Probing the Non-Classicality of Temporal Correlations
Martin Ringbauer
1 2 3
and Rafael Chaves
4
1
Centre for Engineered Quantum Systems, School of Mathematics and Physics, University of Queensland, Brisbane, QLD 4072, Australia
2
Centre for Quantum Computer and Communication Technology, School of Mathematics and Physics, University of Queensland, Brisbane,
QLD 4072, Australia
3
Institute of Photonics and Quantum Sciences, School of Engineering and Physical Sciences, Heriot-Watt University, Edinburgh EH14
4AS, UK
4
International Institute of Physics, Federal University of Rio Grande do Norte, 59070-405 Natal, Brazil
November 15, 2017
Correlations between spacelike separated
measurements on entangled quantum systems
are stronger than any classical correlations and
are at the heart of numerous quantum tech-
nologies. In practice, however, spacelike sepa-
ration is often not guaranteed and we typically
face situations where measurements have an
underlying time order. Here we aim to pro-
vide a fair comparison of classical and quan-
tum models of temporal correlations on a sin-
gle particle, as well as timelike-separated cor-
relations on multiple particles. We use a causal
modeling approach to show, in theory and
experiment, that quantum correlations out-
perform their classical counterpart when al-
lowed equal, but limited communication re-
sources. This provides a clearer picture of the
role of quantum correlations in timelike sepa-
rated scenarios, which play an important role
in foundational and practical aspects of quan-
tum information processing.
Quantum predictions are fundamentally incompat-
ible with the intuitive notion of cause and effect that
underpins all of classical empirical science, as well
as everyday experience. Most famously, Bell’s the-
orem [1] shows that the correlations revealed by mea-
surements on distant parts of entangled quantum sys-
tems cannot be explained in causal terms, under the
assumptions that local measurement outcomes are not
influenced by spacelike separated events—local causal-
ity—and that measurement settings can be chosen
freely. This phenomenon has become known as quan-
tum nonlocality, and has been extensively studied for
variants of the scenario above, where spacelike sepa-
rated parties perform measurements on a composite
quantum system [2].
Yet, in practice we often face situations where
spacelike separation between observers is not guar-
anteed and instead there is some time order underly-
ing the observed physical events. Indeed, it has been
shown that a single quantum system measured at dif-
ferent points in time exhibits Bell-like correlations [3
8]. However, in such a scenario one cannot appeal to
relativity to rule out communication between the time
steps. This can be problematic, since a classical model
with sufficient communication can simulate any cor-
relations. Recent research has thus been devoted to
clarifying to what extent the incompatibility between
classical and quantum descriptions encountered in the
spacelike case carries over to the timelike scenario.
Specifically, one research direction has been to
study how limited resources such as entropy [9], mem-
ory [1012] or dimension [1315] can lead to a quan-
tum advantage over classical models. On the other
hand, one can relax the local causality assumption in
Bell’s theorem, aiming to explain quantum correla-
tions by classical models augmented with communi-
cation [1625]. However, communication is typically
only allowed for the classical system, leading to an un-
fair comparison between classical and quantum mod-
els, and it remains unclear to what extent these re-
sults hold when equal communication power is given
to quantum mechanics [23].
Here we use a causal modeling approach to allow
a fair comparison of classical and quantum models
of timelike separated correlations. First, we show
that for two time-ordered spatially separated mea-
surements, augmented with a limited amount of clas-
sical communication, quantum correlations outper-
form their classical counterpart. Second, we show
that, in contrast to spatial Bell-type scenarios [26],
there are faithful classical causal models reproducing
all the temporal correlations obtained from a series of
projective quantum measurements on a single quan-
tum system. However, we find that non-classical cor-
relations can arise in this scenario, when considering a
slightly weaker classical causal model, that is nonethe-
less strictly stronger than previous results [27], which
are contained as a special case. Finally we derive Bell-
type inequalities for the above scenarios and demon-
strate that they are violated in a photonic experiment.
1
arXiv:1704.05469v2 [quant-ph] 14 Nov 2017
1 Causal modeling and timelike Bell-
scenarios
In the following we employ the formalism of Bayesian
networks [28], which provides a natural framework for
classical causal modeling. A central concept in this
framework is that of a directed acyclic graph (DAG),
which consists of a set of nodes, representing the rel-
evant random variables
1
in the considered situation,
and directed edges, representing the causal relations
between those variables. A set of variables X
1
, . . . , X
n
forms a Bayesian network with respect to some DAG
if and only if the probability distribution p(x
1
, . . . , x
n
)
can be decomposed as
p(x
1
, . . . , x
n
) =
n
Y
i=1
p(x
i
|pa
i
) (1)
where P A
i
stands for the set of graph-theoretical par-
ents of the variable X
i
(i.e. all variables that have a
direct causal influence over X
i
). Without loss of gen-
erality each variable can be understood as a determin-
istic function of its parents plus local noise U
i
that
supplies potential randomness, x
i
= f
i
(pa
i
, u
i
). This
formalism thus enables a distinction between simple
statistical correlations and actual causation by explic-
itly specifying the underlying mechanism generating
the data.
Here we are interested in DAGs containing so-called
latent variables, which are empirically inaccessible. In
the context of Bell’s theorem [1] these are also known
as hidden variables. For any set of observed corre-
lations, there are in general many DAGs with hid-
den variables that could have produced these obser-
vations. Among these, causal inference is particularly
interested in those fulfilling the conditions of minimal-
ity and faithfulness. Minimality requires that, given
two possible causal models, we choose the simplest
one, capable of generating the smallest set of correla-
tions (including the observed one). In turn, faithful-
ness, requires the causal model to be able to explain
the observed data without resorting to fine-tuning of
the causal-statistical parameters. In other words, any
observed (conditional) independence should be a con-
sequence of the causal structure itself, rather than a
specific choice of parameters. Faithful (i.e. non-fine-
tuned) models are therefore robust against changes in
the causal parameters and thus the preferred choice.
To illustrate the last point, consider the paradig-
matic causal structure of Bell’s theorem in Fig. 1a.
