You are on page 1of 16

How Quantum is the D-Wave Machine?

Seung Woo Shin* , Graeme Smith , John A. Smolin , and Umesh Vazirani*

arXiv:1401.7087v1 [quant-ph] 28 Jan 2014

Computer Science division, UC Berkeley, USA.


IBM T.J. Watson Research Center, Yorktown Heights, NY 10598, USA.

Abstract
Recently there has been intense interest in claims about the performance of the D-Wave machine.
Scientifically the most interesting aspect was the claim in [7], based on extensive experiments, that
the D-Wave machine exhibits large-scale quantum behavior. Their conclusion was based on the
strong correlation of the input-output behavior of the D-Wave machine with a quantum model
called simulated quantum annealing, in contrast to its poor correlation with two classical models:
simulated annealing and classical spin dynamics. In this paper, we outline a simple new classical
model, and show that on the same data it yields correlations with the D-Wave input-output behavior
that are at least as good as those of simulated quantum annealing. Based on these results, we
conclude that classical models for the D-Wave machine are not ruled out. Further analysis of the
new model provides additional algorithmic insights into the nature of the problems being solved by
the D-Wave machine.

Introduction

In a future world of quantum devices, it will become increasingly important to test that
these devices behave according to specification. While this is clearly a central issue in the
context of quantum cryptography [4, 27] and certified random number generators [17, 26],
it is also quite fundamental in the context of testing whether a claimed quantum computer
is really quantum [2, 18, 14, 5]. Recently, this last issue has featured prominently in the
context of the D-Wave machine [12, 6, 7, 1, 15], amidst questions about whether it should
really be thought of as a quantum computer and whether it provides speedups over classical
computers. One of the grounds for skepticism is that the decoherence time of D-Waves
qubits is reported to be on the order of nanoseconds, which is comparable to the time for
a single operation and much shorter than the time required to carry out a computation,
which is on the order of microseconds.
By now, claims about speedups over classical computers [16] have been largely refuted [7, 22, 19]. But one of the papers [7] that disproved the claims of quantum speedup
1

also claimed based on extensive experiments that the D-Wave machine exhibits large-scale
quantum behavior. While falling far short of the existence of a working quantum computer,
this conclusion is nonetheless very exciting, since it suggests that it might be possible to
create large-scale entanglement without investing in stable, slowly decohering qubits. For
this reason it is particularly important to closely examine the evidence presented in the
paper.
The paper recorded the input-output behavior of the machine D-Wave One on a thousand randomly chosen inputs and compared it to the input-output behavior obtained by
simulating three models. The first model was simulated quantum annealing1 , and the other
two were classical models simulated annealing, and classical spin dynamics suggested in
[23]. The paper reported that the input-output behavior of D-Wave One correlated highly
with simulated quantum annealing, but despite considerable effort in choosing parameters
did not correlate well with the two classical models. Based on this, the paper concluded
that the D-Wave machine was indeed following quantum dynamics, and exhibited largescale quantum behavior.
At a high level, the nature of this argument is very interesting, and reminiscent of the
algorithmic approaches in [2, 18, 9, 14, 5]. In those papers, the quantum system to be
tested is modeled as a black box, and elaborate protocols are designed whose execution
guarantees that the untrusted quantum systems behaves according to specification. Given
the limited access to the D-Wave machine, which is proprietary and with access controlled
by the company, the approach of [7] may be seen as algorithmic in nature. The evidence of
quantumness is based on broad algorithmic differences between the three models rather
than detailed physical modeling of the system. In particular, the core of the argument is
based on the finding that the D-Wave machine and simulated quantum annealing generally
find the same set of instances to be hard and the same set of instances to be easy.
The hardness of an instance is measured by the number of times that each candidate
algorithm succeeds in finding the optimal solution out of one thousand trials. Of course,
how convincing this evidence is depends upon how comprehensive the list of candidate
classical models is.
In this paper, we consider a third classical model for the D-Wave machine. The model
is quite natural, and it replaces each spin in the D-Wave machine with a magnet pointing
in some direction in the XZ plane. Each magnet is subject to both an external magnetic
field as well as dipole-dipole couplings due to interactions with nearest neighbors on the
so-called Chimera graph. The external magnetic field is attenuated at the same rate as
in [8], while the couplings between magnets also follow the schedule from [8]2 . We consider
1

Simulated quantum annealing refers to a quantum Monte Carlo simulation of quantum annealing. In
this sense, it is a classical algorithm. However, one can argue that the quantum Monte Carlo simulation is
quite involved and the D-Wave machine is surely not carrying out such a complex computation. Therefore,
it may be reasonable to regard this as a quantum process.
2
We note that the schedule we consider here is different from the one reported in [7]. We were informed [24] that the schedule in [7] was based on an error by D-Wave, and the correct schedule is reported

