of energy is given by B(1)/2with B(1) provided in Fig. 1. The Ohmic cutoff frequency is ωc = 5 [B(1)/2~], the Trotter time slice is ∆t = 0.05 [2~/B(1)],the cutoff memory time is τc = 10 [2~/B(1)], and the bond dimension is up to 128. (a) Results for various temperatures and afixed coupling strength η = 0.08. Dashed horizontal lines show the thermal expectation values at temperatures T = 5, 2, and 1from the top. The unit of temperature is given by B(1)/2kB . (b) Results for zero temperature and various coupling strengths.The rightmost eight data points for each coupling strength are fitted with the power law, Eq. (17), and the correspondingexponent α is provided above each data set shown.

Eq. (19). We assume that the KZM applies to the presentopen system case because KZM theory is developed basedonly on the divergence of the relaxation time near a crit-ical point, without recourse to a microscopic Hamilto-nian. We therefore apply the generic Eq. (7) to findthe exponent ν/(1 + zν) = 0.28, about half of the iso-lated case of 0.5 in Eq. (3). Although this open-systemtheoretical value of 0.28 is still different from the ex-perimental values of 0.20 (NASA) and 0.34 (Burnaby),

the spin-boson model of Eq. (18) significantly reducesthe difference between theory and experiment, in com-parison with the closed-system theoretical value of 0.5as illustrated in Fig. 4. It is therefore reasonable toconclude that the Hamiltonian Eq. (18) captures, to afirst approximation, the essential features of the behav-ior of the D-Wave devices near the critical point of theone-dimensional transverse-field Ising model embeddedon the Chimera graph. To achieve more precise quan-

Page 7: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

7

0 0.50.280.20 0.34 0.48 α

D(N) D(B)T(O) T(C)C

FIG. 4. The values of the decay exponent α by various meth-ods. D(N): Data from DW2K at NASA. T(O): Open-systemtheory. D(B): Data from DW2K at Burnaby. C: ClassicalSVMC (Sec. V B). T(C): Closed-system theory.

titative agreement between theory and experiment, wewould need to incorporate additional elements that havenot been taken into account so far. Such may include (i)finite temperature effects not considered in Refs. [73–75];(ii) transient phenomena due to the finite annealing time;and (iii) control errors, i.e., imprecision in the parametersetting in the devices [15].

The impacts of the first two items (i) and (ii) listedabove were studied by some of us in Refs. [76, 77]. Ex-tensive numerical computations using the time-evolvingblock decimation (TEBD) method, as well as infinite-TEBD (iTEBD) combined with the quasi-adiabatic prop-agator path integral (QUAPI) reveal that, as shown inFig. 3, (a) the kink density approaches a temperature-dependent constant as ta becomes very large; (b) the kinkdensity may behave non-monotonically as a function ofta in the transient time range if the temperature is fi-nite; (c) the effective exponent α in ρkink ∝ ta−α arounda given time ta depends on the coupling strength evenwhen the temperature is zero.

More precisely, Fig. 3(a) shows the temperature de-pendence of ρkink for a fixed coupling strength η ob-tained by iTEBD with QUAPI. One can see that thecurve of ρkink for finite temperature deviates upwardsfrom that of the zero temperature case with increasingta and the deviation is more pronounced for higher tem-peratures. The results for the temperatures T = 1, 2, and5 [in units of B(1)/2kB ] imply that ρkink behaves non-monotonically with ta and would approach the thermalaverage, [1− tanh(B(1)/2kBT )]/2, as ta →∞. Since ourdata in Fig. 2 do not show an approach to a constant, wemay conclude that temperature effects in the form con-sidered in Ref. [76] have not come into play in our datafor the present range of annealing time.1

Regarding the observation (c), Fig. 3(b) shows the de-pendence of the slope α on the coupling strength η atzero temperature. Note the transient effects, which in-crease in magnitude with η, and also extend to largerta. We would expect the exponent α to approach a con-stant independent of the coupling constant for sufficientlylarge ta, if we assume consistency with the equilibriumcomputations in Refs. [73–75], which suggest a universalexponent α = 0.28 independent of η as mentioned above.

1 See Sec. V A for the evaluation of the effective temperature in adifferent sense.

We suspect that the deviations of our experimental re-sult 0.20 and 0.34 for the exponent α from the theoreticalequilibrium value of 0.28 are at least in part a result oftransient effects. These effects are difficult to analyzein a precise way because the effective exponent changesas a function of the coupling constant η and the anneal-ing time range, and is therefore non-universal as seen inFig. 3(b) and Fig. 1 of Ref. [76].

Noise amplitude and control errors may qualitativelyexplain the difference between the NASA and Burnabydevices. The latter is a newer, low-noise model, withlower 1/f noise amplitude and more accurate control [61].Better control in the specification of system parameters,the interaction strength J between neighboring qubits aswell as the local longitudinal field (which is nominallyzero in the present problem), results in less randomnessin the problem parameters, which should lead to a morerapid decrease of the kink density as a function of anneal-ing time on the Burnaby device, meaning a larger expo-nent α. Moreover, the less noisy Burnaby device has avalue closer to that of the closed-system value, and so itstands to reason that the fact that the spin-boson valueis intermediate between the noisier NASA device and theBurnaby device is a reflection of the fact that the formerdevice is more closely described as an open system thanthe latter. The extracted exponents of 0.34 (Burnaby)and 0.20 (NASA) are consistent with this picture.

Note that randomness in J from location to locationnecessarily induces more kinks and eventually leads to avery slow inverse-logarithmic law, instead of a polynomialdecay [40, 41, 78].

IV. KINK DISTRIBUTION

We collected statistics of the kink density from 200, 000samples for each ta at L = 800 as a test of the distributionfunction theory developed by one of us as described inSec. II B and Ref. [52]. See also Ref. [36]. One of theimportant predictions of these references is that the ratioof the qth cumulant κq (q ≥ 2) to the first cumulantκ1, the average, is independent of the annealing time ta[Eq. (4)].

Figure 5 shows the ta dependence of three cumulants[panels (a) and (b) for the NASA and Burnaby devices,respectively] and the ratios κ2/κ1 and κ3/κ1 [panels (c)and (d) for the NASA and Burnaby devices, respec-tively]. With the exception of κ3/κ1 for the NASA de-vice, these ratios indeed appear to be independent of ta,as predicted. The experimental values are extracted asκ2/κ1 ≈ 0.61 − 0.63 and κ3/κ1 ≈ 0.23 − 0.25. The the-oretical predictions are 0.586 and 0.134, respectively, sothe former ratio is closer to the theoretical predictionthan the latter. A possible reason is the large uncer-tainty in statistics as reflected in the large error bars inFigure 5(c) and (d) for κ3/κ1. Indeed, the lower ends ofthe error bars of this ratio lie around 0.1, and the theo-retically predicted value of 0.134 is within the error bars.

Page 8: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

8

100

101

100

101

102

kq

ta[ms]

q = 1 q = 2 q = 3

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

100

101

102

kq

/k1

ta[ms]

k2/k1 k3/k1

(b)

(d)

100

101

100

101

102

kq

ta

[ms]

q = 1 q = 2 q = 3

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

100

101

102

ta

[μs]

kq

/k1

k2/k1 k3/k1

(a)

(c)

FIG. 5. Cumulants of qth order κq of the kink distribution. The chain length is L = 800. Error bars indicate the 68% confidenceinterval. Power-law scaling of cumulants from the first-order κ1 to third-order κ3 as functions of the annealing time ta on (a)the DW2KQ at NASA and (b) the DW2KQ in Burnaby. Ratios κ2/κ1 and κ3/κ1 of cumulants on (c) the DW2KQ at NASAand (d) the DW2KQ in Burnaby. The solid lines are optimized fits to constants, κ2/κ1 ≈ 0.61 and κ3/κ1 ≈ 0.23 for the NASAcase, and κ2/κ1 ≈ 0.63 and κ3/κ1 ≈ 0.25 for the Burnaby case.