This structure intuitively reflects the causal assump-
tions of Bell’s theorem, leading to the so-called local
hidden-variable (LHV) models. First, the two parties
are assumed to be spacelike separated, such that the
correlations between the measurements outcomes A
1
We adopt the standard convention that uppercase letters
label random variables while their values are denoted in lower
case.
and B can only be mediated via a common source
Λ, implying that p(a, b|x, y, λ) = p(a|x, λ)p(b|y, λ).
Second, it is assumed that the experimenters can
freely choose which observables to measure (repre-
sented by the random variables X and Y ), inde-
pendently of how the system was prepared, that is,
p(x, y, λ) = p(x, y)p(λ). Note that these constraints
implied by the causal model appear at an unobserv-
able level since they explicitly involve the hidden vari-
ables Λ. Yet, they also imply observable constraints
in the form of no-signaling conditions, expressed as
p(a|x, y) = p(a|x) and p(b|x, y) = p(b|y), and Bell
inequalities [1, 29].
A
X
B
Y
Λ
A
X
B
Y
Λ
M
ba
Figure 1: Causal structure underlying Bell’s theorem. (a)
Two observers, Alice and Bob, each have the choice of two
measurements represented by the random variables X and Y ,
respectively. The correlation between their measurement out-
comes, modeled as random variables A and B, respectively,
are mediated solely by a common cause in their past—the
hidden variable Λ. (b) Bell’s causal model augmented with
one-way communication from Alice to Bob. The initial state
of the joint system is specified by the ontic state Λ. First,
Alice performs a measurement with setting x, obtaining out-
come a. She then sends a message m to Bob, who performs
a measurement with setting y, obtaining outcome b.
While quantum correlations obey the no-signaling
conditions, they violate Bell inequalities [1, 29] and
are thus in conflict with the assumptions behind the
causal structure in Fig. 1a. In order to maintain a
classical causal explanation, the model in Fig. 1a must
therefore be augmented with additional resources;
something that can only be done at the cost of in-
troducing fine-tuning [26]. For instance, the causal
structure in Fig. 1b can reproduce all quantum corre-
lations, but at the same time allows, in principle, for
non-local correlations between X and B. Hence, in
order to satisfy the no-signaling condition p(b|x, y) =
p(b|y) the causal parameters must be chosen from a
set of measure zero [28], a signature of fine-tuning.
Studying such non-local classical models can pro-
vide valuable insights into the relation between clas-
sical and quantum theory, and their applications [2].
However, at the same time such models lead to an
unfair comparison, since allowing for communication
makes not only classical, but also quantum models
more powerful. In practice, it is more natural to
assume a certain underlying causal structure, and
ask what can be achieved with classical and quan-
tum resources? Bell’s theorem is a particular case
of this broader question, referring to spacelike sep-
arated events. However, there are often situations
2
where the events are timelike rather than spacelike
ordered. Examples include central quantum infor-
mation tasks, such as teleportation [30], superdense
coding [31], and measurement-based quantum compu-
tation [32], as well as prepare-and-measure scenarios
[13], sequential Bell scenarios [33, 34] and a sequence
of measurements on a single quantum system [27].
2 Non-classicality of timelike correla-
tions augmented by communication
Consider the scenario in Fig. 1b, where two distant
parties, Alice and Bob, share pre-established cor-
relations (represented by Λ) and are allowed one-
way communication (the message M). As shown in
Ref. [16], a classical model of this form (for a large
enough message M) is enough to reproduce all the
correlations obtainable from local measurements on
two-qubit entangled states as in Fig. 1a, which are
described by
p(a, b|x, y) = Tr
(M
a
x
M
b
y
)ρ
, (2)
where M
a
x
and M
b
y
are measurement operators for Al-
ice and Bob, respectively, and ρ is the density matrix
describing the shared quantum state. If, in contrast,
we impose the causal structure of Fig. 1b also to the
quantum case—that is, Bob’s measurement may de-
pend on Alice’s measurement setting and outcome—
then the set of correlations is described by
p(a, b|x, y) =
X
m
Tr
(M
a
x
M
b
y,m
)ρ
. (3)
Note that Bob’s measurement operator now explicitly
depends on the values of X and A (via the message
M)
2
. Clearly, if there are no restrictions on the di-
mensionality of the message, every distribution of the
form above can be also obtained by the classical hid-
den variable model in Fig. 1b, that is, by a model
respecting the decomposition
p(a, b|x, y) =
X
m,λ
p(λ)p(m|x, a)p(a|x, λ)p(b|y, m, λ).
(4)
In fact, it is enough to choose m = x to reproduce
all possible one-way signalling correlations. To see
this, note that the quantum correlations arising from
Eq. (3) are of the form p(a, b|x, y) = p(b|x, y, a)p(a|x),
and that a can be made a deterministic function a =
f
a
(x, λ). Hence a does not carry any information that
is not already contained in λ and x.
Notwithstanding, this picture changes if we impose
restrictions on the message sent from Alice to Bob.
Consider that each party measures three dichotomic
2
In fact, Bob could apply to his share of the joint quantum
state any completely-positive trace-preserving map dependent
on the message m.
observables (i.e. x, y = 0, 1, 2 and a, b = 0, 1) and
that Alice is bound to send a binary message (m =
0, 1). In this case, every classical model must obey
the inequality [16]
S
1bit
= hA
0
B
0
i + hA
0
B
1
i + hA
0
B
2
i + hA
1
B
0
i (5)
+ hA
1
B
1
i hA
1
B
2
i + hA
2
B
0
i hA
2
B
1
i 6.
