Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Experimental Proof of Nonlocal Wavefunction Collapse for a Single Particle Using

Homodyne Measurement

Maria Fuwa1 , Shuntaro Takeda1 , Marcin Zwierz2,3 , Howard M. Wiseman3 ,∗ and Akira Furusawa1†
1
Department of Applied Physics, School of Engineering, The University of Tokyo,
7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8656, Japan
2
Faculty of Physics, University of Warsaw, Pasteura 5, 02-093 Warsaw, Poland
3
Centre for Quantum Computation and Communication Technology (Australian Research Council),
Centre for Quantum Dynamics, Griffith University, Brisbane, QLD 4111, Australia
(Dated: December 30, 2014)
A single quantum particle can be described by a wavefunction that spreads over arbitrarily large
distances, but it is never detected in two (or more) places. This strange phenomenon is explained in
quantum theory by what Einstein repudiated as “spooky action at a distance”: the instantaneous
nonlocal collapse of the wavefunction to wherever the particle is detected. We demonstrate this
single-particle spooky action, for the first time with no efficiency loophole, by splitting a single
photon between two laboratories and experimentally testing if the choice of measurement in one lab
really causes a change in the local quantum state in the other lab. To this end, we use homodyne
measurements with six different measurement settings and quantitatively verify Einstein’s spooky
arXiv:1412.7790v1 [quant-ph] 25 Dec 2014

action by violating an Einstein-Podolsky-Rosen-steering inequality by 0.042±0.006. Our experiment


also verifies the entanglement of the split single photon even when one side is untrusted.

Einstein never accepted orthodox quantum mechanics bution and terminology of Schrödinger [6], who talked
because he did not believe that its nonlocal collapse of of Alice “steering” the state of Bob’s quantum system.
the wavefunction could be real. When he first made this From a quantum information perspective, EPR-steering
argument in 1927 [1], he considered just a single particle. is equivalent to the task of entanglement verification
The particle’s wavefunction was diffracted through a tiny when Bob (and his detectors) can be trusted but Alice
hole so that it “dispersed” over a large hemispherical area (or her detectors) cannot [4]. This is strictly harder than
prior to encountering a screen of that shape covered in verifying entanglement with both parties trusted [7], but
photographic film. Since the film only ever registers the strictly easier than violating a Bell inequality [8], where
particle at one point on the screen, orthodox quantum neither party is trusted [7].
mechanics must postulate a “peculiar mechanism of ac-
To demonstrate EPR-steering quantitatively it is nec-
tion at a distance, which prevents the wave . . . from pro-
essary and sufficient to violate an EPR-steering inequal-
ducing an action in two places on the screen” [1]. That is,
ity involving Alice’s and Bob’s results [5]. Such a vi-
according to the theory, the detection at one point must
olation has been shown to be necessary for one-sided
instantaneously collapse the wavefunction to nothing at
device-independent quantum key distribution as well [9].
all other points.
Because Alice is not trusted in EPR-steering, a rig-
Here, for the first time, we demonstrate Einstein’s orous experiment cannot use postselection on Alice’s
“spooky action at a distance” [2] using a single parti- side [4, 5, 10–13]. Previous experimental tests of non-
cle (a photon), as in his original conception. We sim- local quantum state collapse over macroscopic distances,
plify Einstein’s 1927 gedankenexperiment so that the sin- without postselection on Alice’s side, have involved the
gle photon is split into just two wavepackets, one sent distribution of entangled states of multiple particles [10–
to a lab supervised by Alice and the other to a distant 12, 14–26]. Experiments demonstrating Bell-nonlocality
lab supervised by Bob. But there is a key difference, (violating a Bell inequality) for a single photon have
which enables us to demonstrate the nonlocal collapse involved postselection on both sides, opening the effi-
experimentally: rather than simply detecting the pres- ciency loophole [27, 28]; these works would otherwise
ence or absence of the photon, we use homodyne mea- have demonstrated EPR-steering of a single photon as
surements. This gives Alice the power to make different well. A recent experimental test of entanglement for a
measurements, and enables Bob to test (using tomogra- single photon via an entanglement witness has no effi-
phy) whether Alice’s measurement choice affects the way ciency loophole [29] but it demonstrates a weaker form
his conditioned state collapses, without having to trust on nonlocality than EPR-steering [4, 5]. Our experiment
anything outside his own laboratory. is the first rigorous demonstration of this “spooky action
at a distance” for a single particle without opening the
The key role of measurement choice by Alice in demon-
efficiency loophole.
strating “spooky action at a distance” was introduced in
the famous Einstein-Podolsky-Rosen (EPR) paper [3] of While the nonlocal properties of a single particle have
1935. In its most general form [4], this phenomenon has spurred much theoretical debate [30–39] and many funda-
been called EPR-steering [5], to acknowledge the contri- mental experiments [27–29, 40–44], it is also recognized
2

