Defects in Quantum Computers - arXiv · logical defects: Spins change orientation, kinks...

8
arXiv:1707.09463v2 [quant-ph] 6 Mar 2018 Defects in Quantum Computers Bartlomiej Gardas, 1, 2, 3 Jacek Dziarmaga, 3 Wojciech H. Zurek, 1 and Michael Zwolak 1 1 Theoretical Division, LANL, Los Alamos, New Mexico 87545, USA 2 Institute of Physics, University of Silesia, 40-007 Katowice, Poland 3 Instytut Fizyki Uniwersytetu Jagiello´ nskiego, ul. Lojasiewicza 11, PL-30-348 Krak´ow, Poland (Dated: March 7, 2018) The shift of interest from general purpose quan- tum computers to adiabatic quantum computing or quantum annealing calls for a broadly appli- cable and easy to implement test to assess how quantum or adiabatic is a specific hardware. Here we propose such a test based on an exactly solv- able many body system – the quantum Ising chain in transverse field – and implement it on the D- Wave machine. An ideal adiabatic quench of the quantum Ising chain should lead to an ordered broken symmetry ground state with all spins aligned in the same direction. An actual quench can be imperfect due to decoherence, noise, flaws in the implemented Hamiltonian, or simply too fast to be adiabatic. Imperfections result in topo- logical defects: Spins change orientation, kinks punctuating ordered sections of the chain. The number of such defects quantifies the extent by which the quantum computer misses the ground state, and is, therefore, imperfect. I. INTRODUCTION Adiabatic quantum computing [13] – an alternative to the quantum Turing machine paradigm – is at its core very simple and very quantum: Evolve a system from the ground state of an “easy” Hamiltonian H 0 to the ground state of H 1 that encodes the solution to the problem of interest by varying the parameter s from 0 to 1 in H (s) = (1 s)H 0 + sH 1 . (1) When H (s) varies slowly enough the system will remain in its ground state, and the answer can be “read off” through a suitable measurement of the final state. It has always been appreciated that adiabatic quan- tum computing will be difficult. For instance, even if the hardware to accurately implement H (s) and measure the final (likely, globally entangled) state were available, how slow is “slow enough” to retain the system in the ground state? This is a difficult question, as H (s) is likely to have – somewhere between H 0 to H 1 – a narrow energy gap ∆ analogous to the critical point of a quantum phase tran- sition in a finite system. The exact size and properties of such a gap are ab initio unknown. Yet, for the com- putation to succeed, this gap should be traversed slowly, on a timescale longer than /∆. Here, we put forth a simple test based on the behavior of the exactly solvable quantum Ising chain in transverse field and deploy it on the D-Wave chip. As we shall see, in addition to the issues of adiabaticity and accessibility of global ground states, there are other practical consid- erations that affect performance of D-Wave computers, and are likely to play a role in similar devices. There are several efforts that aim at such hardware [4]. The D-Wave computer is already available and is the ob- vious guinea pig that we can test. There are by now sev- eral papers that, with varying degrees of success, model the behavior of D-Wave [5]. We applaud such efforts, but aim at a rather different goal – a general TAC. The quantum Ising chain has a Hamiltonian, H = g(t) L i=1 σ x i J (t) L-1 i=1 σ z i σ z i+1 , (2) in the form of Eq. (1). It can be implemented on the D-Wave computer, see Fig. 1. When the transverse field g and coupling J vary, the ground state of the quan- tum Ising chain can undergo a transition from a non- degenerate, paramagnetic state, |··· →→→ ···〉, to a degenerate, ferromagnetic state spanned by | · · · ↑↑↑ · · · 〉 and |··· ↓↓↓ ···〉. The phase transition occurs when g = J . The ground state on the broken symmetry side is a defect-free, ferromagnetically ordered ground state. Quenches that are too fast to be adiabatic, or are in some other way imperfect, would instead lead to a “defective” FIG. 1. The quantum Ising chain implemented in a D-Wave computer. a. An example of an Ising chain on the D-Wave “chimera graph”. The red lines are active couplings between “spins”. b. A typical annealing protocol for a D-Wave annealer. Here J (s)= Jmax · j (s), where j (s) is a predetermined function increasing from j (0) = 0 to its maximal value j (1) and Jmax [-1, 1] is a free parameter that can be turned at will.

Transcript of Defects in Quantum Computers - arXiv · logical defects: Spins change orientation, kinks...

arX

iv:1

707.