Furthermore, it was shown in Ref. [16] that this in-
equality also holds for correlations from local mea-
surements on entangled quantum states, while it
can be violated by more powerful no-signalling cor-
relations. Hence, while one bit of communication
is in this scenario sufficient for a classical model
to simulate quantum correlations (without commu-
nication), it is not sufficient to simulate all pos-
sible no-signalling correlations. However, just like
communication-augmented classical models become
more powerful, so do quantum models. Specifically,
local measurements on entangled quantum states aug-
mented with one bit of communication can indeed
violate inequality (5) [23], thus showing that under
fair comparison, quantum advantage persists in such
a timelike-separated Bell scenario.
state preparation
Alice
Bob
GT
HWP
APD
Source
QWP
PPBS
Figure 2: Experimental setup for studying the non-
classicality of timelike correlations. Pairs of single photons
are produced via spontaneous parametric downconversion in
a β-barium borate crystal (not shown). The two photons
are entangled to arbitrary degree using a non-deterministic
controlled-NOT gate (CNOT), based on nonclassical interfer-
ence in a partially polarizing beam splitter (PPBS) [35]. Alice
and Bob then perform local projective measurements on their
share of the entangled state, which are implemented using a
set of half- (HWP) and quarter-waveplates (QWP), a Glan-
Taylor polarizer (GT) and single-photon counters (APD).
As an example, consider that Alice and Bob share a
maximally entangled state |Ψ
+
i = (|00i + |11i)/
2).
Alice performs local measurements with settings A
0
=
A
1
=
ˆ
X, A
2
=
ˆ
Z and encodes her measurement set-
ting in a message m to Bob. For x = 0 she sends
m = 0 and for x = 1 or x = 2 she sends m = 1. As-
suming that all inputs are equally likely this message
has an entropy of H(m) 0.92 and thus contains less
than 1 bit of information. If Bob receives m = 0, he
measures B
0
= B
1
= B
2
=
ˆ
X, while for m = 1 he
measures B
0
= (
ˆ
X +
ˆ
Z)/
2, B
1
= (
ˆ
X
ˆ
Z)/
2 and
B
2
=
ˆ
X. This protocol achieves S
1bit
= 4
2, thus
violating inequality (5).
Experimentally we can test inequality (5) with
qubits encoded in the polarization of single photons,
see Fig. 2. Two single photons are first entangled
using a non-deterministic controlled-NOT gate, and
3
then distributed to Alice and Bob. By varying the in-
put states this configuration can produce states with
arbitrary degree of entanglement, quantified by the
concurrence C [36], see Appendix B for more details.
For simplicity, the message m has been directly taken
into account in Bob’s measurement basis. The ex-
perimental results in Fig. 3 show a clear violation of
inequality (5), up to S
1bit
= 6.66
+0.02
0.02
.
fixed
optimized
Figure 3: Experimental test of inequality (5) for
temporally-separated spatial correlations. The inequality
is tested for a family of states with varying degree of en-
tanglement as quantified by the concurrence. Data points
include 3σ statistical error bars and the theory prediction is
shown as solid lines. Data points in the grey region are com-
patible with inequality (5), while it is violated for data in
the white region. Blue data and theory corresponds to the
fixed measurement scheme outlined in the text, whereas the
orange data and theory is obtained when the measurement
settings are optimized for a given non-maximally entangled
state, see Appendix B for more details.
3 Classical causal models for a se-
quence of projective measurements
Above we have analyzed the correlations between two
timelike separated parties, whose measurements fol-
low an underlying time order. We will now focus on
a series of projective measurements performed on a
single quantum system. A measurement with setting
x, obtaining outcome a, is described by a set of pro-
jective operators M
x
a
, such that after the measure-
ment the system, initially in state ρ
1
, is left in a state
given by ρ
2
= ρ
x
a
= M
x
a
. In a next time step a mea-
surement with setting y, producing outcome b is per-
formed, leaving the system in a state ρ
3
= ρ
y
b
= M
y
b
,
and so forth.
Similar to a standard Bell experiment, the clas-
sical causal description of this scenario, illustrated
in Fig. 4, involves a random variable Λ
1
—the ontic
state [37]—which fully specifies the initial state of
the system. The probability that a measurement x
produces the outcome a is then given by p(a|x) =
P
λ
1
p(a|x, λ
1
)p(λ
1
). Here we have explicitly used
the measurement independence assumption [22, 38]
p(λ
1
|x) = p(λ
1
), that the measurement setting can
be chosen independently of how the system is pre-
pared. After the measurement, the system will be in
a potentially different state Λ
2
. For a projective mea-
surement, Λ
2
is fully specified by the setting and out-
come of the preceding measurement, and does not de-
pend directly on the pre-measurement state Λ
1
, that
is p(λ
2
|x, a, λ
1
) = p(λ
2
|x, a). This implies that all
correlations between Λ
1
and Λ
2
are mediated via the
measurement. To see this, note that the quantum
probability distribution after a series of three sequen-
tial measurements is given by
p
Q
(a, b, c|x, y, z) = p(c|b, y, z)p(b|a, x, y)p(a|x) , (6)
where p(c|b, y, z) = tr [M
z
c
ρ
y
b
], p(b|a, x, y) = tr [M
y
b
ρ
x
a
]
and p(a|x) = tr [M
x
a
ρ
1
]. In particular, any poten-
tial correlation between the measurement outcome
c in the third time step with the measurement set-
ting and outcome in the first time step (x and a, re-
spectively), is screened-off by the intermediate mea-
surement setting and outcome (y and b), that is,
p(c|a, b, x, y, z) = p(c|b, y, z). This is similar to the
no-signalling constraints arising in a Bell scenario and
thus imposes restrictions to the classical causal mod-
els describing such a scenario. Specifically, a causal
model that reproduces this independence without re-
sorting to fine-tuning cannot contain a causal link of
the form Λ
2
Λ
3
, since such a link can generate un-
wanted correlations between the variables X, A and
C (not mediated by Y, B). In fact, any model that
contains such a link, can only satisfy the condition
p(c|a, b, x, y, z) = p(c|b, y, z) by virtue of causal pa-
rameters chosen from a set of measure zero [28], that
is, the parameters are fine-tuned in such a way that
these correlations are hidden from the observational
data [26].
Following the above description, the case of 3 se-
quential projective measurements on a single qubit
can be represented in terms of the causal structure
in Fig. 4. The temporal correlations p(a, b, c|x, y, z)
compatible with this causal structure can then be de-
composed as (with straightforward generalization to
more time steps)
p(a, b, c|x, y, z) =
X
λ
1
2
3
p(a|x, λ
1
)p(b|y, λ
2
) (7)
p(c|z, λ
3
)p(λ
1
)p(λ
2
|x, a)p(λ
3
|y, b).