Apart from this subtlety, the experimental data are con-sistent with the theory presented in Ref. [52] (see alsoRefs. [36, 79]).

Since the third and higher order cumulants are muchsmaller than the second order cumulant, the distributionis well approximated by the Gaussian distribution func-tion (5). Figure 6 shows the distributions at three valuesof ta. All three cases are very well approximated by thisGaussian, as drawn in solid, dashed, and dotted lines. 2

Additional data are presented in Appendix B.It is remarkable that we find such strong agreement be-

tween the closed-system quantum theory of Ref. [52] andthe experimental results for the kink decay, cumulants,and distribution, given that the experiment is conducted

2 The distribution is close to Gaussian but is not exactly so dueto the small but non-vanishing value of the third cumulant.

on devices whose behavior is described by open-systemdynamics as discussed in the previous section. This sug-gests that these features are robust aspects of the kinkstatistics that lie beyond the KZM theory. This is thefirst time that a quantum simulator predicted a hithertounknown phenomenon. To confirm the reliability of thisresult, we have conducted extensive numerical computa-tions using iTEBD with QUAPI. The result is shown inFig. 7. It is clearly seen that the ratio κ2/κ1 is constantas a function of the annealing time and the constant valueis independent of whether or not the system is coupled tothe environment. Although it is difficult to compute cu-mulants beyond second order due to the large number ofterms that must be summed, the present result supportsthe experimental finding that the theoretical predictionin Ref. [52] holds beyond its scope of an isolated system,at least for κ2/κ1.

Page 9: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

9

0

0.05

0.1

0.15

0.2

0.25

0.3

0 5 10 15 20 25 30

P(n

)

n

ta

=1 μs

ta

=10 μs

ta

=100 μs

0

0.05

0.1

0.15

0.2

0.25

0.3

0 5 10 15 20 25 30

P(n

)

n

ta

=1 μs

ta

=10 μs

ta

=100 μs

(b)

(a)

FIG. 6. Histograms of the number of kinks observed on (a)the DW2KQ at NASA and (b) the DW2KQ in Burnaby. Thechain length is L = 800. From right to left, the annealingtimes ta are 1, 10, 100 µs, respectively. Solid, dashed, anddotted lines are the Gaussian distributions of Eq. (5).

V. TESTS OF CLASSICALITY

In this section we address whether our empirical re-sults and reasonable agreement with a quantum theoryof the KZM can also be understood using a purely classi-cal approach. We first consider a Boltzmann distributionof the kink density of the classical Ising chain, then theclassical spin-vector Monte Carlo model [80], which hasbeen successfully applied to at least partially describethe outcomes of experiments using the D-Wave devicesin past studies [8, 13, 56, 66, 67, 71, 81], and also in recenttheoretical studies of quantum annealing [82, 83].

1 10 100

ta [2ℏ/B(1)]

0.1

1

κ2/κ

1

η = 0

η = 0.5, T = 0

2−√2

FIG. 7. The cumulant ratio κ2/κ1 as a function of the anneal-ing time computed by iTEBD with QUAPI. The annealingschedule and other parameters are the same as in Fig. 3. η isthe strength of the spin-boson coupling. The solid horizontalline denotes κ2/κ1 = 2−

√2, which is the value theoretically

predicted for an isolated (closed) system.

A. Boltzmann distribution and effectivetemperature of the kink distribution

A question of significant interest, e.g., due to the po-tential for quantum-assisted classical machine learningapplications of quantum annealers as Boltzmann ma-chines [84–86], is whether the kink distribution is thermaland well described by a Boltzmann distribution. Vari-ous previous studies have found mixed results in termsof trying to fit such thermal distributions to empiricalquantum annealing data, an issue that is understood interms of the fact that the distribution freezes once quan-tum fluctuations cease at low (but non-zero) values of thetransverse field [69, 86–88]. The associated effective tem-perature is a relevant metric for quantifying how noisyone quantum annealer is relative to another. Given theresult we found in Sec. III, that the Burnaby device gen-erates fewer kinks than the NASA device, we expect theformer to exhibit a lower effective temperature. In thissection we address these issues, and provide an assess-ment of how well a trivial classical model matches theempirical kink distribution we have observed.

Let the empirical distribution be denoted by P (n; ta),where n is the number of kinks, and let Q(n;β′) denotethe Boltzmann distribution for the classical Ising modelof a chain of length L. The latter is easily shown to be:

Q (n;β′) = g(n)e−β

′E(n)

Z, (20)

where g(n) =(L−1n

)is the degeneracy of n kinks, Z =

(eβ′+e−β

′)L−1 is the partition function, E(n) = 2n+1−L

is the dimensionless energy, and β′ is:

β′ =B(1)

2

1

kBT. (21)

Page 10: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

10

We write β′ instead of β to indicate that this is a dimen-sionless effective inverse temperature reflecting all noiseeffects, not the inverse of the real physical temperature.B(1) is the device-dependent value of the Ising schedule[Eq. (1)] at the end of the anneal (at s = 1), kB is theBoltzmann constant, and T is the physical temperature.T = 12.1 mK and B(1)/2 = 6.344 GHz for the NASADW2KQ device; T = 13.5 mK and B(1)/2 = 5.930 GHzfor the DW2KQ in Burnaby.

To estimate β′, we minimize (with respect to β′) theKullback-Leibler (KL) divergence and the trace-normdistance between the experimental distribution and theBoltzmann distribution in Eq. (20). The KL divergenceDKL and the trace-norm DTN are defined by using theempirical distribution P (n; ta) and Q (n, β′) as follows:

DKL(ta) =∑n

P (n; ta) logP (n; ta)

Q (n;β′), (22a)

DTN(ta) =1

2

∑n

|P (n; ta)−Q (n;β′)| . (22b)

To obtain reliable estimates of the effective temperature,we first minimize the KL divergence to obtain the firstapproximation of the effective temperature 1/β′, becausethe KL divergence turns out to be robust against datafluctuations. The β′ thus obtained is then used as theinitial value in the effective temperature estimation basedon the trace norm.

Figure 8(a) shows the effective temperatures thus com-puted for L = 800. It is clear that, as expected, theNASA device has a larger effective temperature, 23%larger at ta = 100 µs, for example. This confirms thelower-noise aspect of the Burnaby device. The decreaseof the effective temperature is consistent with more andmore kinks being annihilated as the annealing time is in-creased (as seen in Fig. 2). Indeed, the kink density forthe Boltzmann distribution is

ρkink =1

L

∞∑n=0

nQ(n;β′) =1− 1/L

1 + e2β′ , (23)

and given that we know that ρkink ∝ ta−α, we expect 1/β′

to decrease as ta increases. Moreover, a larger α valuethen corresponds to a lower value of 1/β′. As shown inFig. 8(a), the effective temperature β′ obtained by equat-ing Eq. (23) to the D-Wave data has almost the samevalue as the effective temperature obtained by Eq. (22b).