0946

3v2

[qu

ant-

ph]

6 M

ar 2

018

Defects in Quantum Computers

Bart lomiej Gardas,1, 2, 3 Jacek Dziarmaga,3 Wojciech H. Zurek,1 and Michael Zwolak1

1Theoretical Division, LANL, Los Alamos, New Mexico 87545, USA2Institute of Physics, University of Silesia, 40-007 Katowice, Poland

3Instytut Fizyki Uniwersytetu Jagiellonskiego, ul. Lojasiewicza 11, PL-30-348 Krakow, Poland

(Dated: March 7, 2018)

The shift of interest from general purpose quan-

tum computers to adiabatic quantum computing

or quantum annealing calls for a broadly appli-

cable and easy to implement test to assess how

quantum or adiabatic is a specific hardware. Here

we propose such a test based on an exactly solv-

able many body system – the quantum Ising chain

in transverse field – and implement it on the D-

Wave machine. An ideal adiabatic quench of the

quantum Ising chain should lead to an ordered

broken symmetry ground state with all spins

aligned in the same direction. An actual quench

can be imperfect due to decoherence, noise, flaws

in the implemented Hamiltonian, or simply too

fast to be adiabatic. Imperfections result in topo-

logical defects: Spins change orientation, kinks

punctuating ordered sections of the chain. The

number of such defects quantifies the extent by

which the quantum computer misses the ground

state, and is, therefore, imperfect.

I. INTRODUCTION

Adiabatic quantum computing [1–3] – an alternativeto the quantum Turing machine paradigm – is at its corevery simple and very quantum: Evolve a system from theground state of an “easy” Hamiltonian H0 to the groundstate of H1 that encodes the solution to the problem ofinterest by varying the parameter s from 0 to 1 in

H(s) = (1 − s)H0 + sH1. (1)

When H(s) varies slowly enough the system will remainin its ground state, and the answer can be “read off”through a suitable measurement of the final state.

It has always been appreciated that adiabatic quan-tum computing will be difficult. For instance, even if thehardware to accurately implement H(s) and measure thefinal (likely, globally entangled) state were available, howslow is “slow enough” to retain the system in the groundstate? This is a difficult question, as H(s) is likely to have– somewhere between H0 to H1 – a narrow energy gap ∆analogous to the critical point of a quantum phase tran-sition in a finite system. The exact size and propertiesof such a gap are ab initio unknown. Yet, for the com-putation to succeed, this gap should be traversed slowly,on a timescale longer than ~/∆.

Here, we put forth a simple test based on the behaviorof the exactly solvable quantum Ising chain in transverse

field and deploy it on the D-Wave chip. As we shall see,in addition to the issues of adiabaticity and accessibilityof global ground states, there are other practical consid-erations that affect performance of D-Wave computers,and are likely to play a role in similar devices.

There are several efforts that aim at such hardware [4].The D-Wave computer is already available and is the ob-vious guinea pig that we can test. There are by now sev-eral papers that, with varying degrees of success, modelthe behavior of D-Wave [5]. We applaud such efforts, butaim at a rather different goal – a general TAC.

The quantum Ising chain has a Hamiltonian,

H = −g(t)

L∑

i=1

σxi − J(t)

L−1∑

i=1

σzi σz

i+1, (2)

in the form of Eq. (1). It can be implemented on theD-Wave computer, see Fig. 1. When the transverse fieldg and coupling J vary, the ground state of the quan-tum Ising chain can undergo a transition from a non-degenerate, paramagnetic state, | · · · →→→ · · · 〉, to adegenerate, ferromagnetic state spanned by | · · · ↑↑↑ · · · 〉and | · · · ↓↓↓ · · · 〉. The phase transition occurs wheng = J . The ground state on the broken symmetry sideis a defect-free, ferromagnetically ordered ground state.Quenches that are too fast to be adiabatic, or are in someother way imperfect, would instead lead to a “defective”

FIG. 1. The quantum Ising chain implemented ina D-Wave computer. a. An example of an Ising chainon the D-Wave “chimera graph”. The red lines are activecouplings between “spins”. b. A typical annealing protocolfor a D-Wave annealer. Here J(s) = Jmax · j(s), where j(s)is a predetermined function increasing from j(0) = 0 to itsmaximal value j(1) and Jmax ∈ [−1, 1] is a free parameterthat can be turned at will.

2

a b

0.6 0.8 1.0 1.2 1.4g/J

10-1

100

∆L/J

10-3 10-2 10-1 100J/τQ

10-3

10-1

101

kinks

L = 20

L = 50

L = 100

L = 200

FIG. 2. Adiabaticity in the quantum Ising chain a.the relative energy gap ∆L/J in Eq. (4) as a function of therelative transverse field g/J . For long chains the gap hasa minimum when g/J = 1. b. The number of kinks in achain of length L after a quench with a quench time τQ. Thedependence crosses over from the power law, Eq. (7), to theLandau-Zener formula, Eq. (5), at τAD in Eq. (6).

state with “kinks”, e.g., | · · · ↑↑↓↓ · · · 〉.