Note that without any restrictions on the dimension-
ality the hidden states Λ
i
can contain the full infor-
mation about the measurement settings and outcomes
of the previous time steps. This implies that, in con-
trast to the spacelike Bell scenario, all temporal cor-
relations of the form of Eq. (6) can be faithfully re-
produced by the classical causal model in Fig. 4.
This naturally raises the question whether further
restrictions on the classical causal model, might re-
veal a quantum advantage. Similar to the so-called
prepare-and-measure scenarios [13], one might expect
that quantum systems of a given dimension give rise
to measurement statistics that cannot be reproduced
4
A
X
B
Y
Λ
1
Z
C
Λ
2
Λ
3
t
1
t
2
t
3
Time
x a
y b z
c
-1
+1
b
a
-1
+1
-1
+1
Figure 4: A sequence of three projective measurement on
a single qubit. (a) A single quantum system is subject to a
sequence of projective measurements at times t
1
, t
2
, and t
3
with settings x, y, z, and obtaining outcomes a, b, c, respec-
tively. (b) A general classical causal model for the scenario
in (a). Note that, although causal graphs are formulated
without reference to any spacetime structure, here we have
drawn the graph such that the horizontal direction can be
identified with time.
by classical systems Λ
i
of the same dimension. Since
restriction on the dimension of Λ
i
in models of the
form of Fig. 4 would lead to non-convex sets that are
technically very challenging to characterize [39, 40],
one typically considers convex relaxations. The re-
sulting models contain the model of interest as a spe-
cial case, but allow for shared randomness between
the parties. For example, using the results of Ref. [17]
we show in the Appendix A that classical hidden
states of dimension 4 (two bits of classical informa-
tion), together with shared randomness between the
parties, are enough to reproduce all correlations from
a series of projective measurements on a single qubit.
For a fair comparison to a qubit, we then considered
hidden states of dimension two and could not find a
difference between quantum and classical correlations.
In light of these results it would be very interesting to
test the model of Fig. 4 with two-dimensional hidden
states and no shared randomness. Due to the com-
plexity of characterizing such non-convex sets, how-
ever, this remains an open question.
Besides restrictions on the dimension of the hid-
den state, certain physical constraints and assump-
tions might naturally lead to weaker causal models
for a sequence of projective measurements on a sin-
gle qubit. A well-known example is the Leggett-Garg
(LG) model [27] for testing macroscopic realism. This
model is based on the assumption of noninvasive mea-
surability, stating that it should be possible, in prin-
ciple, to determine (measure) the state of a system
without perturbing it. This is a special case of Eq. (7),
where the hidden variable state is unchanged by the
measurements and constant throughout the experi-
ment, that is, λ
i
= λ, which implies that
p(a, b, c|x, y, z) =
X
λ
p(a|x, λ)p(b|y, λ)
p(c|z, λ)p(λ).
(8)
Note that this is the usual local hidden variable de-
scription encountered in Bell’s theorem. Further,
since the expectation values of a sequence of projec-
tive measurements on a single qubit are the same as
for local measurements on a pair of entangled parti-
cles [6], quantum correlations also violate macroscopic
realism, manifest as violations of so-called Leggett-
Garg inequalities [41]. The result, however, relies
critically on the assumption of noninvasive measura-
bility, which is difficult to justify for a single quantum
system. In fact, quite generally, the correlations ob-
tained by sequential measurements on a quantum sys-
tem will display signaling (e.g p(b|x, y) 6= p(b|x
0
, y))
as opposed to the model in Eq. (8) that only allow
for non-signaling correlations. In other terms, the
Leggett-Garg model implies independence relations
that are not observed in the experiment. In this sense,
in order to test incompatibility with the Leggett-Garg
model we do not need to take the strength of the cor-
relations into account and simply look for violations
of independence relations implied by the model [42].
This raises the question of whether one can find
examples of temporal quantum experiments to which
classical faithful causal models can reproduce all in-
dependence relations while at the same time being
incompatible with the generated correlations. Next
we show that this is indeed the case.
4 Quantum incompatibility with a
weaker classical model for sequential
projective measurements
From a causal inference perspective, given some ob-
served probability distribution the goal is to find a
faithful causal model reproducing all the (conditional)
independence relations implied by this distribution.
As a concrete example, consider a sequence of three
projective measurements (see Fig. 4a) on an initial
maximally mixed qubit state ρ
1
= /2. It is easy to
verify that for arbitrary projective measurements, the
measurement outcome of the third measurement is in-
dependent of the setting of the first measurement, i.e.
p(x, c) = p(x)p(c). In this case, the causal model in
Fig. 4b is not faithful any longer because it allows for
correlations between the variables X and C. Instead,
the most general causal model reproducing such inde-
pendence relation is shown in Fig. 5 where, in com-
parison with the causal model in Fig. 4, the causal
link between the variable B and Λ
3
is removed. Any
distribution compatible with this model has a decom-
position given by
p(a, b, c|x, y, z) =
X
λ
1
2
3
p(a|x, λ
1
)p(b|y, λ
2
) (9)
p(c|z, λ
3
)p(λ
1
)p(λ
2
|a, x)p(λ
3
|y).
Crucially, this classical model faithfully captures the
observed independence p(x, c) = p(x)p(c) that holds
5
A
X
B
Y
Λ
1
Z
C
Λ
2
Λ
3
t
1
t
2
t
3
b
a
Alice
Bob Charlie
GT
HWP
APD
Source
QWP
Time
CNOT
Figure 5: The causal model of Eq. (9) for a sequence
of projective measurements. (a) Experimentally, the first
and last parties, Alice and Charlie, measure the polarization
of the photon directly, while the intermediate party, Bob,
implements a non-destructive measurement by coupling the
system to a meter photon using the CNOT gate in Fig. 2.
Note that, this experiment naturally satisfy the independence
p(c|x) = p(c). (b) The causal graph corresponding to Eq. (9)
faithfully captures this independence. This model is slightly
weaker than Fig. 4, since the ontic state Λ
3
depends only on
the previous measurement setting Y , but not on the outcome
B.
for a wide range of relevant experimental scenarios.