Figure 8(b) compares the empirical data and the Boltz-mann distribution with the optimized effective temper-ature at ta = 10 µs for L = 800. Although the opti-mized Boltzmann distribution captures the gross shapeof the kink distribution, significant differences are appar-ent. This is consistent with the fact that, as alreadymentioned in Sec. IV, the actual distribution is closeto Gaussian, as predicted by the quantum KZM theory.Thus, the deviation from the purely classical Boltzmanndistribution of the kink density is to be expected. It is,however, possible that a closer agreement would be found

0

0.05

0.1

0.15

0.2

0.25

0 5 10 15 20 25

P (

n)

n

PN (n;ta)

QN (n;β’)

PB (n;ta)

QB (n;β’)

(a)

(b)

0.35

0.4

0.45

0.5

0.55

100

101

102

1/β

ta [μs]

NASA_Eq.(22b)

Burnaby_Eq.(22b)

NASA_Eq.(23)

Burnaby_Eq.(23)

FIG. 8. Effective temperatures and the optimized Boltzmanndistributions on the D-Wave device. The chain length isL = 800. (a) Effective temperature 1/β′ as a function ofthe annealing time ta. Effective temperatures are obtained byequating Eq. (23) to the D-Wave data or minimizing Eq. (22b)with respect to β′. (b) The optimized Boltzmann distribu-tions at ta = 10 µs and the best fit of the Boltzmann distri-bution. Here, PN and PB are observed distributions in theDW2KQ at NASA and Burnaby, respectively. QN and QB

are the Boltzmann distribution optimized for PN and PB, re-spectively.

with the quantum Boltzmann distribution obtained oncequantum fluctuations freeze (at s < 1), but this distribu-tion too would not be Gaussian [69]. We thus concludethat the kink distribution does not thermalize in accor-dance with equilibrium theory expectations, but is ratherbetter described by the KZM and its generalization.

Nevertheless, Fig. 8(b) indicates that the effective tem-perature obtained from our fitting procedure can serve asa reasonable proxy for quantifying the relative overall ef-fect of noise for a comparison between different quantumannealing devices.

Page 11: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

11

B. Test of a classical description by spin-vectorMonte Carlo

A much more stringent model than a simple Boltz-mann distribution is the standard classical model of theD-Wave devices, the spin-vector Monte Carlo (SVMC)model [80]. In our SVMC simulations, we replace the op-erators σxi and σzi by the components of a classical unitvector, sin θi and cos θi, respectively. The Hamiltonian istherefore written as: 3

H =B(sn)

2

L−1∑i=1

cos θi cos θi+1 −A(sn)

2

L∑i=1

sin θi, (24)

where sn is a parameter representing time correspondingto the number of Monte Carlo steps, n. We choose thefollowing parameterization of sn,

sn =n

t′aN0(25)

where N0 is the number of Monte Carlo steps necessaryto reproduce the kink density observed in the D-Wave de-vice at 1 µs. The dimensionless parameter t′a correspondsto the total annealing time, such that the total numberof Monte Carlo steps is n = N0t

′a, and sn = 1 at the

end of a simulation. In the present case, N0 is 1000 and1500 for the NASA and Burnaby devices, respectively.We use the actual NASA and Burnaby annealing sched-ules depicted in Fig. 1 for comparison of the DW2KQdata with SVMC results. We first set all local angles toθi = π/2 and use a Metropolis move with the physicaltemperature of each device, T = 12.1 mK for NASA andT = 13.5 mK for Burnaby, and sequentially update eachlocal state θi. After the dynamical evolution from s = 0to s = 1, we project the final state to +1 if 0 < θi < π/2,and −1 if π/2 < θi < π. We take 1, 000 samples for eacht′a for statistical analysis.

Figure 9 shows the kink density ρkink as a function of t′aas obtained from the SVMC simulations. The power-lawscaling seen for the D-Wave data in Fig. 2 and predictedfrom the KZM theory is observed here as well, but onlyfor short annealing times4. For longer annealing times,the power law breaks down and a more rapid decay ofthe kink-density sets in, with the crossover point increas-ing with chain length L. This is not the case for theDW2KQ results (see Appendix B). Moreover, the expo-nents extracted from the power-law regions, summarizedin Table II, deviate substantially from the DW2KQ ex-ponents summarized in Table I. See also Fig. 4. It isalso shown in Appendix C that the critical exponent ν

3 The coefficient of the second term is −A(sn)/2, not A(sn)/2,following the convention of SVMC [80].

4 The unit of time in SVMC is arbitrary and we should not directlycompare the data for the same values on the horizontal axes inFigs. 2 and 9.

TABLE II. Results from SVMC model simulations for theexponent α of the power-law scaling describing the decay rateof the kink density as shown in Fig. 9. Each exponent isobtained from a fit up to the L-dependent crossover pointseen in Fig. 9.

L NASA Burnaby50 1.891± 0.158 2.218±0.236200 0.580± 0.018 0.618±0.021500 0.496± 0.008 0.506±0.007800 0.477± 0.005 0.482±0.006

TABLE III. Results from SVMC model simulations at 50 mKfor the exponent α of the power-law scaling describing thedecay rate of the kink density as shown in Fig. 16. Each ex-ponent is obtained from a fit up to the L-dependent crossoverpoint seen in Fig. 16 of Appendix D.

L NASA Burnaby50 1.472± 0.134 1.209±0.084200 0.525± 0.012 0.528±0.011500 0.450± 0.003 0.455±0.004800 0.437± 0.002 0.441±0.003

assumes the value 1/2 in the SVMC model in contrast tothe corresponding quantum value of ν = 1 for an isolatedsystem and ν = 0.66 for a system coupled to a bosonicenvironment [73]. This implies that the closeness of thedecay exponent α of the classical SVMC model to thequantum closed-system value of 0.5 is likely to be acci-dental.

Another noticeable difference is that the kink den-sity curves are all quite close, i.e. size-independent, forL ≥ 200 for the DW2KQ results, whereas the correspond-ing curves tend to differ for the SVMC simulations, withthe kink density decaying more slowly for larger chainlengths.

A further test is provided by the cumulants, shownfor SVMC in Fig. 10. The contrast with the DW2KQdata shown in Fig. 5 is clear, with the constancy of theratio seen there, as predicted from generalized KZM the-ory, weaker in Fig. 10. We furthermore provide fits tothe Gaussian distribution predicted by this theory to theSVMC simulation results in Fig. 11. Given the small-ness of the third order cumulants, the Gaussian fits areunsurprisingly quite good, though not as good as to theDW2KQ data, shown in Fig. 6.

Additional results for SVMC at the higher (and hencemore classical) simulation temperature of 50 mK are pro-vided in Appendix D. The overall trends are the same asthose seen in Figs. 9-11, but the agreement with the pre-dictions of generalized KZM theory is in fact closer thanfor the lower temperature simulations above. In particu-lar, agreement with the power-law decay predictions forthe first cumulant extend to larger t′a values, as does theconstancy of the cumulant ratios. The extracted decayexponent α is listed in Table III.

Page 12: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

12

(a) (b)

10-4

10-3

10-2

100

101

102

103

rkink

t'a

L = 50

L = 200

L = 500

L = 800

10-4

10-3

10-2

100

101

102

103

rkink

t'a

L = 50

L = 200

L = 500

L = 800

FIG. 9. Numerically computed kink density ρkink as a function of the dimensionless annealing time t′a from the SVMCmodel (log-log scale). The error bars indicate the 68% confidence interval. (a) SVMC simulation results using the NASADW2KQ annealing schedule at T = 12.1 mK. (b) SVMC simulation results using the Burnaby DW2KQ annealing schedule atT = 13.5 mK. Dashed straight lines are linear fits to a power-law decay. For L = 50, the fitting range is limited from t′a = 5 tot′a = 20. Contrast with the DW2KQ results shown in Fig. 2.

The exponent α = 0.44 is closer to the experimen-tal value 0.20 (NASA)/0.34(Burnaby) than the lower-temperature exponent α = 0.48 is, but the quantum-theoretical prediction 0.28 by the KZM is much closer tothe experimental result. One noteworthy qualitative dif-ference is that for sufficiently large annealing times, theSVMC kink density deviates downward from the power-law (fewer kinks), while in the D-Wave case it deviatesupward (more kinks).