II. TEST OF ADIABATIC COMPUTING

The dynamics of quantum phase transitions was firstunderstood by analyzing the density of kinks in the fi-nal post-transition state (g = 0, J at its maximum) as afunction of the quench timescale τQ [6–8]. Near the crit-ical point, g = J ≡ Jc (see Fig. 1b and Fig. 2a), quenchis well approximeted by

g(t)

J(t)− 1 ≈

t

τQ. (3)

That analysis dealt with the limit of very long chains(L ≫ 1) [6–8] where the generation of kinks was a fore-gone conclusion. However, we are interested in relativelyshort chains where there is a chance for adiabaticity tosurvive. This is determined by the gap size, ∆L, seeFig. 2a. At the critical point, s = sc, where

∆L = Jc2π

L, (4)

the ground and first excited states (that can accommo-date a single pair of kinks) undergo an anti-crossing,where the probability of exciting a pair of kinks is givenby the Landau-Zener (LZ) formula [6, 7]

p = exp(

−2π3JcτQ/~L2)

. (5)

Thus, when τQ exceeds

τAD =~L2

2π3Jc, (6)

we expect exponential suppression of kinks, i.e., quantumannealing should lead to the “correct answer” (in thiscase, all spins pointing in the same direction).

When the condition for adiabaticity is not met, τQ ≪τAD, the quench timescale also governs the density ofexcitations according to

1

1√

2JcτQ/~(7)

for sufficiently long closed chains [8]. The scaling, Eq. (7),conforms with the Kibble-Zurek mechanism (KZM) thatrelates the density of topological defects (and, more gen-erally, excitations) to the critical exponents of phase tran-sition and the rate of the quench [9, 10].

The two regimes – LZ (τQ ≫ τAD) and KZM (τAD ≪τQ) – switch validity when τQ ∼ τAD, see Fig. 2b. A goodindication of the “border” between LZ and KZM – i.e.,between adiabatic and non-adiabatic – is the expectednumber of excitations: When it is fractional, LZ is agood approximation; When there are several, then KZMshould work.

We expect that, in hardware to implement quantumannealing, one should be able to choose g, J , L, and τQ

to cover the range where the ideal quantum Ising chainundergoes a transition from quantum adiabatic LZ be-havior (i.e., a successful computation) to non-adiabaticKZM behavior (i.e., a defective computation). Thus, thequench of the Ising chain gives a simple test of adiabaticcomputing (TAC) for devices that implement quantumannealing. There are other tests that aim at similar goals(e.g., “quench echo” [11] and the symptoms of entangle-ment [12]). The physically motivated TAC proposed herewill be useful in evaluating quantum annealing hardware.

III. RESULTS

In D-Wave computers, L can vary from L = 2 toL ∼ 103 and τQ by over two orders of magnitude. More-over, the maximal value of J at the beginning (and theend) of the quench, respectively, can vary by about twoorders of magnitude. We have implemented the quenchon both the DW2X-SYS4 (based in Burnaby) and theDW2X (based in Los Alamos), as shown in Fig. 1. Thenumber of kinks in long chains as a function of quenchtime from the Los Alamos D-Wave DW2X are shown inFig. 3a (see Methods for details and a compilation ofresults from Burnaby and Los Alamos).

There are several striking and general features ofFig. 3. The plots conform well to a power law with thedensity of kinks proportional to τ−1

Q . This power law dif-

fers from the KZM prediction of τ−1/2

Q . Indeed, all ofthese plots enter the regime where the number of kinksper chain is ∼ 0.1 or less. In this range one expects expo-nential LZ suppression of excitations, though. We havenot found any evidence of such an effect.

Since we do not see the exponential suppression in ei-ther open or closed chains of many different lengths from50 to 500 sites, we search for it in very short closed chains,L = 4, that exhibit LZ crossover for a finite g. Additional

3

a b

10-3 10-1 101

τ−1

Q [MHz]

10-5

10-3

10-1

101kinks

L = 50

A = 0.039 ± 0.001

x = 1.04 ± 0.04

10-3 10-1 101

τ−1

Q [MHz]

10-5

10-3

10-1

101

kinks

L = 100

A = 0.067 ± 0.002

x = 1.2 ± 0.06

10-3 10-1 101

τ−1

Q [MHz]

10-8

10-6

10-4

kinks

Q = 0.73 ± 0.04

Jmax = −1.0

10-3 10-1 101

τ−1

Q [MHz]

10-8

10-6

10-4

kinks

Q = 1.14 ± 0.07

Jmax = −0.99

10-3 10-1 101

τ−1

Q [MHz]

10-5

10-3

10-1

101

kinks

L = 300

A = 0.159 ± 0.002

x = 1.31 ± 0.02

10-3 10-1 101

τ−1

Q [MHz]

10-5

10-3

10-1

101kinks

L = 500

A = 0.381 ± 0.003

x = 1.22 ± 0.02

10-3 10-1 101

τ−1

Q [MHz]

10-8

10-6

10-4

kinks

Q = 2.82 ± 0.32

Jmax = −0.97

10-3 10-1 101

τ−1

Q [MHz]