the input-output behavior of this model on the same set of one thousand inputs. Our
simulations show that it achieves as good or better correlation with the D-Wave machines
input-output behavior than simulated quantum annealing does. Based on these results, it
is safe to say that classical models for the D-Wave machine are not ruled out, and it is
premature to conclude that the D-Wave machine exhibits large-scale quantum behavior.
Matthias Troyer suggested [24] a direct comparison between our model and simulated
quantum annealing, which reveals an extremely high correlation of R 0.99 (Fig 9).
One way to view this result is that our model is the classical analogue of a mean-field
approximation to simulated quantum annealing, and that for the set of problems solved by
D-Wave One, this approximation is very accurate. This helps explain why the correlation
of the D-Wave data with simulated quantum annealing does not preclude a classical model
for D-Wave.
Further analysis of the algorithmic behavior of this new model also provides new insights
into the nature of the computational problem being solved, namely the problem of finding
the ground state of a classical Ising spin glass on the so-called Chimera graphs. The
Chimera graph consists of super-nodes or clusters of 8 qubits (connected in a complete
bipartite graph K4,4 ) arranged in a two-dimensional grid. The D-Wave machine consists
of superconducting qubits arranged on the vertices of such a graph, with a z z coupling
between neighboring qubits. An analysis of the new model suggests that the existence of
these super-nodes greatly reduces the effective size of the computational problem being
solved, so that heuristically the size of the problem should be thought of as the number
of super-nodes m = 16, rather than the number of qubits n = 108. To the extent that
the model accurately represents the dynamics of the D-Wave machine, it therefore has
great implications for the performance of the D-Wave machine as it is scaled up to larger
numbers of qubits.

Background

We start by defining the computational problem that D-Wave machines are designed to
solve. The native problem of D-Wave machines is the problem of finding the ground
state of a classical Ising spin glass, which can be described as follows. The input to the
problem is a weighted graph with n vertices, where each vertex represents a spin and
each edge represents interaction between the two incident vertices. The edge weights Jij
can be either 1 or 1, corresponding to ferromagnetic and antiferromagnetic interaction
respectively. The problemP
is simply to assign a spin value zi {1, 1} to each vertex i
so that the energy H = i<j Jij zi zj is minimized, i.e. finding the ground state of the
given interaction graph.
An additional constraint imposed by D-Wave machines is that the input graph must be a
in [8]. Our results are reported for this new schedule. All the claims we report here apply equally well to
the previous schedule in [7].

Supplementary material for Quantum annealing with more

12

16

20

24

28

13

17

21

25

29

10

14

18

22

26

30

II.

THE CHIMER

The qubits and cou


thought of as the ve
bipartite graph, called
figure 1. This graph
eight qubits each. W
couplers realise a com
each of the four qub
the four on the right
left is furthermore co
in the unit cell above
on the right is horizo
ing qubits in the uni
appropriate modificat
the 128 qubits in the
in the experiments ar
between them are ma
For our scaling an
cedure for scaling of
FIG.
1:
Qubits
and
couplers
in
the
D-Wave
device.
Figure 1: The Chimera graph structure implemented by the D-Wave One. Grey
dots
sidering
the chimera
D-Wave
One Rainer
consists
4 Figure
4 unitiscells
represent The
the twenty
non-working
qubitschip
in the
D-WaveofOne.
fromof
[7]. with an eight-site un
eight qubits, connected by programmable inductive couplers
tions. The sizes we ty
as shown by lines.
are L = 1, .
subgraph of the so-called Chimera graph, which is depicted in Figure 1. The Chimeraulations
graph
consists of copies of complete bipartite graphs K4,4 s which are laid out in 2D lattice8,as32,
in 72, 128, 200, 288
Figure 1. We will refer to each copy of K4,4 as supernode. Note that the left qubits in
each
lated annealers and
supernode are coupled vertically in the 2D lattice and the right qubits horizontally.above
More we used a perf
specifically, each left qubit is coupled with the corresponding left qubit in the supernodes
128 where we compar
immediately above and below it, and each right qubit to the corresponding right qubits in
qubits
supernodes immediately to the right and left. On the coarser level, this structure can
be within selectio
the
shown in fi
viewed as a 2D lattice of size 4 4. While the problem of finding the ground state of graph
a
I. OVERVIEW
In references [29,
classical Ising spin glass is NP-hard, it is known to have a polynomial time approximation
scheme when the input graph is a Chimera graph [20].
sation problem on a
The D-Wave machine places a superconducting flux qubit at each node of the Chimera
can be mapped to an
graph with Npvertice
Here we provide additional details in support of the
tree width of N men
main text. Section II shows details of the chimera graph
this mapping. See S
4
used in our study and the choice of graphs for our simulaabout the tree width
tions. Section III expands upon the algorithms employed
in our study. Section IV presents additional success probability histograms for dierent numbers of qubits and for
III. CLAS
instances with magnetic fields, explains the origin of easy
and hard instances, and explains how the final state can
A. S
be improved via a simple error reduction scheme. Section
3

11

15

19

23

27

31

32

36

40

44

48

52

56

60

33

37

41

45

49

53

57

61

34

38

42

46

50

54

58

62

35

39

43

47

51

55

59

63

64

68

72

76

80

84

88

92

65

69

73

77

81

85

89

93

66

70

74

78

82

86

90

94

67

71

75

79

83

87

91

95

96

100

104

108

112

116

120

124

97

101

105

109

113

117

121

125

98

102

106

110

114

118

122

126

99

103

107

111

115

119

123

127

Figure 2: The annealing schedule of the D-Wave One. Figure is from [8].
graph. The time-dependent Hamiltonian of D-Wave machines is given by
X
X
H(t) = A(t)
ix B(t)
Jij iz jz
i<j