that a single photon split between two spatially distant detection. This allows Bob to do quantum tomography
modes is a very flexible entanglement resource for quan- on his state [50, 51], and gives Alice the power to do
tum information tasks: they have been teleported [42], different types of measurement (which is necessary for
swapped [43] and purified with linear optics [44]. Spatial- an EPR-steering test) by controlling her local oscillator
mode entanglement [34] more generally has a broad po- (LO) phase θ.
tential ranging from long distance quantum communica- If Alice were simply to detect the presence or absence of
tion [45, 46], quantum computation [42–44, 47], to the a photon, then Bob’s measurement of the same observ-
simulation of quantum many-body systems in tabletop able will be anticorrelated with Alice’s, as in Ref. [41].
implementations [48]. Our work is the first one-sided But this does not prove that Alice’s measurement af-
device-independent verification of this type of entangle- fected Bob’s local state, because such perfect anticorrela-
ment. tions would also arise from a classical mixture of |0iA |1iB
and |1iA |0iB , in which Bob’s measurement simply re-
veals a pre-existing local state for him, |1iB or |0iB . To
RESULTS demonstrate nonlocal quantum state collapse, measure-
ment choice by Alice is essential [4].
Our gedankenexperiment First we explain in detail Following Alice’s homodyne measurement of the θ-
θ
our simplified version (Fig. 1) of Einstein’s original single- quadrature XA , yielding result xθA ∈ R, Bob’s local state
particle gedankenexperiment described above, formalized is collapsed to
as the task of entanglement verification with only one √ √ √
xA ΨiAB ∝ R |1iB − 1 − R e−iθ 2 xθA |0iB , (2)

θ
trusted party, as proposed in Ref. [38]. That is, assuming
only that Bob can reliably probe the quantum state at
his location, he can experimentally prove that the choice where the proportionality
√ factor is c(xθA ) =
θ 2
of measurement by the distant Alice affects his quantum exp[−(xA ) /2]/ π. Thus by changing her LO phase θ,
4

state. This is exactly the “spooky action at a distance” Alice controls the relative phase of the vacuum and one
that Einstein found objectionable [49]. photon component of Bob’s conditioned state (modulo π,
We start with a pure single photon |1i incident on a depending on the sign of the xθA she obtains). Because of
beam splitter of reflectivity R. As a result, the state this, it is convenient for Alice to coarse-grain her result
of the single photon becomes spread out between two to s(xθA ) = sign(xθA ) ∈ {+, −}. It is possible that a more
spatially separated modes A and B: sensitive EPR-steering inequality could be obtained that
√ √ makes use of a finer-grained binning of Alice’s results,
|ΨiAB = R |0iA |1iB − 1 − R |1iA |0iB . (1) but two bins are sufficient for our experiment.
Independently of Alice’s measurement, Bob performs
The transmitted mode is sent to Alice, and the reflected full quantum state tomography using homodyne detec-
one to Bob. We allow Alice and Bob to use homodyne tion on his portion of the single photon. This enables
him to reconstruct his state, for each value of Alice’s LO
Blackbox to Bob in an EPR-steering test setting θ, and coarse-grained result s. Because of the
coarse-graining, even under the idealisation of the pure
Various measurements
state as in Eq. (2), Bob’s (normalized) conditioned state
will be mixed:
- + Alice
sign ρ̂θs = R |1i h1| + (1 − R) |0i h0|
50:50 LO p
− s R(1 − R)2/π(eiθ |1i h0| + e−iθ |0i h1|). (3)
50:50 The idealized theoretical prediction for the unconditioned
LO quantum state is