Finally, we also analyzed the D-Wave and SVMC databy computing their trace-norm distance from the Boltz-mann distribution. Here the goal is to test the predictionof the adiabatic theorem for open quantum systems, thatthis distance decreases following a power law as a func-tion of time ta for sufficiently large ta [89]. Figures 12(a) and (b) show the result for L = 800. We computedthe trace-norm distance according to Eq. (22b), whichuses the optimized Boltzmann distribution Q (n;β′) inEq. (20), and the kink distributions of D-Wave andthe SVMC simulations as empirical distributions P (n).We fixed β′ of Q (n;β′) to that already obtained atta = 100 µs (D-Wave) or t′a = 100 (SVMC) because weare interested in how the trace-norm distance approachesthe Boltzmann distribution at this annealing time, thelargest reliable value available to us (see Appendix A).It is seen from Fig. 12 that the decrease of the trace-norm distance of SVMC fits an exponential (solid line),while the behavior of the D-Wave is closer to a powerlaw for sufficiently large ta although the difference is notlarge. Thus it is reasonable to conclude that the D-Waveresults are in closer agreement with the adiabatic theo-rem for open quantum systems than the classical SVMCsimulation results.

Given all the discrepancies we have found, it is reason-

able to conclude that the SVMC model does not explainthe behavior of the DW2KQ devices reported here.

VI. DISCUSSION

We have reported on extensive experiments for the one-dimensional transverse-field Ising model performed usingthe NASA and low-noise Burnaby DW2KQ devices. Wedemonstrated that the kink density decays in a powerlaw t−αa with the annealing time ta, in qualitative agree-ment with the theoretical prediction of the Kibble-Zurekmechanism (KZM). In more detail, we found that the ex-ponent α describing the rate of power-law decay differsfrom the KZM prediction derived under the assumptionof an isolated, closed quantum system. The differencebetween the theoretical value of α = 0.5 (which, coinci-dentally, is close to the outcome of the classical SVMCmodel as well), and the empirical values of α = 0.20(NASA) and α = 0.34 (Burnaby), can be understood toa first approximation by modeling the coupling of thesystem to a bosonic environment with an Ohmic spectraldensity, which reduces the theoretical value to α = 0.28.Although it is difficult to quantitatively explain the re-maining discrepancy, it is reasonable to suppose that itoriginates in other factors, which lead to a non-universalvalue of the exponent, such as parameter control errors,and transient effects due to short annealing times. In-deed, the larger exponent (a faster decrease of the kinkdensity) of the Burnaby device than the NASA deviceis consistent with the lower-noise characteristics of theformer.

The lower noise aspect of the Burnaby device was fur-ther verified in our study by computing the effective tem-

Page 13: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

13

(b)

(d)

(a)

(c)

10-1

100

101

100

101

102

103

kq

t'a

q = 1 q = 2 q = 3

10-1

100

101

100

101

102

103

kq

t'a

q = 1 q = 2 q = 3

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

100

101

102

103

t'a

k2/k1 k3/k1

kq

/k1

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

100

101

102

103

t'a

kq

/k1

k2/k1 k3/k1

FIG. 10. Numerically computed cumulants of qth order κq of the kink distribution. The chain length is L = 800. Error barsindicate the 68% confidence interval. Solid lines are power-law fits of the cumulants from the first-order κ1 to third-order κ3

as functions of the dimensionless annealing time t′a with the annealing schedule of (a) the DW2KQ at NASA at T = 12.1mKand (b) the DW2KQ in Burnaby at T = 13.5mK. Ratios κ2/κ1 and κ3/κ1 of cumulants for the annealing schedules of (c) theDW2KQ at NASA and (d) the DW2KQ in Burnaby. Solid lines are fits to constants for the power-law decay region of thecumulants. Contrast with the DW2KQ results shown in Fig. 5.

perature of the best-fit Boltzmann distribution at the endof the anneal, which shows that the effective temperatureis about 23% higher on the NASA device. Note that thiseffective temperature reflects the combined effects of thedilution refrigerator temperature (which is in fact slightlyhigher for the Burnaby device) and a wide range of othernoise sources including coupling to the environment andcontrol errors.

Related work was reported by Gardas et al. [50] on theprevious generation D-Wave 2X devices, at Los AlamosNational Laboratory and in Burnaby. They also founda power-law decay of the kink density but the value ofthe exponent α depended strongly on experimental con-ditions such as the choice of the device and the sign ofinteractions (ferromagnetic or antiferromagnetic). Thevalues reported range from α = 0.24 to α = 1.31, andthey left the explanation for further work after listingseveral possible options. In contrast, our work quite

definitively establishes that quantum simulation usingthe newer DW2KQ devices is capable of demonstratingand probing the KZM and its generalization, in par-ticular using the lower-noise version in Burnaby withthe DW 2000Q 5 solver.

Other closely related work is the recent experimentprobing the two-dimensional transverse-field Ising modelon the square lattice by Weinberg et al. [51], whichdemonstrated non-monotonicity in the kink density as afunction of the annealing time ta. In the short annealingtime regime, the kink density decreases as a function ofta, as in our case. The value of the exponent they foundin this shorter time range, α = 0.74, is close to the the-oretical value for an isolated system in two dimensions,α = 0.77. In contrast, for long annealing times, the kinkdensity increases as ta increases. This kind of behav-ior is often referred to as anti-Kibble-Zurek scaling andcan result from environmentally-induced heating [48, 49].

Page 14: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

14

(b)

(a)

0

0.1

0.2

0.3

0.4

0.5

0 5 10 15 20 25 30

P(n

)

n

t'a

=1

t'a

=10

t'a

=100

0

0.1

0.2

0.3

0.4

0.5

0 5 10 15 20 25 30

P(n

)

n

t'a

=1

t'a

=10

t'a

=100

FIG. 11. Histograms of the number of kinks computed us-ing the SVMC model with the annealing schedule of (a) theDW2KQ at NASA at T = 12.1mK and (b) the DW2KQ inBurnaby at T = 13.5mK. The chain length is L = 800. Fromright to left, the annealing times t′a are 1, 10, 100 µs, respec-tively. Solid, dashed, and dotted lines are the Gaussian distri-butions of Eq. (5).Contrast with the DW2KQ results shownin Fig. 6.

Weinberg et al. also attribute this latter behavior tothe effects of noise, including thermal fluctuations. In-deed, numerical calculations for the one-dimensional sys-tem shown in Fig. 3 and presented in Ref. [51] as well asin Refs. [46, 76] show that the kink density can be non-monotonic if the temperature is finite. A possible reasonthat our experiment did not find such non-monotonicityin the range ta ≤ 100 µs is that this annealing time istoo short for the temperature effects to appear in the one-dimensional problem. Our data in the range of very longannealing times up to 2000 µs show non-monotonicity insome cases and possibly reflect thermal and other devia-tions from the ideal quantum simulation, as discussed in

10-1

100

100

101

102

DT

N

ta (D-Wave) or t'

a (SVMC)

D-Wave at NASA

SVMC

10-1

100

100

101

102

DT

N

ta (D-Wave) or t'

a (SVMC)

D-Wave at Burnaby

SVMC

(a)

(b)

FIG. 12. Trace-norm distance between the kink density ofthe D-Wave device or SVMC simulations and the Boltzmanndistribution with the same β′, chosen such that the distancesare minimized at ta = 100 µs. The chain length is L = 800.The solid black lines are exponentials, D ∝ e−γta , fitted tothe SVMC data. The dotted lines are polynomials, D ∝ t−τa ,fitted to the D-Wave data for ta ∈ [30, 100] µs using (a) theNASA and (b) Burnaby annealing schedules.