1in

where A(t) and B(t) represent the annealing schedule of the machine, which is shown in
Figure 2, and the process is carried out at a finite temperature. As mentioned in a footnote
in the introduction, this schedule is different from the one reported in [7], which was based
on an error by D-Wave [24], although this did not change their conclusions. Our paper is
based entirely on the new corrected schedule [8]. We note that our conclusions also hold
equally well for both schedules.
If the entire process were carried out sufficiently slowly and at zero temperature, then
it would be an implementation of adiabatic quantum optimization [13]. Such
Pa process is
guaranteed to end up in the ground state of the target Hamiltonian H = i<j Jij Zi Zj .
For this to happen, the total time for the process has to scale as (1/2 ), where is the
minimum spectral gap of the time-dependent Hamiltonian where the minimum is taken
over the annealing schedule. It is known that in the worst case scales exponentially in
the problem size [13, 25, 3]. Quantum annealing can be thought of as a noisy version of
adiabatic quantum computing which is carried out at a finite temperature.
Finally, we summarize the previous results on the question of whether D-Wave machines
exhibit large-scale quantum behavior. We note that there were a few earlier papers that
studied related questions [12, 6], but the first paper to study this question at a sufficiently
large scale was [7], in which the behavior of the 108-qubit D-Wave One was compared to
simulations of quantum and classical models. The method of comparison was as follows.
First, one thousand instances of the Ising spin glass problem were generated randomly.
5

Number of instances

500
400

300

200

200

100

100

0
0.0
400

0.2

0.4

0.6

0.8

0
0.0

1.0

500

C) SA

400

300

300

200

200

100

100

0
0.0

0.2 0.4 0.6 0.8


Success probability

B) SQA

400

300

500
Number of instances

500

A) DW

0
0.0

1.0

0.2

0.4

0.6

0.8

1.0

D) SD

0.2 0.4 0.6 0.8


Success probability

1.0

Figure 3: Histogram of success probabilities from [8]. It is observed that the histogram is
bimodal for D-Wave, simulated quantum annealing, and classical spin dynamics, whereas
it is unimodal for SA. This means that the former three algorithms divide the problem
instances into two groups, namely easy and hard. They succeed almost always on the
easy instances and fail almost always on the hard instances.
The problem size was 108 for all of these instances, in order that all of the 108 working
qubits of the D-Wave One are utilized. Then, each of the following three algorithms were
run one thousand times on each of these one thousand problem instances: the D-Wave
One, simulated quantum annealing, and simulated (classical) annealing. Based on this,
the success probability of each algorithm was recorded for each instance, where success is
defined to be the finding of the exact optimum, i.e. the ground state, for the instance. The
histogram of success probabilities of each algorithm, shown in Figure 3, served as the main
evidence for the papers conclusion, which was that the D-Wave One was consistent with
simulated quantum annealing rather than the classical models and therefore D-Wave One
is indeed performing quantum annealing. A parallel conclusion of the paper was that their
simulated annealing code was in fact up to 6 times faster than the D-Wave One, indicating
no speedup over classical computers.
This result was immediately questioned by Smolin and Smith in [23], who pointed out
that the unimodal histogram of simulated annealing was simply a consequence of statistical
independence between different runs and that it was not a signature of all classical algorithms. They went on to propose a classical model for the D-Wave machine based on O(2)
rotor model, which also exhibited bimodal signature in the success probability histogram.
6

20"

instances"

40"

0"
FIG. 3: Correlations. Panels A-C show scatter plots of
correlations of the success probabilities p(s) obtained from
dierent methods. The red lines indicate perfect correlation.
Panel A is for the D-Wave device between two sets of eight
dierent gauges. This data shows the baseline imperfections
in the correlations due to calibration errors in the D-Wave device. Panel B is for the simulated quantum annealer (SQA)
and the D-Wave device, with the latter averaged over 16 random gauges. This correlation is nearly as good as in panel A,
indicating good correlations between the two methods.. Panel
C is for the classical spin dynamics and the D-Wave device,
and shows poor correlation. Panel D shows the correlation
between success probability and the mean Hamming distance
of excited states found at the end of the annealing for N = 108
spin instances with local random fields. Easy (hard) instances
tend to have a small (large) Hamming distance. The colour
scale indicates how many of the instances are found in a pixel
of the plots.

Figure 4: Scatterplot of success probabilities from [8]. The correlation between D-Wave
and SQA is noticeably better than that between D-Wave and the classical models.
FIG. 2: Energy-success distributions. Shown is the joint
probability distribution p(s, ) (colour scale) of success probability s and the final state energy
measured relative to
the ground state. We find very similar results for the D-Wave
device (panel A) and the simulated quantum annealer (panel
B). The distribution for simulated classical annealing (panel
C) matches poorly and for spin dynamics (panel D) matches
only moderately. For the D-Wave device and SQA the hardest instances result predominantly in low-lying excited states,
while easy instances result in ground states. For SA most
instances concentrate around intermediate success probabilities and the ground state as well as low-lying excited states.
For classical spin dynamics there is a significantly higher incidence of relatively high excited states than for DW, as well as
far fewer excited states for easy instances. The histograms of
figure 1, representing p(s), are recovered when summing these
distributions over
. SA distributions for dierent numbers
of sweeps are shown in the supplementary material.