Bob
X
ρ̂ = P (s|θ)ρ̂θs = R |1i h1| + (1 − R) |0i h0| , (4)
Tomography s

where P (s|θ) = 0.5 is the relative frequency for Alice to


FIG. 1: Our simplified version of Einstein’s original gedanken-
report sign s given setting θ.
experiment. A single photon is incident on a beam splitter of We once again emphasize the intrinsic lack of trust
reflectivity R and then subjected to homodyne measurements that Bob has with respect to anything that happens in
at two spatially separated locations. Alice is trying to con- Alice’s lab. Neither her honesty nor the efficiency or
vince Bob that she can steer his portion of the single photon accuracy of her measurement devices is assumed in an
to different types of local quantum states by using different EPR-steering test. On the other hand, Bob does trust his
values of the local oscillator (LO) phase θ on her side. own measurement devices. From the experimental point
3

a b c d

2 2 2 2
Quadrature

Quadrature

Quadrature

Quadrature
0 0 0 0

-2 -2 -2 -2

Bob’ s Phase (Rad) Bob’ s Phase (Rad) Bob’ s Phase (Rad) Bob’ s Phase (Rad)

0.20 0.20 0.20 0.20

0.10 0.10 0.10 0.10

0.00 0.00 0.00 0.00

3 0 -3 -3 0 3 3 0 -3 -3 0 3 3 0 -3 -3 0 3 3 0 -3 -3 0 3

Re Im Re Im Re Im Re Im

FIG. 2: Bob’s unconditioned and conditioned quantum states for R = 0.50. (a) Bob’s unconditioned data. (b) Bob’s data
conditioned on θ = 0, positive quadrature sign s (x0A > 0). (c) θ = 0, negative quadrature sign s (x0A < 0). (d) θ = −π/3,
−π/3
positive quadrature sign s (xA > 0). (Top) Bob’s marginal distributions (the variation in the number of data points as a
function of Bob’s LO phase φ is due to unevenness in Bob’s phase scanning, and does not affect the reconstructions). (Middle)
Reconstruction of Wigner functions (top view in the inset); here ~ = 1 and (q, p) label canonically conjugate quadratures.
(Bottom) Reconstruction of truncated density matrices defined on the {|0i , |1i} subspace; negative elements in darker shade.

of view this means that his photoreceivers do not have analyzing his homodyne data (taken by scanning his LO
to be efficient, and that he can post-select on finding his phase φ from 0 to 2π) for each value of Alice’s LO phase
system in a particular subspace. In particular, for our θ and result s. The reconstructed density matrices ρ̂θs in
experiment (where there are small two-photon terms), the {|0iB , |1iB } subspace and the corresponding Wigner
he can restrict his reconstructed state ρ̂θs to the qubit functions give complementary ways to to visualize how
subspace spanned by {|0iB , |1iB }. Despite Bob’s lack Bob’s local quantum state can collapse in consequence of
of trust in Alice, she can convince him that her choice Alice’s measurement.
of measurement setting, θ, steers his quantum state ρ̂θs , The results of Bob’s tomography, which take into ac-
proving that his system has no local quantum description. count inefficiency in Bob’s detection system by using
We now present data showing this effect qualitatively, the maximum likelihood method [51] during quantum
prior to our quantitative proof of EPR-steering for this state reconstruction, are presented in Fig. 2, for the case
single-photon system. R = 0.50. There is good qualitative agreement between
Bob’s Tomography Results In our experiment, the these results and the theoretical predictions for the ideal
(heralded single-photon) input state to the beam splitter case in Eqs. (3) and (4). In particular, there are four
comprises mostly a pure single-photon state |1i, but it features to note.
has some admixture of the vacuum state |0i, and a (much First, Bob’s unconditioned quantum state is a phase-
smaller) admixture of the two-photon state |2i. (for more independent statistical mixture of the vacuum and single-
details on state preparation see the METHODS). Follow- photon components [see Fig. 2(a)], as in Eq. (4). The
ing the beam splitter and Alice’s measurement, Bob re- Wigner functions are rotationally invariant, and the off-
constructs his conditioned quantum states by separately diagonal terms in ρ̂ are zero. The vacuum component
4