Appendix A. We excluded this time range from our anal-ysis since the data appears unstable with large uncer-tainties. In the short time region of the two-dimensionalexperiment of Ref. [51] where the KZM is likely to apply,the system seems to be much less susceptible to noiseand the exponent α is close to the theoretical value of anideal, isolated system as mentioned above. These obser-vations suggest that how noise affects the system behav-ior strongly depends on the problem type as well as onthe annealing time range.

We further investigated the distribution of the kinkdensity at the end of the anneal, which encodes signa-tures of universality beyond the original predictions ofKZM theory. We found agreement with the theoreticalprediction that the ratio of the second and higher or-

Page 15: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

15

der cumulants κq (q ≥ 2) to the first order cumulantκ1 is independent of the annealing time ta [52]. We alsofound very good agreement with the prediction of a Gaus-sian distribution of the kink density. This agreementwith a quantum theory constructed for a closed, isolatedsystem suggests that these are robust features that re-main largely intact even in the presence of coupling toan environment. Extensive numerical computations us-ing iTEBD with QUAPI support this conclusion. In theclassical KZM, it is indeed the case that the constancyof cumulant ratios is a robust feature [79], but a gener-alization of this classical theory to the quantum case isnon-trivial.

Given the history of challenges via classical models toexperiments involving the D-Wave devices [80, 90, 91],and their rebuttals [65–68, 71, 92–94], we tested whetherclassical models alternatively explain the experimentaldata as well. We first tried a simple fit to a Boltzmanndistribution but did not find satisfactory agreement. Wealso tried the standard classical limit of the D-Wave de-vice, the spin-vector Monte Carlo (SVMC) model [80]and found that the quantum theory of the generalizedKZM provides better qualitative as well as quantitativeagreement.

The two DW2KQ devices we tested therefore serve toa good approximation as quantum simulators of the one-dimensional transverse-field Ising model under the influ-ence of a dephasing Ohmic bosonic environment. Thesebosons do not represent thermal effects because we haveobserved neither an approach of the kink density to aconstant, nor a non-monotonic behavior of the kink den-sity as a function of annealing time (over the annealingtime range where we have confidence in the reliabilityof the data). Instead, the bosons possibly correspondto dynamical fluctuations of the normal current flowingthrough Josephson junctions [95]. It is a difficult but in-teresting future direction of work to identify the natureof these fluctuations and to try to find a way to reducethem for better agreement with the closed quantum sys-tem limit.

VII. CONCLUSION

In the wake of a phase transition, topological defectsform. The Kibble-Zurek mechanism predicts that thedensity of defects scales as a universal power-law withthe time scale used to cross the transition. We haveshown that this prediction can be tested on quantumannealers by using them for analog quantum simulation,i.e., as a test-bed for non-equilibrium statistical mechan-ics. Specifically, our work has tested the Kibble-Zurekmechanism in the one-dimensional transverse-field Isingmodel. Our analysis of the quantum annealer data showsthat the behavior of the devices is consistent with theimplementation of this model coupled to a bosonic envi-ronment. Our work thus provides experimental evidenceof universal Kibble-Zurek scaling in an open quantum

system.

By probing the full counting statistics of topologicaldefects (kinks), we furthermore established signatures ofuniversality beyond the original prediction of the Kibble-Zurek scaling, which is focused on the average kink num-ber. In particular, we found that the power-law scalingwith the annealing time of the average kink number isshared by its variance and the third cumulant. Our ex-perimental and numerical results thus indicate that theuniversal scaling recently predicted for the cumulants ofthe kink number distribution in an isolated quantum crit-ical system also holds under open-system quantum dy-namics,

ACKNOWLEDGMENTS

The research of YB, DL, and HN is based upon worksupported by the Office of the Director of National Intelli-gence (ODNI), Intelligence Advanced Research ProjectsActivity (IARPA), via the U.S. Army Research Officecontract W911NF-17-C-0050. The views and conclusionscontained herein are those of the authors and shouldnot be interpreted as necessarily representing the officialpolicies or endorsements, either expressed or implied, ofthe ODNI, IARPA, or the U.S. Government. The U.S.Government is authorized to reproduce and distributereprints for Governmental purposes notwithstanding anycopyright annotation thereon.

Appendix A: Dependence of the kink density on theinteraction type and annealing time

Figure 13 shows the kink density obtained for threetypes of interactions up to ta = 2000 µs on the NASADW2KQ device. Figure 13(a) displays the case of fer-romagnetic interactions (J = −1), (b) antiferromagnetic(J = 1), and (c) random gauge. In the latter case we firstset all the interactions to be ferromagnetic, randomlychoose L/2 qubits, and change the sign of their interac-tions with their two nearest neighbors. In one dimensionwith free boundaries, these three cases are theoreticallyequivalent and the kink density should behave identically.

Figures 13(a), (b), and (c) clearly indicate that the dif-ferences are small for the short time regime ta ≤ 100 µs.Beyond this regime, marked deviations emerge in theferromagnetic case, whereas the antiferromagnetic andrandom gauge data remain close even up to the longestannealing time ta = 2000 µs. This difference may im-ply the presence of a systematic bias toward ferromag-netic states in the D-Wave devices, which becomes promi-nent at larger annealing times. Antiferromagnetic andrandom-gauge interactions may be interpreted to havecaused cancellations of such a bias. For these reasons wechoose to use the data from antiferromagnetic interac-tions in the main text.

Page 16: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

16

10-4

10-3

10-2

100

101

102

103

rkink

ta

[ms]

L = 50

L = 200

L = 500

L = 800

10-4

10-3

10-2

100

101

102

103

rkink

ta

[ms]

L = 50

L = 200

L = 500

L = 800

10-4

10-3

10-2

100

101

102

103

rkink

ta

[ms]

L = 50

L = 200

L = 500

L = 800

(a) (b) (c)

FIG. 13. Kink density as a function of the annealing time with different interactions on the DW2KQ at NASA. The error barsindicate the 68% confidence interval. (a) Ferromagnetic interactions (J = −1); (b) Antiferromagnetic interactions (J = 1); (c)Random gauge, where we randomly select L/2 sites i and flip the sign of interactions between qubits i− 1, i and i, i+ 1.

(a) (b)

10-2

100

101

102

103

rkink

ta

[ms]

10-3

10-2

100

101

102

103

rkink

ta

[ms]

FIG. 14. Kink density for L = 800 as a function of the annealing time from ta = 1 µs to ta = 2000 µs on (a) the NASADW2KQ device and (b) the Burnaby DW2KQ device. In both cases we used antiferromagnetic interactions. The error barsindicate the 68% confidence interval. The dashed lines are power law fits to the data up to 100 µs.

Additional data displayed in Fig. 14 show that thekink density obeys a power-law decay very accuratelyup to ta = 100 µs on both the NASA and the BurnabyDW2KQ devices with antiferromagnetic interactions andchain length L = 800. Beyond ta = 100 µs, deviationsfrom a power law become apparent. This data set, alongwith Fig. 13, motivated us to use the empirical data onlyfor ta ≤ 100 µs, in order to eliminate artifacts other thanthe coupling to bosonic environment.

Appendix B: Kink distribution for different chainlengths

Figure 15 supplements Fig. 6 by showing histograms ofthe kink density for different chain lengths on the two dif-ferent devices, (a) NASA and (b) Burnaby, at ta = 10µs.Dashed lines are the Gaussian distribution of Eq. (5).The data are well described by this distribution in allcases.

Appendix C: Critical exponent ν for the SVMCmodel

We show in this Appendix that the critical exponentν is 1/2 for the one-dimensional ferromagnetic SVMCmodel. The Hamiltonian is

H = −JN∑j=1

sin θj sin θj+1 − Γ

N∑j=1

cos θj . (C1)

We have exchanged sin θj and cos θj from the conven-tional notation of Eq. (24) for later convenience. A peri-odic boundary is assumed.