The authors of [7] soon responded in [28] that their conclusions were in fact based on more
detailed correlation analysis, showing that the correlation between the D-Wave success
probabilities and SQA success probabilities is much higher than the correlation between
the D-Wave success probabilities and success probabilities of the classical models (Figure
In panel C) we show the correlation between the classi4). Along with these results, they also released
the dataset from their experiments as a
cal spin dynamics model and the device. Some instances
are easily solved by the classical mean-field dynamics,
service
the community,
to perform
the
studyand
reported
this paper.
classicalto
annealer
(panel C) and spinwhich
dynamicsallowed
(panel D). us simulated
quantum
annealing,
the device.in
However,

The third test, shown in figure 3, is perhaps the most


enlightening, as it plots the correlation of the success
probabilities between the DW data and the other models.
As a reference for the best correlations we may expect,
we show in panel A) the correlations between two dierent sets of eight gauges (dierent embeddings of the same
problem on the device, see Methods and supplementary
material): no better correlations than the device with itself can be expected due to calibration errors. Panel B)
shows a scatter plot of the hardness of instances for the
simulated quantum annealer and the D-Wave device after
gauge averaging. The high density in the lower left corner (hard for both methods) and the upper right corner
(easy for both methods) confirms the similarities between
the D-Wave device and a simulated quantum annealer.
i for instances of interThe two are also well correlated
mediate hardness. The similarity to panel A) suggests
almost perfect correlation with SQA, to within calibration uncertainties.

Results

as can be expected from inspection of their respective


distributions in figure 1, there is no apparent correlation
between the hard instances for the spin dynamics model
and the success probability on the device, nor does there
appear to be a correlation for instances of intermediate
hardness, in contrast to the correlations seen in panel A).
Similarly, there are poor correlations [22] with a classical
spin dynamics model of reference [23].
The correlations between the simulated classical annealer and the D-Wave device, shown in the supplementary
i material, are significantly worse than between SQA
and the device.
We next provide evidence for the bimodality being due
to quantum eects. Our first evidence comes from the
simulated quantum annealer. When lowering the temperature thermal updates are suppressed, quantum tunneling dominates thermal barrier crossing, and we observe
a stronger bimodality; indeed a similar bimodal distribution arises also in an ensemble of (zero-temperature)

In this paper, we consider the following classical model: each spin i is modeled by a
classical magnet pointing in some direction in the XZ plane (since there is no y term
in our time-dependent Hamiltonian, we can assume that there is no Y component). We
assume that the angle is measured in relation to x
. We further assume that there is an
external magnetic field of intensity A(t) pointing in the x
-direction, and that neighboring
magnets are coupled via either ferromagnetic or anti-ferromagnetic coupling, according to
the specification of the input graph. Our Hamiltonian can then be rewritten as follows:
X
X
H(t) = A(t)
cos i B(t)
Jij sin i sin j .
i<j

1in

Thus, in the absence of noise, i.e. at zero temperature, each spin i willPsimply align with
the net effective field at that location which is given by A(t)
x + B(t)
z j Jij sin j .
To model the noise in the system, we perform a Metropolis-type update inside this
model. That is, at each time step, we do the following:
1. For each spin i, pick i0 at random and compute the change in energy that would result
if the spin were to change its state to i0 :
X
Ei = A(t)(cos i0 cos i ) B(t)
Jij sin j (sin i0 sin i )
j

300"

Number'of'instances'

250"
200"
150"
100"
50"
0"
0.2"

0.4"
0.6"
0.8"
Success'probability'
Our"model"
DWave"

1"
Our models success probability

Figure 5: Histogram and scatterplot of our classical model. The correlation coefficient R
between the D-Wave One and our model is about 0.89.
2. Update spin is state to i0 with probability eEi /T where T is the temperature of
the system which is kept constant throught the simulation.
Note that this model is related to the O(2) rotor model of [23], but differs in that we
perform Metropolis-type updates instead of integrating the equation of motion. Indeed the
update procedure is the direct analogue of the Metropolis algorithm in O(2) model.
Our experimental results (Figure 5) show that this simple model not only exhibits a
bimodal signature similar to that of the D-Wave One or simulated quantum annealing, but
it also achieves a very high correlation with the success probabilities of the D-Wave One
which were published in [28]. Our success probabilities were obtained by simulating the
above model one thousand times on each instance, where each run consisted of 500,000
steps and the system temperature was fixed at T = 0.22. We note that the correlation of
0.89 achieved by our model is higher than the correlation achieved by simulated quantum
annealing in [8].

Discussion

It is instructive to analyze the behavior of our classical model and to compare and contrast
it with simulated annealing. First, if the transverse field were set to 0, i.e. if A(t) = 0 for all
t, then the model simplifies to an O(2) analogue of simulated annealing. To see this, note
that for the Metropolis acceptance probability function eEi /T , keeping the temperature
T constant and increasing the coupling strength B(t) has the same effect as keeping B(t)
constant and decreasing the temperature T over time. Simple experiments confirm that
8

ir%conclusions%were%
sis.%In%parHcular,%the%
er%than%the%correlaHon%
3
ing%rotor%models.%
odel%for%D;Wave%that%

g Woo Shin

Role of transverse field

What happ

!If%we%remove%the%transverse%eld%A(t)%from%our%model,%we%obtain
O(2)%simulated%annealing,%which%shows%unimodal%signature.%
!A%case%study:%the
!How%does%the%transverse%eld%aect%the%evoluHon?%