This includes arbitrary measurements on an initially
maximally mixed state, as well as arbitrary initial
states in the xy-plane of the Bloch-sphere for the mea-
surements in the experimental implementation below.
As detailed in Appendix A any correlations com-
patible with Eq. (9) must respect the inequality
S
τ
3
= hA
0
B
0
i + hA
0
B
1
i + hA
1
B
0
i hA
1
B
1
i (10)
hBC
001
i hBC
101
i hBC
010
i hBC
110
i 6,
where the joint expectation values are defined as
hA
x
B
y
i =
P
a,b
(1)
a+b
p(a, b|x, y) and hBC
xyz
i =
P
a,b,c
(1)
b+c
p(a, b, c|x, y, z). This inequality can,
however, be violated by a sequence of projective mea-
surements on any initial qubit state. For instance,
choosing measurement settings A
0
=
ˆ
Z, A
1
=
ˆ
X,
B
0
= C
1
= (
ˆ
Z +
ˆ
X)/
2, and B
1
= C
0
=
(
ˆ
Z
ˆ
X)/
2 (where
ˆ
X and
ˆ
Z are the Pauli operators)
obtains a value of S
τ
3
= 4
2 > 6. Furthermore, for
any initial state in the xy-plane of the Bloch sphere,
the resulting probability distribution p(a, b, c|x, y, z)
respects the independence relation p(c|x) = p(c) im-
plied by the model under test. Through unitary ro-
tations this implies that for any fixed initial quantum
state, one can generate temporal correlations that
cannot be explained by non-fine-tuned models.
Experimentally we test inequality (10) with pho-
tonic polarization qubits for an initial maximally
mixed state, see Fig. 5a. For the intermediate mea-
surement the system is coupled to a meter in the
state |0i. A measurement of the meter in the com-
putational basis {|0i, |1i} achieves a projective mea-
surement of the system in a basis that is chosen by
appropriate single-qubit unitaries applied to the sys-
tem before and after the interaction. Notably, this
measurement design can be straightforwardly general-
ized to more than three parties by replicating the von
Neumann measurement. The experimental results in
Fig. 6 demonstrate a clear violation of inequality (10)
by a series of three projective measurements on a sin-
gle qubit, achieving a value of S
τ
3
= 6.65
+0.01
0.01
.
Figure 6: Experimental results for testing Inequality (10)
with a sequence of measurements on a single quantum
systems. The top part shows the experimental value of S
τ
3
(vertical black bar) together with the 3σ statistical confi-
dence region (grey shaded area). It also shows the individual
contributions from the 8 expectation values (alternate blue
and orange shadings), which are compared to their theoreti-
cal expectation in the bottom part of the figure. Here, blue
shaded bars correspond to theoretical expectation values for
ideal states and measurement and the horizontal black bars
represent the experimental values with 3σ statistical error
bars shown as grey shaded areas (almost within the black
bars).
5 Discussion
Our results show, both in theory and experiment, that
the discrepancy between the classical and quantum
descriptions typically associated with spatial correla-
tions extends to temporal correlation scenarios. The
latter arise naturally in situations where spacelike sep-
aration cannot be practically guaranteed, or when a
sequence of measurements is performed on the same
quantum system. In the case of spatial correlations
with an underlying time order, we have shown that
a Bell-type inequality, designed to test classical mod-
els augmented with limited communication, displays
a quantum violation if and only if the quantum pro-
tocol is also augmented with communication power.
This highlights that a quantum advantage persists in
a fair comparison of classical and quantum resources.
In a purely temporal-correlations scenario we have
shown that, as opposed to spatial Bell scenarios [26],
there is a faithful (non-fine-tuned) classical model ca-
pable of simulating all correlations from projective
measurements on quantum states. It remains an
6
open question whether these models are equivalent
when imposing the same dimensionality to quantum
and classical models. Notwithstanding, we identified
quantum violations of inequalities associated with a
slightly less powerful classical model that includes the
Leggett-Garg model as special case [27]. We have
experimentally observed such a quantum violation,
which demonstrates a stronger form of non-classicality
of the correlations arising from a temporal sequence
of projective measurements. Furthermore, our exper-
imental design allows for tuning of the intermediate
measurement strength, which will enable future stud-
ies of non-Markovian models with some residual cor-
relations between the ontic states Λ
i
over multiple
time-steps [43].
Our theoretical results are based on a causal mod-
eling approach and thus formulated without reference
to a background spacetime structure. The direction
of time is only implicitly deduced from the flow of
information between the parties or sequence of mea-
surements. Experimentally, the considered scenarios
are not subject to the locality loophole. However, the
results rely on the related assumption that there is no
hidden communication channel—other than the ones
implied by the models in Figs. 1b and 4—between
the different time steps of the scenario under consid-
eration. Practically, we also rely on a fair-sampling
assumption to contend with imperfect detection effi-
ciencies.
Temporal correlation scenarios play an important
role in communication complexity problems [44] and
in the search for a physical principle behind quan-
tum nonlocality, such as information causality [45].
Our results thus provide an avenue towards a more
systematic understanding of the quantum advantage
arising in such scenarios, that not only may lead to
new ways of processing information but also to new
insights into the nature of quantum correlations.
Acknowledgments
We thank C. Budroni, F. Costa, A. Fedrizzi,
and A.G. White for helpful discussions and feed-
back, and T. Vulpecula for experimental assistance.
This work was supported in part by the Centres
for Engineered Quantum Systems (CE110001013)
and for Quantum Computation and Communica-
tion Technology (CE110001027), the Engineering and
Physical Sciences Research Council (grant number
EP/N002962/1), the Brazilian ministries MEC and
MCTIC, the FQXi Fund, and the Templeton World
Charity Foundation (TWCF 0064/AB38).
References
[1] J. S. Bell, On the Einstein–Podolsky–Rosen para-
dox, Physics 1, 195 (1964).
[2] N. Brunner, D. Cavalcanti, S. Pironio,
V. Scarani, and S. Wehner, Bell nonlocal-
ity, Rev. Mod. Phys. 86, 419 (2014).