Since we are interested in how the system behaves aswe decrease Γ from a very large value (where the systemis in the paramagnetic phase with θj ≈ 0 ∀j) toward atransition point, it is reasonable to expand the Hamilto-

Page 17: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

17

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20

P(n

)

n

ta

= 10 [μs]

0

0.1

0.2

0.3

0.4

0.5

0 5 10 15 20

n

ta

= 10 [μs]

0

0.05

0.1

0.15

0.2

0.25

0.3

0 5 10 15 20

n

ta

= 10 [μs]

0

0.05

0.1

0.15

0.2

0.25

0.3

0 5 10 15 20

n

ta

=10 [μs]

L = 50 L = 800L = 500L = 200

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20

P(n

)

n

ta

= 10 [μs]

0

0.1

0.2

0.3

0.4

0.5

0 5 10 15 20

n

ta

= 10 [μs]

0

0.05

0.1

0.15

0.2

0.25

0.3

0 5 10 15 20

n

ta

= 10 [μs]

0

0.05

0.1

0.15

0.2

0.25

0.3

0 5 10 15 20

n

ta

=10 [μs]

L = 50 L = 800L = 500L = 200

(a)

(b)

FIG. 15. Kink distribution at ta = 10 µs for different chain lengths L = 50, 200, 500, 800 (left to right) from the DW2KQdevices at (a) NASA and (b) Burnaby. The dashed lines are the theoretical prediction of Eq. (5).

nian to quadratic order as

H = −J∑j

θjθj+1 +Γ

2

∑j

θ2j , (C2)

where we have ignored a constant term. Let us check ifthe paramagnetic state is stable by Fourier transforma-tion,

θj =1

N

N∑k=1

e−i2πkj/Nφk (C3)

as

H =1

2

∑k

Ak|φk|2, Ak =1

N

(Γ− 2J cos

2πk

N

). (C4)

It is observed that φk = 0 ∀k is the stable state (groundstate) configuration for Γ > 2J , and a second-order phasetransition exists at Γc = 2J .

The correlation is given by

G(r) = 〈sin θ0 sin θr〉 ≈1

N

N∑l=1

〈θlθl+r〉

=1

N2

∑k

ei2πkr/N 〈|φk|2〉, (C5)

where the angular brackets 〈· · · 〉 stand for the statistical-mechanical average with respect to the Hamiltonian(C4). We have averaged over l using translation symme-try. Since the Hamiltonian (C4) represents independentGaussian fields, we easily find

〈|φk|2〉 =1

βAk=

kBTN

Γ− 2J cos(2πk/N). (C6)

Thus the correlation function (C5) becomes

G(r) =kBT

N

∑k

ei2πkr/N1

Γ− 2J cos(2πk/N). (C7)

The behavior as r � 1 is evaluated as

G(r) =kBT

∫ ∞0

eiyr1

Γ− 2J + Jy2dy ∝ e−r/ξ, (C8)

where

ξ =

√J

Γ− 2J. (C9)

Thus the exponent is ν = 1/2. Notice that the tem-perature is kept small but finite. If T = 0 exactly, nofluctuations exist classically and the spin configuration isfixed to θj = 0 in the paramagnetic phase.

Appendix D: SVMC results at T = 50 mK

In the main text we provided our SVMC results atthe dilution refrigerator temperatures of the NASA andBurnaby DW2KQ devices. Here we provide additionalSVMC results computed at a higher simulation temper-ature of 50 mK in Figs. 16, 17, and 18. The α valuescorresponding to Fig. 16 are reported in Table III. Thesevalues are somewhat closer to the DW2KQ values thanthose for the dilution refrigerator temperatures, and wealso note that the qualitative behavior of the kink densitycurves seen in Fig. 16 is more like the DW2KQ resultsseen in Fig. 2, in that the curves for the two largest L

Page 18: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

18

values now nearly overlap. In these respects the warmerSVMC model is a closer match to the DW2KQ data thanat the dilution refrigerator temperatures.

Page 19: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

19

(a) (b)

10-4

10-3

10-2

100

101

102

103

rkink

t'a

L = 50

L = 200

L = 500

L = 80010-4

10-3

10-2

100

101

102

103

rkink

t'a

L = 50

L = 200

L = 500

L = 800

FIG. 16. Same as Fig. 9 but for T = 50 mK: numerically computed kink density ρkink as a function of the dimensionlessannealing time t′a from the SVMC model (log-log scale). The error bars indicate the 68% confidence interval. (a) SVMCsimulation results using the NASA DW2KQ annealing schedule. (b) SVMC simulation results using the Burnaby DW2KQannealing schedule. Dashed straight lines are linear fits to a power-law decay. For L = 50, the fitting range is limited fromt′a = 5 to t′a = 20.

(b)

(d)

(a)

(c)

10-1

100

101

100

101

102

103

kq

t'a

q = 1 q = 2 q = 3

10-1

100

101

100

101

102

103

kq

t'a

q = 1 q = 2 q = 3

0

0.2

0.4

0.6

0.8

1

100

101

102

103

t'a

k2/k1 k3/k1

kq

/k1

0

0.2

0.4

0.6

0.8

1

100

101

102

103

t'a

kq

/k1

k2/k1 k3/k1

FIG. 17. Same as Fig. 10 but for T = 50 mK: numerically computed cumulants of qth order κq of the kink distribution. Thechain length is L = 800. Error bars indicate the 68% confidence interval. Solid lines are power-law fits of the cumulants fromthe first-order κ1 to third-order κ3 as functions of the dimensionless annealing time t′a with the annealing schedule of (a) theDW2KQ at NASA and (b) the DW2KQ in Burnaby. Ratios κ2/κ1 and κ3/κ1 of cumulants for the annealing schedules of (c)the DW2KQ at NASA and (d) the DW2KQ in Burnaby. Solid lines are fits to constants for the power-law decay region of thecumulants.

Page 20: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

20

(b)

(a)

0

0.1

0.2

0.3

0.4

0.5

0 5 10 15 20 25 30

P(n

)

n

t'a

=1

t'a

=10

t'a

=100

0

0.1

0.2

0.3

0.4

0.5

0 5 10 15 20 25 30

P(n

)

n

t'a

=1

t'a

=10

t'a

=100

FIG. 18. Same as Fig. 11 but for T = 50 mK: histogramsof the number of kinks computed using the SVMC modelwith the annealing schedule of (a) the DW2KQ at NASA and(b) the DW2KQ in Burnaby. The chain length is L = 800.From right to left, the annealing times t′a are 1, 10, 100 µs,respectively. Solid, dashed, and dotted lines are the Gaussiandistributions of Eq. (5).

Page 21: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

21

[1] T. Kadowaki and H. Nishimori, Phys. Rev. E 58, 5355(1998).

[2] A. Das and B. K. Chakrabarti, Rev. Mod. Phys. 80, 1061(2008).

[3] T. Albash and D. A. Lidar, Rev. Mod. Phys. 90, 015002(2018).

[4] P. Hauke, H. G. Katzgraber, W. Lechner, H. Nishimori,and W. D. Oliver, arXiv:1903.06559 (2019).

[5] S. Boixo, T. F. Ronnow, S. V. Isakov, Z. Wang,D. Wecker, D. A. Lidar, J. M. Martinis, and M. Troyer,Nat. Phys. 10, 218 (2014).

[6] T. F. Rønnow, Z. Wang, J. Job, S. Boixo, S. V. Isakov,D. Wecker, J. M. Martinis, D. A. Lidar, and M. Troyer,Science 345, 420 (2014).