Umesh Vazirani

t=0.1725%up%to%two
Computer Science
division
Provides%gradaHon%in%the%magnitude%of%z;components.%
!Comparison%of%th
3
orientaHon%betwee
UC Berkeley

transverse field

weak bias in z direction

with transverse field

weak bias in z direction

1%

0.5%

z;component%

without transverse field

0%

during%the%branchi
!As%t%increases,%sm
!Other%instances%s

;0.5%

Rotor model
ll figures are from [1]

move%the%transverse%eld%A(t)%from%our%model,%we%obtain%
;1%
ated%annealing,%which%shows%unimodal%signature.%
FIG. 3: Correlations. Panels A-C show scatter plots of
correlations of the success probabilities p(s) obtained from
es%the%transverse%eld%aect%the%evoluHon?%
dierent methods. The red lines indicate perfect correlation.
strong bias in z-direction

strong bias in z-direction

Spins%(sorted%in%decreasing%order)%
with%X%eld%
without%X%eld%

and%Smith%[2],%but%
! Reminiscent%of%the%relaHonship%between%cuts%and%eigenvectors%
Panel A is for the D-Wave device between
two
sets6:ofRole
eightof transverse field
des%gradaHon%in%the%magnitude%of%z;components.%
Figure
her%than%integraHng%
dierent gauges. This data shows the
baseline
imperfections
of%graphs%in%spectral%graph%theory.%

in the correlations
due
to calibration
ansverse
field
with
transverse
field errors in1%the D-Wave de-

vice. Panel B is for the simulated quantum annealer (SQA)


and the D-Wave device, with the latter averaged over 16 rancatter
plots ofThis correlation is nearly as good
0.5% as in panel A,
dom gauges.
ector%poinHng%at%
obtained
indicatingfrom
good correlations between!
the
two
methods.. Panel
At%zero%temperature,%i.e.%in%the%absence%of%noise,%the%simulaHon%
given%by%
ect
C correlation.
is for the classical spin dynamicscan%be%viewed%as%a%xed%point%iteraHon%procedure.%%
and the D-Wave device,
0%
wo
sets
of eight
s in
z direction
weak bias in z direction
and
shows
poor correlation.
Panel D shows the correlation
imperfections
Je ij
sin isuccess
sin jprobability and the!
By%a%theorem%in%topology,%all%xed%points%(equilibria)%are%
between
mean
Hamming distance
the
D-Wave
de;0.5%
of excited states found at the end of the
annealing
for N = 108
reachable%from%the%starHng%point.%
annealer
(SQA)
spin
instances
with
local
random
fields.
Easy
(hard)
instances
ns%with%the%net%
!Thus,%as%Hme%t%changes,%the%model%traces%the%xed%points%
edtend
overto16have
ran- a small (large) Hamming
distance.
The
colour
nite%temperature,%
;1%
dter
as
in
panel
A,
(equilibria)%of%the%Hme;dependent%update%funcHon.%
scale
indicates
in a pixel
plots
of how many of the instances are foundSpins%(sorted%in%decreasing%order)%
methods..
Panel strong bias in z-direction
ofz-direction
the plots.
with%X%eld%
without%X%eld%
s in
tained
from
D-Wave device,
5%
correlation.
Annealing'schedule'of'D/Wave'(approximate)'
A(t)%
the
correlation
4.5%
Figure
7:
Snapshot
of
a
typical
run:
z-components
of the spins, sorted in decreasing order.
sets
of
eight
mming distance
B(t)%
4%
owing%update:%%
In
panel
C)
we
show
the
correlation
between
the
classi3.5%
mperfections
ng for N = 108
y%at%random%and%
cal spin
dynamics model and the device.
Some instances
3%
hard)
instances
D-Wave
deour
model
indeed exhibits
a unimodal
signature if the transverse field is removed. This
2.5%mean-field
E
%%%%%%%%%if%the%spin%was%
are
easily
solved
by
the
classical
dynamics,
i
ce. The
colour
ealer
(SQA)
2%
indicates
that the
transverse
field is the very source of bimodal signature.
simulated
quantum
annealing,
and
ound
in aEpixel
1.5%the device. However,
i
To%conrm%our%in
over
16
ranTo
understand
this
better,
a closer look at the model shows that an important!
effect
ase can Tbe expected from inspection
1% of their respective
y%%%%%%%%%%%%%%%.%%%
of
transverse
field
is
to
provide
gradation
in
the
magnitude
of
z-components
of
spins.
in
panel A, in figure 1, there is no
high%correlaHon%wi
0.5% apparent correlation
distributions
0% illustrates typical states of spins with and without transverse field.
Figure 6 schematically
hods..
Panel
also%achieves%high%
between the hard instances for the spin
dynamics
0% 0.05%
0.1% 0.15%model
0.2% 0.25% 0.3% 0.35% 0.4% 0.45% 0.5% 0.55% 0.6% 0.65% 0.7% 0.75% 0.8% 0.85% 0.9% 0.95%
Note that in the absence
of transverse
field, each spin will simply align in the direction
Wave
device,
Time%t%
andthe
theclassisuccess probability on the device, nor does there
een
of the z-component of the net field at that location. That is, the spin will want to point!Up%to%t=0.4,
correlation
appear
to
be
a
correlation
for
instances
of
intermediate
ome instances
~0.08
provablyofdeterministic
completely up or completely down
regardless
the actual strength of the field, as this is the!At%t=0.4,%let
ing
distance
hardness,
in contrast
to
the
correlations
seen
in
panel
A).
eld
dynamics,
configuration that minimizes energy. ~0.15
The empirically
presence ofdeterministic
a transverse field completely alters
for
NHowever,
= 108there are
Similarly,
poor
correlations
[22]
with
a
classical
vice.
branching of fixed
pointsthe%temperatu
this phenomenon, because the relative strengths of the z-0.15~0.4
and x- components
determine
dynamics model
of reference
[23].of the spin. behaves like SA 0.4~
d)spin
instances
non;evolving%
heir
respective
the preferred
direction
z;component%