10-8

10-6

10-4

kinks

Q = 4.93 ± 1.33

Jmax = −0.92

FIG. 3. Defect generation in a quench of the quantum Ising chain on a D-Wave computer. a. The number ofdefects versus quench time for two different length chains (Jmax = −1 for all) on the Los Alamos machine. Solid line shows thebest fit to the function A · τ −x

Q . b. The number of defects versus quench time for a short chain (L = 4) for different values of

Jmax. Solid line shows the best fit to the function A · τ −1

Q . These results were obtained by averaging over different runs andrealizations of the same chain on the chimera graph. Errors are the standard errors of the mean. Note the dramatic change inthe behavior between the quenches that start with maximum initial coupling strength (upper right corner) and only 8% smallerinitial coupling strength (lower right corner).

motivation for this search comes from the scaling ≈ τ−1Q

in Fig. 3a. It is known [13] that decoherence with energyeigenstates as the pointer states [14] results in a τ−1

Q de-pendence for the LZ regime.

The results for L = 4 closed chains are in Fig. 3d.The scaling with τ−1

Q is still present. This tempts us toregard it as evidence of an anti-crossing in presence ofdecoherence [13]. The number of kinks, though, seems tobe larger than the theory can accommodate. Moreover,we found evidence against this “LZ with decoherence”interpretation. For one, quenches with a slightly smallervalue of the maximal Jmax behave differently. The num-ber of kinks can be nearly independent of τQ, see Fig. 3b.That qualitative change is rather abrupt. Furthermore,quenches with long chains seem to show little dependenceon chain length, while one expects the kink number toincrease with chain length.

We do not see how these features can be accommo-dated within any known general theories (e.g., LZ, KZM,LZ with decoherence). Furthermore, we find significantdifferences between Ising chains of the same length im-plemented using different “spins” (i.e., Josephson junc-tions) on the D-Wave chip, as well as differences betweenthe Los Alamos and Burnaby machines. In particular,the number of defects, as well as the scatter, is signifi-cantly smaller in the Los Alamos machine compared to

the Burnaby machine for similar Ising chains, quenchrates, etc. (see Methods).

IV. QUANTUM ISING CHAIN IN A HOSTILEENVIRONMENT

Many factors can be contributing to this unusual be-havior: Heating, randomness in couplings, eigenstatedecoherence, local decoherence, self-interactions, non-Markovian effects, noise, etc. Many of these issues willlikely be encountered in other settings. We note that,in our case, some of them can be ruled out, while oth-ers can not. The following discussion is inspired by ourthinking of what can happen to a quantum Ising chainimplemented on a D-Wave chip. Essentially, we discussthe behavior of Ising chains that are not completely iso-lated from their environment. We do not aim to be ex-haustive: We have selected models of decoherence thatcan be described relatively simply (which does not meanthat they can be readily solved!). We have also focusedon models that can be simply parametrized (thus, for ex-ample, we have avoided discussing “mixtures” of modelsthat – like models of noise – have several components).

This selection of what is to be discussed is in accordwith the goal we have – understanding of the role of

4

a b

10 30 50 70 90

L

0.0

0.4

0.8

1.2∆

L[G

Hz]

kBT1

kBT2

kBT3

Gap

Fit

0.5 1.5x

3.0

4.0

5.0

6.0

∆4

δ = 0

δ = 0.1

δ = 0.3

δ = 0.5

FIG. 4. Energy gaps. a. The gap ∆L at the criti-cal point sc for the Ising chain implemented on the D-Wavechip. Red dots where obtained from numerical calculationsusing the D-Wave protocol, Fig. 1(a). The solid line is thefit ∆L = ∆0 · L−x, where ∆0 = (10.84 ± 0.06) [GHz] andx = 0.973 ± 0.002. Dotted lines show the thermal energyfor the DW2X (Los Alamos) chip: T1 = (15.7 ± 1.0) mK,DW2X (Los Alamos) chip: T2 = (14 ± 1.15) mK and DW2X-SYS4 (Burnaby): T3 = (26 ± 5) mK. b. The gap ∆4 for theclosed random model (18) where both gi ∈ [x − δ, x + δ] andJi ∈ [1 − δ, 1 + δ] are drawn from a uniform distribution. Thedisorder has weak perturbative effect even at the anti-crossingcenter.

external factors in the dynamics of phase transitions asrepresented by a quantum Ising chain. This may comehandy not just in benchmarking of adiabatic quantumcomputers, but also in future condensed matter exper-iments where quantum many-body systems are driventhrough a symmetry breaking transitions in presence ofthe inevitable coupling with their environment. Thus,while the D-Wave chip is “on our mind”, we feel thatmany of the problems we shall encounter in the discussionof its physics will be also encountered in other settings.