0.7
pu0 = 0.55 is slightly greater than the single-photon com-
ponent pu1 = 0.45 due to the less than unit efficiency 0.6
p1 = 0.857 ± 0.008 < 1 of single-photon generation (for

EPR-steering inequality
0.5
more details see the METHODS).
Second, Bob’s conditioned quantum states are not 0.4
phase-independent, but rather exhibit coherence between
0.3
the vacuum and single-photon components [see Fig. 2(b)–
(d)] as predicted by Eq. (3). The Wigner functions 0.2
are not rotationally invariant (and have a mean field:
0.1
hq + ipi =6 0), and the off-diagonal terms in ρ̂ are non- RHS
LHS
zero. Furthermore, the negative dips observed in the con- 0
0 0.2 0.4 0.6 0.8 1
ditioned Wigner functions prove the strong nonclassical Beam splitter reflectivity
character of Bob’s local quantum state [52].
Third, depending on Alice’s result s ∈ {+, −}, Bob’s FIG. 3: The left- and right-hand sides (LHS, RHS) of the
local quantum state is collapsed into complementary EPR-steering inequality given in Eq. (5) plotted as functions
states (in the sense that they sum to the unconditioned of the beam splitter reflectivity R. The solid lines show the
theoretically predicted values, which take into the account
state) [compare columns (b) and (c) in Fig. 2.] This ef-
the experimental losses, the imperfections associated with the
fect is manifested most clearly by a relative π rotation single-photon source, and the imperfections of Alice’s homo-
between the conditioned Wigner functions Ws0 (q, p) and dyne measurements such as phase fluctuations. Four sets of
the opposite signs of the off-diagonal elements of ρ̂0s , as data were taken for each R to confirm reproducibility of the
expected from Eq. (3). experimental results. To assess the tomographic errors, error
Finally, Alice can steer Bob’s possible conditioned bars were calculated using the bootstrap method, a widely
states by her choice of measurement setting θ, as pre- used technique of resampling the data to estimate confidence
intervals; for details see [54].
dicted. Comparing the results in columns (b) and (d)
of Fig. 2, it is immediately clear that the conditioned
−π/3
Wigner function W+ (q, p) is phase shifted with re-
side thus correlates Bob’s tomographic reconstruction
spect to W+0 (q, p) by an angle θ = −π/3. Moreover,
with Alice’s announced result s, but makes no assump-
we also notice the decrease in the value of the real off-
tions about how Alice generates this result. On the
diagonal elements and the emergence of the imaginary
right-hand-side, ρ̂ is Bob’s unconditioned state, while
off-diagonal elements in the conditioned density matrix z
−π/3 σ̂B = |1i h1| − |0i h0|.
ρ̂+ as compared to ρ̂0+ . Naturally, the described EPR-
For the ideal case considered above, theory predicts a
steering effect can be demonstrated for all possible values
violation of the EPR-steering inequality (5) for n ≥ 2 and
of Alice’s LO phase θ.
any value of R (apart from 0 and 1). [However, exper-
The above results suggest that Bob’s portion of the
imental imperfections associated with the single-photon
single photon cannot have a local quantum state prior
source and the inefficiency of Alice’s photoreceivers make
to Alice defining her measurement setting θ. However a
it more difficult to obtain a violation. For details of
proof that this is the case requires much more quantita-
the theoretical predictions, see Supplementary section 2]
tive analysis, which we now present.
While the inequality is most easily violated for n = ∞
The EPR-steering inequality From Eq. (2), Bob’s [for which f (∞) = 2/π ≈ 0.6366] for our experiment it
portion of the single photon is a qubit (a quantum sys- was sufficient to use n = 6 [for which f (6) = 0.6440].
tem spanned by |0iB and |1iB ). In the experiment The experimental results in Fig. 3 well match the the-
there are small terms with higher photon numbers, but, oretical predictions calculated using independently mea-
as explained above, Bob is allowed to restrict to the sured experimental parameters; see Supplementary sec-
{|0iB , |1iB } subspace. Here we consider a nonlinear tions 3 and 4. The EPR-steering inequality is violated
EPR-steering inequality for the qubit subspace. It in- for R = 0.08, 0.38, and 0.50, but not for R = 0.90; it is
volves n different measurement settings θj by Alice, and most violated at R = 0.38 by 0.042 ± 0.006.
is given by [38]
The violation of the EPR-steering inequality by seven
n standard-deviations is a clear proof that Bob’s quantum
1 XX
q
θ z ρ̂]2 .
P (s|θj ) s Tr[σ̂Bj ρ̂θsj ] ≤ f (n) 1 − Tr [σ̂B state cannot exist independently of Alice, but rather
n j=1 s is collapsed by Alice’s measurement. We were able to
(5) rigorously demonstrate this for the first time for a sin-
Here σ̂ θ ≡ e−iθ |1i h0| + eiθ |0i h1| and f (n) is a mono- gle particle without opening the efficiency loophole by
tonically decreasing positive function of the number of using the combination of multiple (n = 6) measure-
measurement settings defined in Eq. (4.15) of Ref. [38], ment settings and highly efficient phase-sensitive homo-
under the assumption that θj = π(j/n). The left-hand- dyne measurements for Alice (ηh = 0.96 ± 0.01), cou-
5