Evolution of the system

Energy%(GHz)%

scent%of%the%relaHonship%between%cuts%and%eigenvectors%
n%spectral%graph%theory.%

Discussion

on of the system

emperature,%i.e.%in%the%absence%of%noise,%the%simulaHon%
wed%as%a%xed%point%iteraHon%procedure.%%
orem%in%topology,%all%xed%points%(equilibria)%are%
from%the%starHng%point.%
%Hme%t%changes,%the%model%traces%the%xed%points%
The
correlations between the simulated classical anent
correlation
The
colour
theoretically equivalent to SA 0.6~
)%of%the%Hme;dependent%update%funcHon.%
nealer
and the D-Wave device, shown in the supplemennamics
model
nd
in a pixel

tary
material,
nor
does
there are significantly worse than between SQA9
and the device.
f intermediate
Annealing'schedule'of'D/Wave'(approximate)'
next
provide
evidence for the bimodality being due
en
panel
A).
)% inWe
models success probability
the
with
classical eects. Our first evidence comes
2.7 from 13.3
)% to aquantum
n the
classi(t=0.15)
(t=0.20)
simulated quantum annealer. When lowering the temper5.6
28.1
provided
in [3], which is
1
instances
thermal
tunnel- (t=0.30)
de ature
classical
an- updates are suppressed,
(t=0.175)
(t=0.10)quantum
ments
of [1].
ing
dominates
thermal barrier crossing, and we observe
hedynamics,
supplemen# of fixed points
steps.
a
stronger
bimodality;
indeed a similar bimodal
distribetween
SQA
. However,
(empirical data)
bout
0.89.
bution arises also in an ensemble of (zero-temperature)

28.3
(t=0.40)

!Resume%the
transverse%e
evoluHon%is%eq
!The%number%of%branches%is%
at%an%already%
surprisingly%low,%considering%the%
problem%size%(N=108).%
!The%model%indica
!Olen%there%exists%a%strongly%
stable%local%congu
preferred%branch,%making%the%
breaking%their%two
process%even%more%determinisHc
other%supernodes.%

Figure 7, in which we take a snapshot of a typical simulation and plot z-components


of the spins, confirms that the above intuition is true. The values of z-components are
concentrated around 1 and 1 when there is no transverse field (red line), whereas they
are more evenly distributed when there is transverse field (blue line). From an algorithmic
viewpoint, it is worth drawing an analogy to spectral graph theory, in particular to the
relationship between cuts and eigenvectors of graphs. In spectral graph theory, a cut of a
graph is represented by a {1, 1}-valued vector, while entries of an eigenvector of a graph
can take on arbitrary real values. It is a well known fact that there exists a very interesting
relationship between the two, which allows us to find a sparse cut in a given graph using
information about its eigenvectors [11].
Why does the transverse field cause the model to behave more deterministically? Here
we make the observation that a simulation of our model can be viewed as a fixed point
iteration procedure which traces the equilibria of the time-dependent Hamiltonian. Note
that there is a unique equilibrium at t = 0, namely the configuration where every spin
points in the x direction. In fact, it is straightforward to prove that under the annealing
schedule of Figure 2 and the assumption that the underlying interaction graph is a subgraph
of the Chimera graph, H(t) has a unique equilibrium when t < 0.02. Moreover, we can
invoke a theorem in topology [10, 21] to prove that all later equilibria are reachable from
this unique equilibrium.
If we set the noise term in our model to 0, the evolution of the system is completely
deterministic and therefore the same equilibrium will be reached on each run. In this case,
simulation results are similar to that of [23] and it fails on too many instances to correlate
well with the D-Wave machine. When there is noise, however, different equilibria are
reached with different probabilities. Interestingly, our results (Figure 5) show that success
probabilities are greatly enhanced with the introduction of noise. To obtain an intuitive
understanding of this phenomenon, we designed the following experiment:
1. Up to t = tthres , simulate our model as usual, with noise.
2. At t = tthres , equilibrate the system under H(tthres ).
3. Resume the original simulation from t = tthres , but with the transverse field A(t)
completely turned off.
Remarkably, we were able to obtain a very high correlation of R 0.99 between our
original model and this modified model using tthres = 0.15. (The correlation with D-Wave
One remained at about 0.88.) This result suggests the following picture for our model.
In the first part of the evolution, i.e. when t < 0.15, we can understand the system to be
making a probabilistic choice of one of the equilibria of H(tthres ). The probability with
which each equilibrium is reached is governed by the behavior of the noise model. On the
other hand, the second part of the evolution, i.e. when t > 0.15, can be understood as
simulated annealing that begins at an effective temperature T /B(t) which is already quite
low. This means that it can roughly be thought of as local search with small perturbations
which explores a very small neighborhood of the given state. Therefore, this algorithm will
10