Thermal excitation. Heating of the Ising chains isan obvious culprit that would add excitations – generatekinks. We do not believe that, in the D-Wave setting,it is dominant. The heating will be most effective nearthe critical point, as the temperatures of the two D-Wavechips we have worked with exceed the size of the gap onlyin its vicinity for the chains we studied: Figure 4 showsthe minimal energy gap (near the quantum critical point)for different lengths of the quantum Ising chain. Thus,kink generating transitions will be only effective for aperiod of time that is roughly proportional to τQ. If thiseffect was dominant it should result in the number ofkinks increasing with τQ. We observe the opposite trend

(e.g., τ−1Q in the Los Alamos chip). Furthermore, for

very short chains (e.g., squares) there is over an order ofmagnitude difference between the minimal gap and kBTof the chip, suppressing thermal excitations.

For above reasons, we conclude that “heating” is un-likely to be the dominant effect behind the generation ofkinks above the Landau-Zener theory predictions.

Coupling to the spins not in the chain. It is

a b

10-3 10-1 101

J/τQ

10-2

100

102

kinks

γ = 0

γ = 10−5

γ = 10−3

γ = 10−1

10-3 10-1 101

J/τQ

10-5

10-3

10-1

101

kinks

FIG. 5. Ghost spins - defects with decoherence. a.The number of kinks as a function of quench rate J/τQ fordifferent decoherence strengths γ and L = 100. Here, g(t) =J(1 − t/τQ) and γ = γ/J . b. The same as in a. but for aperiodic chain of length L = 4 (i.e., a “square”).

known that the spins on the D-Wave chip also coupleto the spins from which they are nominally decoupled.That is, setting the coupling Jkl = 0 between spins kand l does not guarantee that this coupling is indeednegligible. There are also reasons to believe that thiscoupling is predominantly “Ising” (∼ σz

kσzl ) rather than,

e.g., Heisenberg.We believe we have seen evidence of such spurious cou-

plings in the behavior of the Ising chains. For instance,the “compact chains” (that cover relatively small area ofthe chip) yield fewer kinks than “spread out chains” of thesame length. This would happen if the spurious couplingwith spins that should be decoupled from the chain re-sulted in the couplings between different fragments of thechain. This would have two related effects: The Hamil-tonian of Eq. (2) is no longer the whole story (as it willbe dressed with the couplings to the spins from whichit should be nominally decoupled). We will not modelthis effect (in part because it requires a detailed accountof how these spurious couplings occur and, in part, webelieve it may turn out to be too D-Wave-specific).

The second effect that we will model recognizes thatsuch “ghost spins” act as an environment that will deco-here fragments of the quantum Ising chain – “ghost spins”monitor the orientation of the spins inside the chain. Thisis of interest, and is likely to be ubiquitous in other real-izations of the quantum Ising systems, both in condensedmatter and quantum information processing devices.

We model this effect in Methods for both open chainsof varying length, see Fig. 5(a), and closed “squares”, seeFig. 5(b). There is a generic pattern that emerges: Whendecoherence due to “ghost spins” acts for sufficiently longtime, the number of kinks begins to increase with τQ untilit saturates at ≃ L/2. A similar effect was studied beforein Ref. [15] where it was described as anti-Kibble-Zurekbehaviour.

Randomness in the Hamiltonian. It is nowknown that the implementation of the Ising Hamiltonian,

5

Eq. (2), suffers from errors both in the value of the cou-plings between spins [i.e., J(t)] and the bias field g(t).These errors are difficult to characterize in detail espe-cially in the critical region where g(t) ≈ J(t). They tendto be several percent of the maximal values of g and J 1.The relative error, though, in g−J near the critical point,however, could be large.

Such randomness has a profound effect on the dynam-ics and kink generation that to some extent has been an-alyzed [16–18]. Random couplings and transverse fields,which we allow for in the Hamiltonian of Eq. (18), al-ters the universality class for a long enough chain. Wenote that we take the randomness in gi(t) and Ji(t) toinclude static random fluctuations around uniform g(t)and J(t), respectively. The number of kinks after a

quench is no longer a power law ∼ τ−1/2

Q predicted byKZM for a homogeneous chain but a logarithmic decay∼ (ln τQ)−2 [16–18].

This slow decay might possibly explain the absence ofthe exponential LZ decay for long enough chains: In thepresence of disorder, the adiabaticity estimate in Eq. (6)is no longer valid and much longer quench times are re-quired. However, the longer chains seem to conform to apower law rather than the logarithmic decay and, whatis more important, the power law persists even in shortchains like the L = 4 squares. As seen in Fig. 4(a), thesquare has a relatively large gap even at the anti-crossingso the disorder could only have a weak perturbative effecton the outcome of the quench, see Fig. 4(b). Thus, weconclude that disorder is not the main culprit for the ob-served discrepancies with respect to the pure Ising chain.