[3] C. Brukner, S. Taylor, S. Cheung, and V. Vedral,
Quantum Entanglement in Time, arXiv preprint
quant-ph/0402127 (2004).
[4] M. Markiewicz, P. Kurzyński, J. Thompson, S.-
Y. Lee, A. Soeda, T. Paterek, and D. Kasz-
likowski, Unified approach to contextuality, non-
locality, and temporal correlations, Phys. Rev. A
89, 042109 (2014).
[5] C. Budroni, T. Moroder, M. Kleinmann, and
O. Gühne, Bounding Temporal Quantum Corre-
lations, Phys. Rev. Lett. 111, 020403 (2013).
[6] T. Fritz, Quantum correlations in the temporal
Clauser–Horne–Shimony–Holt (CHSH) scenario,
New J. Phys. 12, 083055 (2010).
[7] A. Fedrizzi, M. P. Almeida, M. A. Broome, A. G.
White, and M. Barbieri, Hardy’s Paradox and
Violation of a State-Independent Bell Inequality
in Time, Phys. Rev. Lett. 106, 200402 (2011).
[8] M. Pawłowski, J. Kofler, T. Paterek,
M. Seevinck, and Č. Brukner, Non-local
setting and outcome information for violation
of Bell’s inequality, New J. Phys. 12, 083051
(2010).
[9] R. Chaves, J. B. Brask, and N. Brunner, Device-
Independent Tests of Entropy, Phys. Rev. Lett.
115, 110501 (2015).
[10] E. F. Galvão and L. Hardy, Substituting a Qubit
for an Arbitrarily Large Number of Classical Bits,
Phys. Rev. Lett. 90, 087902 (2003).
[11] A. Montina, Exponential complexity and ontolog-
ical theories of quantum mechanics, Phys. Rev.
A 77, 022104 (2008).
[12] M. Żukowski, Temporal inequalities for sequen-
tial multi-time actions in quantum information
processing, Front. Phys. 9, 629 (2014).
[13] R. Gallego, N. Brunner, C. Hadley, and A. Acín,
Device-Independent Tests of Classical and Quan-
tum Dimensions, Phys. Rev. Lett. 105, 230501
(2010).
[14] S. Brierley, A. Kosowski, M. Markiewicz, T. Pa-
terek, and A. Przysi¸eżna, Nonclassicality of
Temporal Correlations, Phys. Rev. Lett. 115,
120404 (2015).
[15] J. Bowles, N. Brunner, and M. Pawłowski, Test-
ing dimension and nonclassicality in communica-
tion networks, Phys. Rev. A 92, 022351 (2015).
[16] D. Bacon and B. F. Toner, Bell Inequalities with
Auxiliary Communication, Phys. Rev. Lett. 90,
157904 (2003).
[17] B. F. Toner and D. Bacon, Communication cost
of simulating Bell correlations. Phys. Rev. Lett.
91, 187904 (2003).
[18] S. Pironio, Violations of Bell inequalities as lower
bounds on the communication cost of nonlocal
correlations, Phys. Rev. A 68, 062102 (2003).
7
[19] T. Vértesi and E. Bene, Lower bound on the com-
munication cost of simulating bipartite quantum
correlations, Phys. Rev. A 80, 062316 (2009).
[20] M. J. W. Hall, Relaxed Bell inequalities and
Kochen-Specker theorems, Phys. Rev. A 84,
022102 (2011).
[21] K. Maxwell and E. Chitambar, Bell inequalities
with communication assistance, Phys. Rev. A 89,
042108 (2014).
[22] R. Chaves, R. Kueng, J. B. Brask, and
D. Gross, Unifying Framework for Relaxations
of the Causal Assumptions in Bell’s Theorem,
Phys. Rev. Lett. 114, 140403 (2015).
[23] J. B. Brask and R. Chaves, Bell scenarios with
communication, J. Phys. A 50, 094001 (2017).
[24] M. Ringbauer, C. Giarmatzi, R. Chaves,
F. Costa, A. G. White, and A. Fedrizzi, Ex-
perimental test of nonlocal causality, Sci. Adv. 2,
e1600162 (2016).
[25] R. Chaves, D. Cavalcanti, and L. Aolita, Causal
hierarchy of multipartite Bell nonlocality, Quan-
tum 1, 23 (2017).
[26] C. J. Wood and R. W. Spekkens, The lesson
of causal discovery algorithms for quantum cor-
relations: Causal explanations of Bell-inequality
violations require fine-tuning, New J. Phys. 17,
033002 (2015).
[27] A. J. Leggett and A. Garg, Quantum mechan-
ics versus macroscopic realism: Is the flux there
when nobody looks? Phys. Rev. Lett. 54, 857
(1985).
[28] J. Pearl, Causality (Cambridge University Press,
2009).
[29] J. S. Bell, The theory of local beables, Epistem.
Lett. 9, 11 (1976).
[30] C. H. Bennett, G. Brassard, C. Crépeau,
R. Jozsa, A. Peres, and W. K. Wootters,
Teleporting an unknown quantum state via dual
classical and Einstein-Podolsky-Rosen channels,
Phys. Rev. Lett. 70, 1895 (1993).
[31] C. H. Bennett and S. J. Wiesner, Communication
via one- and two-particle operators on Einstein-
Podolsky-Rosen states, Phys. Rev. Lett. 69, 2881
(1992).
[32] H. J. Briegel, D. E. Browne, W. Dür,
R. Raussendorf, and M. Van den Nest,
Measurement-based quantum computation, Nat.
Phys. , 19 (2009).
[33] S. Popescu, Bell’s Inequalities and Density Ma-
trices: Revealing “Hidden” Nonlocality, Phys.
Rev. Lett. 74, 2619 (1995).
[34] R. Gallego, L. E. Würflinger, R. Chaves, A. Acín,
and M. Navascués, Nonlocality in sequential cor-
relation scenarios, New J. Phys. 16, 033037
(2014).
[35] N. Langford, T. Weinhold, R. Prevedel,
K. Resch, A. Gilchrist, J. O’Brien, G. Pryde, and
A. White, Demonstration of a Simple Entangling
Optical Gate and Its Use in Bell-State Analysis,
Phys. Rev. Lett. 95, 210504 (2005).