succeed if and only if the first part of the evolution brings the system into a state which is
sufficiently close to the global optimum.
Another result of our simulations was that the number of different equilibria reached in
the first part of the process was often surprisingly low. Averaged across all instances, we
only saw 20.2 distinct equilibria at t = 0.15, which indicates that the size of the state space
explored in the first part of the evolution is not comparable with our usual expectations
for a problem of size n = 108. Moreover, in most cases we found that there existed a
few equilibria that were strongly preferred to the others, making the state space even
smaller in effect and the process more deterministic. Together with the picture suggested
in the previous paragraph, this offers an approximate high-level explanation of the bimodal
behavior of our model.
An interesting question that remains to be studied is what happens at the so-called
branching points, namely at those values of t where the number of equilibria of H(t)
jumps from 1 to 2, 2 to 3, and so on. In other words, what choices are being made at
those branching points and in what ways do distinct branches differ? A case study on
the instance 13-55-29 offers a very interesting insight into this question. The instance has
two distinct equilibria at t = 0.054, up to the two-fold symmetry of z-flips of all spins.
Figure 8 shows how the two equilibria differ from each other. The figure suggests that the
choice being made at this particular branching point is the orientation between the white
cluster and the black cluster. Note that if a wrong branch is selected here, the second
part of the process will not be able to correct it, as it will have to flip the entire black
cluster. This is because as we already saw, the second part of the process may be thought
of as simulated annealing at low temperature, and in this case the number of spins that
have to be flipped simultaneously is too large.
Similar analysis on other instances and other values of t confirms this intuition; branching points always correspond to points at which the orientation between large clusters gets
determined. Moreover, as t increases, we see smaller and smaller clusters become involved
in the process. That is, we eventually see branching points which determine the orientation
between one or two supernodes and the rest of the spins. From an algorithmic viewpoint,
this phenomenon has significant implications. Namely, it indicates that supernodes with
highly stable local configurations get fixed early in the process, and therefore the main
challenge in finding the global optimum lies in breaking their two-fold symmetry based on
interaction with other supernodes. Note that the energy contribution from interactions
within a particular supernode remains unchanged if every spin in the supernode is flipped.
However, the energy contribution from interactions between supernodes depends crucially
on the relative orientation between those supernodes.
Since simulated annealing performed in the second part of the process is not able
to correct more than a handful of spins at once, the main difficulty of the problem lies
in choosing a correct branch in the first part of the process. However, we have seen that
different branches roughly correspond to different choice of orientation between supernodes.
This observation has an important implication for the computational problem being solved,
11

Figure 8: Branching of equilibria (Instance 13-55-29, t = 0.054). Black dots indicate


spins for which the sign of z-components differ between the two branches. Figure shows
the less frequent of the two branches. Blue edge indicates ferromagnetic interaction and
red edge antiferromagnetic interaction.
because it provides a sense in which the effective problem size may be thought of as closer
to the number of supernodes m = 16 rather than the number of spins n = 108. This could
be one explanation of the fact that heuristic methods such as simulated annealing perform
rather well on this type of problems despite the large problem size. Therefore, to the extent
that our classical model accurately represents the dynamics of the D-Wave machine, it is
reasonable to conjecture that the D-Wave machine, as well as any other heuristic method,
will fail to solve these problems effectively when the number of supernodes m becomes
sufficiently large.
It is worth emphasizing that the goal of this paper is not to provide a classical model
12

Figure 9: Correlation between simulated quantum annealing of [8] and our classical model.
The correlation coefficient R is about 0.99.
for the D-Wave machine, but rather to critically examine the claim that there is no simple
local classical model that correlates with its input-output behavior. The classical model
introduced here is useful for the purposes of studying the large-scale algorithmic features
of the D-Wave machine. The task of finding an accurate model for the D-Wave machine
(classical, quantum or otherwise), would be better pursued with direct access, not only to
programming the D-Wave machine, but also to its actual hardware.

Conclusions

This paper examined the interesting claims from a recent paper by [7] based on extensive experiments, that the 108 qubit D-Wave One machine exhibits large-scale quantum
behavior. The nature of the argument in that paper is algorithmic: a strong correlation
was observed between the input-output behavior of the D-Wave machine with a quantum
model called quantum simulated annealing, whereas it correlated poorly with two classical
models: simulated annealing and classical spin dynamics. In this paper, we suggest a simple classical model, which treats each spin in the D-Wave machine as a classical magnet,
which couples to neighboring magnets as well as an external field. The evolution of the
system proceeds according to a Metropolis-type dynamics. Simulation of this model on
the same instances reported in [7] yields higher correlation with the D-Wave input-output
behavior than those of simulated quantum annealing. One way to view our model is as a
classical mean field approximation to simulated quantum annealing. A direct comparison
between our model and simulated quantum annealing [24] reveals an extremely high correlation of R 0.99 (Fig 9). Hence, the correlation of the D-Wave data with simulated
13

quantum annealing does not impose any barrier to a classical model for D-Wave.
Further analysis of the new model indicates that the dynamics do not exhibit the the
kind of combinatorial explosion one might expect from a 108-bit problem. The reason
appears to be related to the 8-bit clusters in the interaction graph that appear to have
the effect of reducing the effective problem size to 16 rather than 108. To the extent that
this model captures the essential algorithmic features of the D-Wave One machine, this
suggests that scaling to machines with larger numbers of qubits should lead to increasing
problems with combinatorial explosion.

Acknowledgments
SWS and UV were supported by ARO Grant W911NF-09-1-0440 and NSF Grant CCF0905626. We thank Matthias Troyer, Sergio Boixo, and Troels F. Rnnow for useful discussions and for sharing their unpublished experimental data with us.