Decoherence in energy eigenstates. A model ofan anti-crossing with decoherence via Lindblad super-operators that are diagonal in the instantaneous energyeigenstates turns out to be exactly solvable [13]. More-over, for short chains (i.e., squares), decoherence thatfavors energy eigenstates can be relevant (as it tends toset in whenever the separation of energy levels is largecompared to the other relevant energy scales [14]).

In this regime, the probability of a transition to theexcited state is given by the equation

p =ε~

2∆2L

Q

(

∆L

)

, (8)

where γ and ε ∼ τ−1Q are the decoherence and transition

rates, respectively, and Q is a simple function with a max-imum value of ∼ 0.65 [13]. Our results with squares yieldvalues of Q that are close to Q ∼ 1 and that sometimes“dip” to within the region below 0.65 consistent with theequation above. We note that our estimates of the pa-rameters in Eq. (8) can be significantly affected by thecaveats listed above, so we cannot rule out significanceof this model for squares.

1 Private communication with D-Wave Inc.

In particular, the probability of kink formation forboth ferro and anti-ferro cases exhibits the same quench-rate dependence (τ−1

Q ) consistent with Eq. (8) only in theLos Alamos machine and when the scale of J is set to itsmaximal range, see Fig. 6(b). However, even a relativelymodest change of that scale from the maximum leadsto a fairly dramatic change in the behavior undermininghope in the utility of Eq. (8) for the problem at hand,see Fig. 3(b).

V. DISCUSSION

Complex behavior of quantum annealers demandsglobal tests of adiabaticity and quantumness, as evenwhen components of the device work, their integrationraises questions of decoherence, control, and what “slowenough” is. We propose a global test based on a quenchin the quantum Ising chain. It can assess reliability of thewhole device. Such general tests will prove valuable inestablishing adiabaticity and benchmarking/comparingdifferent implementations of adiabatic quantum comput-ers expected in the near future.

In spite of the outcome of the TAC, D-Wave may,in some cases, find the right or at least approximate,solutions to problems. Obviously, a more precise imple-mentation would result in a more successful adiabaticquantum computation/quantum anneal. Indeed, thenoticeable decrease in the number of defects between thetests of Burnaby and Los Alamos machines is likely dueto the improvements in hardware. One can hope that thenext generation of quantum annealers will be even better.

Methods

Numerical simulations. To obtain the results inFig. 2b, we first brought the Hamiltonian (2) into itsfermionic representation [8],

H = 2

L∑

n=1

gnc†ncn −

L∑

n=1

gn

L−1∑

n=1

Jn

(

c†ncn+1 + c†

nc†n+1 + h.c.

)

,

(9)

using the Jordan-Wigner transformation [19]

σzn =

(

c†n + cn

)

m<n

(

1 − 2c†mcm

)

, (10)

σxn = 1 − 2c†

ncn, (11)

where cn (c†n) is the fermionic annihilation (creation) op-

erator for site n. For the quadratic correlation functionsxpq := 〈c†

pcq〉 and ypq := 〈c†pc†

q〉, this gives the closed

6

system of equations

ixp,q = −Jp xp+1,q − Jp−1 xp−1,q + Jq xp,q+1

+ Jq−1 xp,q−1 + Jq yp,q+1 − Jq−1 yp,q−1

+ Jp y∗p+1,q − Jp−1 y∗

p−1,q

+ 2 (gp − gq) xp,q, q ≥ p;

(12)

and

iyp,q = −Jp yp+1,q − Jp−1 yp−1,q − Jq yp,q+1

− Jq−1 yp,q−1 − Jq xp,q+1 + Jq−1 xp,q−1

+ Jp x∗p+1,q − Jp−1x∗

p−1,q − Jp δp+1,q

+ 2(gp + gq)yp,q, q > p,

(13)

with ypp = 0. The above equations are solved with theinitial condition corresponding to the system’s groundstate when J > 0 and with the boundary conditions c0 =cL+1 = 0. To carry out numerical computations, we usedan adaptive Adams method from LSODA. Finally, thenumber of kinks was obtained from

kinks =L − 1

2−

L−1∑

p=1

ℜ (xp,p+1 + yp,p+1) . (14)

Here ℜ means real part. Both the ground state and thegap depicted in Fig. 2(a) where calculated using tech-niques described in Ref. [20].

Thus, to compute the number of kinks we used thefollowing formula

kinks =N |Jmax| + E

2|Jmax|, (15)

N is the numbers of couplings in the chain. The finalenergy E can be read in directly from the D-Wave solver.

Burnaby versus Los Alamos chip. One wouldexpect different chips of the same generation of anneal-ers to generate roughly the same number of kinks forferro and anti-ferro cases. However, the DW2X based inLos Alamos seems to perform better (i.e., generates lesskinks) when J < 0, see Fig. 6(a). The Burnaby machine,though, has the same behavior for ferro and anti-ferrocases, see Fig. 6(b).