[36] S. Hill and W. K. Wootters, Entanglement of a
Pair of Quantum Bits, Phys. Rev. Lett. 78, 5022
(1997).
[37] N. Harrigan and R. W. Spekkens, Einstein, In-
completeness, and the Epistemic View of Quan-
tum States, Found. Phys. 40, 125 (2010).
[38] M. J. W. Hall, Local Deterministic Model of Sin-
glet State Correlations Based on Relaxing Mea-
surement Independence, Phys. Rev. Lett. 105,
250404 (2010).
[39] R. Chaves, Polynomial Bell Inequalities, Phys.
Rev. Lett. 116, 010402 (2016).
[40] C. M. Lee and R. W. Spekkens, Causal Infer-
ence via Algebraic Geometry: Feasibility Tests
for Functional Causal Structures with Two Bi-
nary Observed Variables, J. Causal Inference 5,
6 (2017).
[41] C. Emary, N. Lambert, and F. Nori,
Leggett–Garg inequalities, Rep. Prog. Phys. 77,
016001 (2014).
[42] L. Clemente and J. Kofler, Necessary and suf-
ficient conditions for macroscopic realism from
quantum mechanics, Phys. Rev. A 91, 062103
(2015).
[43] M. Ringbauer, C. J. Wood, K. Modi,
A. Gilchrist, A. G. White, and A. Fedrizzi,
Characterizing Quantum Dynamics with Initial
System-Environment Correlations, Phys. Rev.
Lett. 114, 090402 (2015).
[44] H. Buhrman, R. Cleve, S. Massar, and R. De
Wolf, Nonlocality and communication complex-
ity, Rev. Mod. Phys. 82, 665 (2010).
[45] M. Pawłowski, T. Paterek, D. Kaszlikowski,
V. Scarani, A. Winter, and M. Zukowski, Infor-
mation causality as a physical principle. Nature
461, 1101 (2009).
[46] T. Christof and A. Löbel, PORTA POlyhedron
Representation Transformation Algorithm,
(2009).
[47] J. Clauser, M. Horne, A. Shimony, and R. Holt,
Proposed Experiment to Test Local Hidden-
Variable Theories, Phys. Rev. Lett. 23, 880
(1969).
[48] R. Horodecki, P. Horodecki, and M. Horodecki,
Violating Bell inequality by mixed spin-12 states:
necessary and sufficient condition, Phys. Lett. A
200, 340 (1995).
8
A Classical models for a sequence of
projective measurements on a qubit
A.1 Classical simulation with hidden states of
dimension 4
Note that the sequential measurement scenario can
be mapped to an equivalent sequence of quantum
teleportations [30]. In the first time step Alice mea-
sures the state in her possession generating some lo-
cal probability distribution p(a|x), which she uses to
prepare different states ρ
x
a
that she sends to Bob via
the usual quantum teleportation protocol. Then Bob
measures the teleported system generating a proba-
bility distribution p(b|x, a, y) and prepares states ρ
y
b
that are teleported to Charlie. Clearly, any classical
(hidden variable) protocol simulating the statistics of
the measurements performed on the teleported states
will immediately lead to a simulation of the equiva-
lent sequence of projective measurements. As shown
in Ref. [17], this can be achieved using only two bits
of classical communication between the parties and
an arbitrary amount of shared randomness, implying
that hidden states of dimension 4 and shared random-
ness are enough to obtain all quantum correlation ob-
tained by a sequence of projective measurements on
a single qubit.
For a fair comparison of classical and quantum re-
sources we have also considered whether a classical
message of dimension 2 (1 bit of communication) is
enough to simulate all projective measurements on a
qubit state. To that aim we considered a scenario
with two time steps. If Alice has the choice between
two measurements (i.e. x = 0, 1) then one classical bit
can carry all the information about the input X and,
using the argument in the main text, all the one-way
signalling correlations in this case can be simulated
using a classical message of dimension 2. We thus
considered the case were X assumes at least 3 differ-
ent values and fully characterized the set of classical
correlations for dichotomic measurements (a, b = 0, 1)
and with x = 0, 1, 2 and y = 0, 1. The polytope cor-
responding to this scenario is described by 864 in-
equalities (many being equivalent under the allowed
symmetries given by party, input and output permu-
tations). Considering qubit states and arbitrary pro-
jective qubit measurements we could not find any vi-
olation of these inequalities.
It is interesting to note that among these inequal-
ities we find the dimension-witness inequality from
Ref. [13], given by
hB
00
i + hB
01
i + hB
10
i hB
11
i hB
20
i 3, (A1)
where hB
xy
i =
P
b=0,1
(1)
b
p(b|x, y). As shown in
Ref. [13], this inequality can be violated by measure-
ments on a qubit. There, however, a slightly differ-
ent situation is considered, the so-called prepare-and-
measure scenario, where the variable X uniquely iden-
tifies the state ρ
x
being prepared and to be measured
in the second time step. In our case the states ρ
a
x
to be measured in the second time step will depend
on X, but also on the measurement outcome A, a
feature that seems to be enough to preclude any vio-
lation of the inequality above (or any other defining
the scenario). It would be interesting to derive in-
equalities for more general scenarios including more
measurement settings or time steps to see whether
any violations can be found.
A.2 Derivation of Bell-type inequalities bound-
ing classical models for a sequence of projective
measurements
In order to derive Bell-type inequalities for the tem-
poral scenario in Fig. 5, first note that all correla-
tions compatible with such a model are also compat-
ible with a model implying that
p(a, b, c|x, y, z) =
X
λ
p(λ)p(a|x, λ) (A2)
p(b|a, x, y, λ)p(c|y, z, λ).