References
[1] S. Aaronson. D-Wave: Truth finally starts to emerge. http://www.scottaaronson.
com/blog/?p=1400, 2013.
[2] D. Aharonov, M. Ben-Or, and E. Eban. Interactive proofs for quantum computation.
In Proceedings of Innovations of Computer Science, pages 453469, 2010.
[3] B. Altshuler, H. Krovi, and J. Roland. Anderson localization casts clouds over adiabatic quantum optimization. Proceedings of the National Academy of Sciences of the
United States of America, 107(28):1244612450, 2010.
[4] J. Barrett, L. Hardy, and A. Kent. No signalling and quantum key distribution.
Physical Review Letters, 95(010503), 2005.
[5] S. Barz, J. F. Fitzsimons, E. Kashefi, and P. Walther. Experimental verification of
quantum computation. Nature Physics, 9:727731, 2013.
[6] S. Boixo, T. Albash, F. M. Spedalieri, N. Chancellor, and D. A. Lidar. Experimental
signature of programmable quantum annealing. Nature Communications, 4:2067, 2012.
[7] S. Boixo, T. F. Rnnow, S. V. Isakov, Z. Wang, D. Wecker, D. A. Lidar, J. M. Martinis,
and M. Troyer. Quantum annealing with more than one hundred qubits. E-print
arXiv:1304.4595. http://arxiv.org/pdf/1304.4595v2.pdf (An updated version will
appear in [8]), 2013.
[8] S. Boixo, T. F. Rnnow, S. V. Isakov, Z. Wang, D. Wecker, D. A. Lidar, J. M. Martinis,
and M. Troyer. Evidence for quantum annealing with more than one hundred qubits.
Nature Physics, in press, 2014.

14

[9] A. Broadbent, J. F. Fitzsimons, and E. Kashefi. Universal blind quantum computation.


In Proceedings of the 50th Annual IEEE Symposium on Foundations of Computer
Science, pages 517526, 2009.
[10] F. E. Browder. On continuity of fixed points under deformations of continuous mappings. Summa Brasil Math, 4:183191, 1960.
[11] F. Chung. Spectral Graph Theory. American Mathematical Society, 1992.
[12] N. G. Dickson et al. Thermally assisted quantum annealing of a 16-qubit problem.
Nature Communications, 4:1903, 2013.
[13] E. Farhi, J. Goldstone, S. Gutmann, and M. Sipser. Quantum computation
by adiabatic evolution. E-print arXiv:quant-ph/0001106. http://arxiv.org/pdf/
quant-ph/0001106v1.pdf, 2000.
[14] J. F. Fitzsimons and E. Kashefi. Unconditionally verifiable blind computation. E-print
arXiv:1203.5217. http://arxiv.org/pdf/1203.5217v2, 2012.
[15] N. Jones. Google and NASA snap up quantum computer D-Wave Two.
http://www.scientificamerican.com/article.cfm?id=
google-nasa-snap-up-quantum-computer-dwave-two, 2013.
[16] C. C. McGeoch and C. Wang. Experimental evaluation of an adiabiatic quantum
system for combinatorial optimization. In Proceedings of the 2013 ACM Conference
on Computing Frontiers, 2013.
[17] S. Pironio et al. Random numbers certified by Bells theorem. Nature, 464:10211024,
2010.
[18] B. W. Reichardt, F. Unger, and U. Vazirani. Classical command of quantum systems.
Nature, 496:456460, 2013.
[19] T. F. Rnnow, Z. Wang, J. Job, S. Boixo, S. V. Isakov, D. Wecker, J. M. Martinis, D. A. Lidar, and M. Troyer. Defining and detecting quantum speedup. E-print
arXiv:1401.2910. http://arxiv.org/pdf/1401.2910v1.pdf, 2014.
[20] R. Saket. A PTAS for the classical Ising spin glass problem on the Chimera graph
structure. E-print arXiv:1306.6943. http://arxiv.org/pdf/1306.6943v2, 2013.
[21] H. Schirmer. Fixed point sets of homotopies.
108(1):191202, 1983.

Pacific Journal of Mathematics,

[22] A. Selby. D-Wave: comment on comparison with classical computers.


http://www.archduke.org/stuff/d-wave-comment-on-comparison/, 2013.
[23] J. A. Smolin and G. Smith. Classical signature of quantum annealing.
arXiv:1305.4904. http://arxiv.org/pdf/1305.4904v1.pdf, 2013.

E-print

[24] M. Troyer. Personal communication, 2014.


[25] W. van Dam, M. Mosca, and U. Vazirani. How powerful is adiabatic quantum computation? E-print arXiv:0206003. http://arxiv.org/pdf/quant-ph/0206003v1, 2002.
15

[26] U. V. Vazirani and T. Vidick. Certifiable quantum dice - or, exponential randomness
expansion. In ACM Symposium on Theory of Computing, 2012.
[27] U. V. Vazirani and T. Vidick. Fully device-independent quantum key distribution. In
The 5th Innovations in Theoretical Computer Science, 2014.
[28] L. Wang, T. F. Rnnow, S. Boixo, S. V. Isakov, Z. Wang, D. Wecker, D. A. Lidar, J. M.
Martinis, and M. Troyer. Comment on: classical signature of quantum annealing.
E-print arXiv:1305.5837. http://arxiv.org/pdf/1305.5837v1.pdf, 2013.

16

You might also like