Thus, chips that belong to different architectures maybehave differently. For instance, Fig. 7 compares theDW2X based in Los Alamos and a previous generationDW2X-SYS4 in Burnaby. Not only do the number ofkinks differ between these two systems but it also exhibits

different quench-time dependence (τ−1Q versus τ

−1/2

Q ).Decoherence by “ghost spins”. Numerical re-

sults presented in Fig. 5 are obtained using the followingLinbdlad master equation [19]:

ρ(t) =1

i~[H(t), ρ(t)] + γD[ρ(t)], (16)

where the superoperator is

D[ρ(t)] = −1

2

L∑

n=1

[σzn, [σz

n, ρ(t)]] (17)

a

10-3 10-1 101

τ−1

Q [MHz]

10-1

100

101

kinks

Jmax = 1.0

A = 3.18 ± 0.033

x = 0.24 ± 0.01

10-3 10-1 101

τ−1

Q [MHz]

10-5

10-3

10-1

101

kinks

Jmax = −1.0

A = 0.159 ± 0.002

x = 1.31 ± 0.02

b

10-3 10-1 101

τ−1

Q [MHz]

10-1

100

101

kinks

Jmax = 1.0

A = 3.519 ± 0.018x = 0.236 ± 0.003

10-3 10-1 101

τ−1

Q [MHz]

10-1

100

101

kinks

Jmax = −1.0

A = 3.798 ± 0.015x = 0.249 ± 0.002

FIG. 6. Comparison between the same D-Wave archi-tectures (L = 300). a. DW2X system based in Los Alamos.The number of kinks is different for ferro and anti-ferro cases.Smaller number of defects when J < 0 suggest that the LosAlamos chips performs better in this regime. b. DW2X sys-tem based in Burnaby. As one would expect, the number ofkinks is roughly the same for both ferro and anti-ferro cases.

and H(t) takes the form

H(t) = −

L∑

n=1

gn(t)σxn −

L−1∑

i=n

Jn(t)σznσz

n+1, (18)

where we allow for time and spatial dependence in bothJn and gn.

Expectation values, 〈O〉 = Tr(Oρ), of an operator Oevolve according to

d

dt〈O〉 =

1

i~〈[O, H ]〉 −

γ

2

L∑

n=1

〈[σzn, [σz

n, O]]〉. (19)

This equation is solved using the Jordan-Wigner trans-formation [20]

σzn =

(

c†n + cn

)

m<n

(

1 − 2c†mcm

)

, (20)

σxn = 1 − 2c†

ncn, (21)

where cn (c†n) is a fermionic annihilation (creation) oper-

ator. For an open chain, the above transformation brings

7

a

10-3 10-1 101

τ−1

Q [MHz]

10-5

10-3

10-1

101kinks

L = 50

A = 0.039 ± 0.001x = 1.04 ± 0.04

10-3 10-1 101

τ−1

Q [MHz]

10-5

10-3

10-1

101

kinks

L = 100

A = 0.067 ± 0.002x = 1.2 ± 0.06

b

10-3 10-1 101

τ−1

Q [MHz]

10-3

10-1

101

kinks

L = 50

A = 0.12x = 0.43

10-3 10-1 101

τ−1

Q [MHz]

10-3

10-1

101kinks

L = 100

A = 0.48x = 0.5

FIG. 7. Comparison between different D-Wave ar-chitectures (Jmax = −1). a. DW2X system based in LosAlamos. b. A previous generation DW2X-SYS4 in Burn-aby. The DW2X chip is better, i.e. it produces less kinks.Moreover, these two architectures exhibit different quench-

time dependence (τ −1

Q versus τ−1/2

Q ).

the Hamiltonian, Eq. (18), to the following form [8]

H = 2

L∑

n=1

gnc†ncn − gnL

L−1∑

n=1

Jn

(

c†ncn+1 + c†

nc†n+1 + h.c.

)

.

(22)

The string operators in Eq. (21) cancel out in the Lind-blad contribution to the right hand side of Eq. (19), hencethe quadratic fermionic correlation functions xpq :=〈c†

pcq〉 and ypq := 〈c†pc†

q〉 satisfy a closed set of equations(q ≥ p),

ixp,q = −Jp xp+1,q − Jp−1 xp−1,q + Jq xp,q+1

+ Jq−1 xp,q−1 + Jq yp,q+1 − Jq−1 yp,q−1

+ Jp y∗p+1,q − Jp−1 y∗

p−1,q

+ 2 (hp − hq) xp,q + γD[xpq ],

(23)

where the Lindblad superoperator D[xpq ] reads

D[xpq ] =

{

1 − 2xpp if p = q,

2ℜ(ypq) − 2|q − p|xpq if q > p,(24)

together with ypp = 0 and (q > p)

iyp,q = −Jp yp+1,q − Jp−1 yp−1,q − Jq yp,q+1

− Jq−1 yp,q−1 − Jq xp,q+1 + Jq−1 xp,q−1

+ Jp x∗p+1,q − Jp−1x∗

p−1,q − Jp δp+1,q

+ 2(hp + hq)yp,q + γD[ypq],

(25)

with D[ypq] = 2ℜ(xpq) − 2|q − p|ypq.These equations are to be solved with the initial con-

dition corresponding to the system’s ground state whenJ > 0 and with the boundary conditions c0 = cL+1 =0 [20]. The number of kinks is then given by

kinks =L − 1

2−

L−1∑

p=1

ℜ (xp,p+1 + yp,p+1) . (26)