[7] I. Hen, J. Job, T. Albash, T. F. Ronnow, M. Troyer, andD. A. Lidar, Phys. Rev. A 92, 042325 (2015).

[8] J. King, S. Yarkoni, J. Raymond, I. Ozfidan, A. D.King, M. M. Nevisi, J. P. Hilton, and C. C. McGeoch,arXiv:1701.04579 (2017).

[9] J. King, S. Yarkoni, M. M. Nevisi, J. P. Hilton, and C. C.McGeoch, arXiv:1508.05087 (2015).

[10] V. S. Denchev, S. Boixo, S. V. Isakov, N. Ding, R. Bab-bush, V. Smelyanskiy, J. Martinis, and H. Neven, Phys.Rev. X 6, 031015 (2016).

[11] S. Mandra, Z. Zhu, W. Wang, A. Perdomo-Ortiz, andH. G. Katzgraber, Phys. Rev. A 94, 022337 (2016).

[12] W. Vinci and D. A. Lidar, Phys. Rev. Applied 6, 054016(2016).

[13] T. Albash and D. A. Lidar, Phys. Rev. X 8, 031016(2018).

[14] S. Mandra and H. G. Katzgraber, Quantum Sci. Technol.3, 04LT01 (2018).

[15] A. Pearson, A. Mishra, I. Hen, and D. A. Lidar, npjQuant. Inf. 5, 107 (2019).

[16] A. D. King, J. Carrasquilla, J. Raymond, I. Ozfidan,E. Andriyash, A. Berkley, M. Reis, T. Lanting, R. Har-ris, F. Altomare, K. Boothby, P. I. Bunyk, C. En-derud, A. Frechette, E. Hoskinson, N. Ladizinsky, T. Oh,G. Poulin-Lamarre, C. Rich, Y. Sato, A. Y. Smirnov,L. J. Swenson, M. H. Volkmann, J. Whittaker, J. Yao,E. Ladizinsky, M. W. Johnson, J. Hilton, and M. H.Amin, Nature 560, 456 (2018).

[17] A. D. King, J. Raymond, T. Lanting, S. V.Isakov, M. Mohseni, G. Poulin-Lamarre, S. Ejtemaee,W. Bernoudy, I. Ozfidan, A. Y. Smirnov, M. Reis,F. Altomare, M. Babcock, C. Baron, A. J. Berkley,K. Boothby, P. I. Bunyk, H. Christiani, C. Enderud,B. Evert, R. Harris, E. Hoskinson, S. Huang, K. Jooya,A. Khodabandelou, N. Ladizinsky, R. Li, P. A. Lott,A. J. R. MacDonald, D. Marsden, G. Marsden, T. Med-ina, R. Molavi, R. Neufeld, M. Norouzpour, T. Oh,I. Pavlov, I. Perminov, T. Prescott, C. Rich, Y. Sato,B. Sheldan, G. Sterling, L. J. Swenson, N. Tsai, M. H.Volkmann, J. D. Whittaker, W. Wilkinson, J. Yao,H. Neven, J. P. Hilton, E. Ladizinsky, M. W. Johnson,and M. H. Amin, arXiv:1911.03446 (2019).

[18] R. Harris, Y. Sato, A. J. Berkley, M. Reis, F. Al-tomare, M. H. Amin, K. Boothby, P. Bunyk, C. Deng,C. Enderud, S. Huang, E. Hoskinson, M. W. Johnson,E. Ladizinsky, N. Ladizinsky, T. Lanting, R. Li, T. Med-ina, R. Molavi, R. Neufeld, T. Oh, I. Pavlov, I. Perminov,

G. Poulin-Lamarre, C. Rich, A. Smirnov, L. Swenson,N. Tsai, M. Volkmann, J. Whittaker, and J. Yao, Sci-ence 361, 162 (2018).

[19] R. Islam, C. Senko, W. C. Campbell, S. Korenblit,J. Smith, A. Lee, E. E. Edwards, C. C. J. Wang, J. K.Freericks, and C. Monroe, Science 340, 583 (2013).

[20] J. Smith, A. Lee, P. Richerme, B. Neyenhuis, P. W. Hess,P. Hauke, M. Heyl, D. A. Huse, and C. Monroe, Nat.Phys. 12, 907 (2016).

[21] J. Zhang, G. Pagano, P. W. Hess, A. Kyprianidis,P. Becker, H. Kaplan, A. V. Gorshkov, Z. X. Gong, andC. Monroe, Nature 551, 601 (2017).

[22] T. W. Kibble, J. Phys. A: Math. Gen. 9, 1387 (1976).[23] W. H. Zurek, Nature 317, 505 (1985).[24] We use both terms interchangeably.[25] A. Polkovnikov, Phys. Rev. B 72, 161201(R) (2005).[26] B. Damski, Phys. Rev. Lett. 95, 035701 (2005).[27] J. Dziarmaga, Phys. Rev. Lett. 95, 245701 (2005).[28] W. H. Zurek, U. Dorner, and P. Zoller, Phys. Rev. Lett.

95, 105701 (2005).[29] A. Polkovnikov, K. Sengupta, A. Silva, and M. Vengalat-

tore, Rev. Mod. Phys. 83, 863 (2011).[30] A. del Campo and W. H. Zurek, Int. J. Mod. Phys. A

29, 1430018 (2014).[31] M. Uhlmann, R. Schutzhold, and U. R. Fischer, Phys.

Rev. Lett. 99, 120407 (2007).[32] M. Uhlmann, R. Schutzhold, and U. R. Fischer, Phys.

Rev. D 81, 025017 (2010).[33] X.-Y. Xu, Y.-J. Han, K. Sun, J.-S. Xu, J.-S. Tang, C.-F.

Li, and G.-C. Guo, Phys. Rev. Lett. 112, 035701 (2014).[34] J.-M. Cui, Y.-F. Huang, Z. Wang, D.-Y. Cao, J. Wang,

W.-M. Lv, L. Luo, A. del Campo, Y.-J. Han, C.-F. Li,and G.-C. Guo, Sci. Rep. 6, 33381 (2016).

[35] M. Gong, X. Wen, G. Sun, D.-W. Zhang, D. Lan,Y. Zhou, Y. Fan, Y. Liu, X. Tan, H. Yu, Y. Yu, S.-L.Zhu, S. Han, and P. Wu, Sci. Rep. 6, 22667 (2016).

[36] J.-M. Cui, F. J. Gomez-Ruiz, Y.-F. Huang, C.-F. Li, G.-C. Guo, and A. del Campo, Communications Physics 3,44 (2020).

[37] H. Bernien, S. Schwartz, A. Keesling, H. Levine, A. Om-ran, H. Pichler, S. Choi, A. S. Zibrov, M. Endres,M. Greiner, V. Vuletic, and M. D. Lukin, Nature 551,579 (2017).

[38] D. Sen, K. Sengupta, and S. Mondal, Phys. Rev. Lett.101, 016806 (2008).

[39] R. Barankov and A. Polkovnikov, Phys. Rev. Lett. 101,076801 (2008).

[40] J. Dziarmaga, Phys. Rev. B 74, 064416 (2006).[41] T. Caneva, R. Fazio, and G. E. Santoro, Phys. Rev. B

76, 144427 (2007).[42] W. H. Zurek and U. Dorner, Phil. Trans. R. Soc. A 366,

2953 (2008).[43] J. Dziarmaga and M. M. Rams, New J. Phys. 12, 055007

(2010).[44] A. del Campo, T. W. B. Kibble, and W. H. Zurek, J.

Phys.: Cond. Matt. 25, 404210 (2013).[45] F. J. Gomez-Ruiz and A. del Campo, Phys. Rev. Lett.

122, 080604 (2019).[46] D. Patane, A. Silva, L. Amico, R. Fazio, and G. E.