This follows from the fact that the arrow between
the hidden variable at a given time step and the
next measurement outcome (e.g. Λ
2
B) can be
replaced by directed arrows from the measurement
choices (or measurement outcomes) of the previous
step to the next one (e.g, X B and A B)
plus a local noise variable (Λ
B
B). Note that
these noise terms are implicitly present in the model
in Fig. 5, where they have been absorbed into Λ
1
,
Λ
2
, Λ
3
. When making the above replacement, how-
ever, these local noise terms have to be introduced
explicitly. They can then be combined into a vari-
able Λ = (Λ
A
, Λ
B
, Λ
C
) that acts as a common ances-
tor and source of randomness for all the measurement
outcomes, see Fig. A1. In principle, one would further
have to impose the independence of the local noise
terms, that is p(λ
A
, λ
B
, λ
C
) = p(λ
A
)p(λ
B
)p(λ
C
).
This, however, would define a non-convex set that
is very difficult to characterize [39, 40]. Instead, we
consider a more general convex relaxation of this set
which contains the case of independent noise variables
as a special case.
As detailed in Ref. [25], the probability distribu-
tions compatible with Eq. (A2) define a convex poly-
tope that can be characterized in terms of finitely
many extremal points. Given the list of extremal
points one can resort to standard convex optimiza-
tion software [46] to find the dual description in terms
of linear (Bell-type) inequalities. For quantum cor-
relations arising from a sequence of three projective
measurements it follows that hABCi = hAihBCi. In
other words, the full (three-point) correlator does not
carry any information that is not already contained in
the bipartite and single correlators. For this reason
we focus our attention on inequalities involving the
9
A
X
B
Y Z
C
Λ
Figure A1: A DAG for the tripartite temporal correlations
scenario that contains the DAG in Fig. 4 of the main text as
a special case. As shown in Ref. [25] the correlations com-
patible with this DAG define a convex set such that Bell-type
inequalities can be derived using standard convex optimiza-
tion software [46].
expectation values hABi and hBCi, which, due to the
structure of the problem, can be defined in different
ways. In the following we use
hAB
x,y
i =
X
a,b
(1)
a+b
p(a, b|x, y) (A3)
hBC
x,y,z
i =
X
a,b,c
(1)
b+c
p(a, b, c|x, y, z). (A4)
Since there is a causal link between X and B, the
correlations between B and C can explicitly depend
on X. We then compute the Bell-type inequalities in
this subspace, one of which is inequality (10) in the
main text.
A.3 Quantum violation of inequality (10)
In order to search for a possible quantum violation
of inequality (10), we consider a single qubit in the
initial pure state
|Ψi = cos θ
0
|0i + e
0
sin θ
0
|1i. (A5)
At each time step we measure observables
ˆ
O
i
,
parametrized by θ
i
and φ
i
ˆ
O
i
= cos φ
i
sin θ
i
ˆ
X + sin φ
i
sin θ
i
ˆ
Y + cos θ
i
ˆ
Z. (A6)
The full correlator can then be expressed, as
hABCi = hAihBCi with hAi = cos θ
0
cos θ
1
+
cos (φ
0
φ
1
) sin θ
0
sin θ
1
and hBCi = cos θ
2
cos θ
3
+
cos (φ
2
φ
3
) sin θ
2
sin θ
3
. Hence, as expected, the
projective measurement at the second time step (cor-
responding to the outcome b) destroys any correla-
tions between the first and third time steps (outcomes
a and c, respectively). Similar to hBCi, the correla-
tions between the first and second time steps are given
by hABi = cos θ
1
cos θ
2
+ cos (φ
1
φ
2
) sin θ
1
sin θ
2
.
For projective measurements on a qubit state we
thus obtain hBC
x=0,y,z
i = hBC
x=1,y,z
i. Since the
first four terms of inequality (10) are nothing else than
CHSH, this part can be maximized using the stan-
dard settings: A
0
=
ˆ
Z, A
1
=
ˆ
X, B
0
= (
ˆ
X +
ˆ
Z)/
2,
B
1
= (
ˆ
X +
ˆ
Z)/
2. The last four terms are max-
imized by setting C
0
= B
1
and C
1
= B
0
, which
obtains a value of 4. In summary, quantum correla-
tions can obtain S
τ
3
= 4 + 2
2 > 6, thus violating
inequality (10). Curiously, this holds for any initial
single-qubit state.
B Experimental details
To test inequality (5) in the main text we prepare
a single-parameter family of two-qubit states of the
form
|Ψi =
1
2
p
1 + κ
2
(|HHi + |V V i)
+
p
1 κ
2
(|HV i + |V Hi)
.
(A7)
We generate this family of states by subjecting the
separable state |Di(
1 + κ|Hi+
1 κ|V i)/
2 to
a controlled-NOT gate, as shown in Fig. 2 in the main
text. The parameter 0 κ 1 turns out to be equal
to the concurrence C of the resulting bipartite state,
such that the generated amount of entanglement can
be easily controlled by the initial state of the meter
photon.
Using the fixed measurement protocol in the main
text, the states of Eq. (A7) achieve a value of S
1bit
=
4 + (1 + κ)
2 in inequality (5). It is, however, pos-
sible to optimize the measurement settings such that
ideally every pure entangled quantum state violates
inequality (5). Specifically, this amounts to mod-
ifying the protocol in the main text such that for
m = 0 Bob measures B
0
= B
1
= B
2
=
ˆ
X, while
for x = 1 he measures B
0
= (
ˆ
X + κ
ˆ
Z)/
1 + κ
2
,
B
1
= (
ˆ
X κ
ˆ
Z)/
1 + κ
2
and B
2
=
ˆ
X. Using
this protocol, states of the form of Eq. (A7) achieve
S
1bit
= 4 + 2
1 + κ
2
6.
This is reminiscent of the observation that, us-
ing optimized measurement settings, the CHSH in-
equality [47] can be violated by every pure two-
qubit entangled quantum state [48]. In fact, inequal-
ity (5) contains a CHSH inequality for the settings
A
1
, A
2
, B
0
, B
1
, which can be violated in the usual way,
while the additional communication in our protocol
can be used to maximize the remaining four terms up
to the maximal value of 4 for every state of the form
Eq. (A7). Hence, the expectation value of inequal-
ity (5) can be written as S
1bit
= 4 + S
chsh
, where
S
chsh
is the CHSH parameter corresponding to the
first four terms of inequality (5).
10