Bibliography

[1] Kadowaki, T. & Nishimori, H. Quantum annealing in thetransverse ising model. Phys. Rev. E 58, 5355 (1998).

[2] Farhi, E., Goldstone, J., Gutmann, S. & Sipser, M.Quantum computation by adiabatic evolution. Preprintat arXiv:0001106v1 (2000).

[3] Aharonov, D., van Dam, W., Kempe, J., Landau, Z.,Lloyd, S. & Regev, O. Adiabatic quantum computa-tion is equivalent to standard quantum computation.SIAM J. Comput. 37, 166 (2007).

[4] Ladd, T. D., Jelezko, F., Laflamme, R., Nakamura,Y., Monroe, C. & O’Brien, J. L. Quantum computers.Nature 464, 45 (2010).

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

[6] Zurek, W. H., Dorner, U. & Zoller, P.Dynamics of a quantum phase transition.Phys. Rev. Lett. 95, 105701 (2005).

[7] Dziarmaga, J. Dynamics of a quantum phase tran-sition: Exact solution of the quantum Ising model.Phys. Rev. Lett. 95, 245701 (2005).

[8] Dziarmaga, J. Dynamics of a quantum phasetransition and relaxation to a steady state.Adv. Phys. 59, 1063 (2010).

[9] Kibble, T. W. B. Topology of cosmic domains andstrings. J. Phys. A: Math. Gen 9, 1387 (1976).

[10] Zurek, W. H. Cosmological experiments in superfluid he-lium?. Nature 317, 505 (1985).

[11] Quan, H. T. & Zurek, W. H. Testing quantum adiabatic-ity with quench echo. New J. Phys 12, 093025 (2010).

[12] Boixo, S., Datta, A., Davis, M. J., Flammia, S. T., Shaji,A. & Caves, C. M. Quantum metrology: Dynamics versusentanglement. Phys. Rev. Lett. 101, 040403 (2008).

[13] Avron, J. E., Fraas, M., Graf, G. M. & Grech, P.Landau-Zener tunneling for dephasing lindblad evolu-tions. Commun. Math. Phys. 305, 633 (2011).

[14] Paz, J. P. & Zurek, W. H. Quantum limit of deco-herence: Environment induced superselection of energyeigenstates. Phys. Rev. Lett. 82, 5181 (1999).

8

[15] Dutta, A., Rahmani, A. & del Campo, A. Anti-Kibble-Zurek behavior in crossing the quantum critical point of athermally isolated system driven by a noisy control field.Phys. Rev. Lett. 117, 080402 (2016).

[16] Dziarmaga, J. Dynamics of a quantum phase tran-sition in the random ising model: Logarithmic de-pendence of the defect density on the transition rate.Phys. Rev. B 74, 064416 (2006).

[17] Caneva, T., Fazio, R. & Santoro, G. E. Adiabatic quan-tum dynamics of a random ising chain across its quantumcritical point. Phys. Rev. B 76, 144427 (2007).

[18] Cincio, L., Dziarmaga, J., Meisner, J. & Rams, M. M.Dynamics of a quantum phase transition with decoher-ence: Quantum ising chain in a static spin environment.Phys. Rev. B 79, 094421 (2009).

[19] Dziarmaga, J., Zurek, W. H. & Zwolak, M. Non-local quantum superpositions of topological defects.Nat. Phys. 8, 49 (2012).

[20] Lieb, E., Schultz, T. & Mattis, D. Two soluble models ofan antiferromagnetic chain. Ann. Phys. 16, 407 (1961).

Acknowledgments

We appreciate fruitful discussions with our Los Alamoscolleagues Marcus Daniels, Seth Lloyd, Scott Pakin, andRolando Somma, as well as Edward Dahl and TrevorLanting of D-Wave Systems. Work of J.D. and B.G. wassupported by Narodowe Centrum Nauki (National Sci-ence Center) under Project No. 2016/23/B/ST3/00830and 2016/20/S/ST2/00152, respectively. Work ofW.H.Z. was supported by the US Department of Energyunder the Los Alamos LDRD program. This researchwas supported in part by PL-Grid Infrastructure.

Author contributions

B.G., J.D., W.H.Z., M.Z. contributed equally to thiswork.

Additional information

Competing interests: The authors declare no compet-ing interests.