Santoro, Phys. Rev. Lett. 101, 175701 (2008).

Page 22: Kibble-Zurek Mechanism and Beyond - arXiv3Department of Physics, Tohoku University, Sendai 980-8578, Japan 4Graduate School of Information Sciences, Tohoku University, Sendai 980-8579,

22

[47] P. Nalbach, S. Vishveshwara, and A. A. Clerk, Phys.Rev. B 92, 014306 (2015).

[48] A. Dutta, A. Rahmani, and A. del Campo, Phys. Rev.Lett. 117, 080402 (2016).

[49] R. Puebla, A. Smirne, S. F. Huelga, and M. B. Plenio,arXiv:1911.06023 (2019).

[50] B. Gardas, J. Dziarmaga, W. H. Zurek, and M. Zwolak,Sci. Rep. 8, 4539 (2018).

[51] P. Weinberg, M. Tylutki, J. M. Ronkko, J. Westerholm,J. A. Astrom, P. Manninen, P. Torma, and A. W. Sand-vik, Phys. Rev. Lett. 124, 090502 (2020).

[52] A. del Campo, Phys. Rev. Lett. 121, 200601 (2018).[53] V. Choi, Quant. Inf. Proc. 7, 193 (2008).[54] K. L. Pudenz, T. Albash, and D. A. Lidar, Nat. Com-

mun. 5, 3243 (2014).[55] A. Mishra, T. Albash, and D. A. Lidar, Quant. Inf. Proc.

15, 609 (2016).[56] A. Mishra, T. Albash, and D. A. Lidar, Nat. Commun.

9, 2917 (2018).[57] H.-P. Breuer and F. Petruccione,

The Theory of Open Quantum Systems (Oxford Univer-sity Press, Oxford, 2002).

[58] H. Nishimori and G. Ortiz,Elements of Phase Transitions and Critical Phenomena(Oxford, United Kingdom: Oxford University Press,2011).

[59] B. Damski and W. H. Zurek, Phys. Rev. A 73, 063405(2006).

[60] L. Cincio, J. Dziarmaga, M. M. Rams, and W. H. Zurek,Phys. Rev. A 75, 052321 (2007).

[61] QPU-Specific Physical Properties: DW 2000Q 5 , D-WaveSystems Inc., 3033 Beta Ave Burnaby, BC V5G 4M9Canada (2019).

[62] A. J. Leggett, S. Chakravarty, A. T. Dorsey, M. P. A.Fisher, A. Garg, and W. Zwerger, Rev. Mod. Phys. 59,1 (1987).

[63] T. Lanting, M. H. S. Amin, M. W. Johnson, F. Al-tomare, A. J. Berkley, S. Gildert, R. Harris, J. Johans-son, P. Bunyk, E. Ladizinsky, E. Tolkacheva, and D. V.Averin, Phys. Rev. B 83, 180502(R) (2011).

[64] T. Albash, S. Boixo, D. A. Lidar, and P. Zanardi, NewJ. Phys. 14, 123016 (2012).

[65] S. Boixo, T. Albash, F. M. Spedalieri, N. Chancellor,and D. A. Lidar, Nat. Commun. 4, 2067 (2013).

[66] T. Albash, W. Vinci, A. Mishra, P. A. Warburton, andD. A. Lidar, Phys. Rev. A 91, 042314 (2015).

[67] T. Albash, T. F. Rønnow, M. Troyer, and D. A. Lidar,Eur. Phys. J. Spec. Top. 224, 111 (2015).

[68] T. Albash, I. Hen, F. M. Spedalieri, and D. A. Lidar,Phys. Rev. A 92, 062328 (2015).

[69] M. H. Amin, Phys. Rev. A 92, 052323 (2015).[70] T. Albash and D. A. Lidar, Phys. Rev. A 91, 062320

(2015).[71] S. Boixo, V. N. Smelyanskiy, A. Shabani, S. V. Isakov,

M. Dykman, V. S. Denchev, M. H. Amin, A. Y. Smirnov,M. Mohseni, and H. Neven, Nat. Commun. 7, 10327

(2016).[72] H. Munoz-Bauza, H. Chen, and D. Lidar, npj Quant.

Inf. 5, 2 (2019).[73] P. Werner, K. Volker, M. Troyer, and S. Chakravarty,

Phys. Rev. Lett. 94, 047201 (2005).[74] S. Pankov, S. Florens, A. Georges, G. Kotliar, and

S. Sachdev, Phy. Rev. B 69, 054426 (2004).[75] S. Sachdev, P. Werner, and M. Troyer, Phys. Rev. Lett.

92, 237003 (2004).[76] S. Suzuki, H. Oshiyama, and N. Shibata, J. Phys. Soc.

Jpn. 88, 061003 (2019).[77] H. Oshiyama, S. Suzuki, and N. Shibata,

arXiv:2005.05621 (2020).[78] D. A. Huse and D. S. Fisher, Phys. Rev. Lett. 57, 2203

(1986).[79] F. J. Gomez-Ruiz, J. J. Mayo, and A. del Campo, Phys.

Rev. Lett. (accepted); arXiv:1912.04679 (2019).[80] S. W. Shin, G. Smith, J. A. Smolin, and U. Vazirani,

arXiv:1401.7087 (2014).[81] A. D. King, E. Hoskinson, T. Lanting, E. Andriyash, and

M. H. Amin, Phys. Rev. A 93, 052320 (2016).[82] Y. Susa, Y. Yamashiro, M. Yamamoto, I. Hen, D. A. Li-

dar, and H. Nishimori, Phys. Rev. A 98, 042326 (2018).[83] Y. Yamashiro, M. Ohkuwa, H. Nishimori, and D. A.

Lidar, Phys. Rev. A 100, 052321 (2019).[84] M. H. Amin, E. Andriyash, J. Rolfe, B. Kulchytskyy, and

R. Melko, Phys. Rev. X 8, 021050 (2018).[85] M. Benedetti, J. Realpe-Gomez, R. Biswas, and

A. Perdomo-Ortiz, Phys. Rev. X 7, 041052 (2017).[86] R. Y. Li, T. Albash, and D. A. Lidar, arXiv:1910.01283

(2019).[87] J. Marshall, E. G. Rieffel, and I. Hen, Phys. Rev. Applied

8, 064025 (2017).[88] W. Vinci and D. A. Lidar, Phys. Rev. A 97, 022308

(2018).[89] L. C. Venuti, T. Albash, D. A. Lidar, and P. Zanardi,

Phys. Rev. A 93, 032118 (2016).[90] J. A. Smolin and G. Smith, Frontiers in Physics 2, 52

(2014).[91] S. W. Shin, G. Smith, J. A. Smolin, and U. Vazirani,

arXiv:1404.6499 (2014).[92] L. Wang, T. F. Rønnow, S. Boixo, S. V. Isakov, Z. Wang,

D. Wecker, D. A. Lidar, J. M. Martinis, and M. Troyer,arXiv:1305.5837 (2013).

[93] T. Lanting, A. J. Przybysz, A. Y. Smirnov, F. M.Spedalieri, M. H. Amin, A. J. Berkley, R. Harris, F. Al-tomare, S. Boixo, P. Bunyk, N. Dickson, C. Enderud,J. P. Hilton, E. Hoskinson, M. W. Johnson, E. Ladizin-sky, N. Ladizinsky, R. Neufeld, T. Oh, I. Perminov,C. Rich, M. C. Thom, E. Tolkacheva, S. Uchaikin, A. B.Wilson, and G. Rose, Phys. Rev. X 4, 021041 (2014).

[94] J. Job and D. Lidar, Quantum Science and Technology3, 030501 (2018).

[95] A. Caldeira and A. Leggett, Ann. Phys. 149, 374 (1983).