pled with a high single-photon occupation probability Single Photon Characterization The experimentally
(p1 = 0.857±0.008). Without the close-to-unity values of generated heralded single-photon input state to the beam
these parameters, the nonlocal collapse of the single pho- splitter is well modeled by an incoherent mixture of vacuum,
ton wavefunction could not have been detected, as in the single-photon and two-photon contributions given by
case of Ref. [40] (see Ref. [38] for a detailed discussion).
ρ̂in = p0 |0i h0| + p1 |1i h1| + p2 |2i h2| . (6)

The input state parameters are p0 = 0.120 ± 0.007, p1 =


DISCUSSION
0.857 ± 0.008, and p2 = 0.02 ± 0.01 as measured by setting
the beam splitter reflectivity to R = 1, that is, sending all
We have demonstrated, both rigorously and in the easy heralded single photons to Bob, and performing single mode
visualized form of nonclassical Wigner functions, the non- tomography. The higher order contributions add up to a neg-
locality of a single particle using a modern and simplified ligible value of ph = 0.004 ± 0.002.
version of Einstein’s original gedankenexperiment. That Imperfections of Alice’s Measurements The imper-
fections of Alice’s measurements include the inefficiency and
is, we demonstrated Einstein’s “spooky action at a dis- inaccuracy of homodyne detection as well as other losses in
tance” in that Bob’s quantum state (of his half of a single Alice’s apparatus. The inefficiency of Alice’s homodyne detec-
photon) was provably dependent on Alice’s choice of mea- tion of 1 − ηh = 0.04 ± 0.01 can be attributed to the imperfect
surement (on the other half), and could not have been spatial mode-match of 0.017 ± 0.004, homodyne detector cir-
pre-existing. Quantitatively, we violated a multi-setting cuit noise of 0.011±0.002, and the inefficiency of photo-diodes
nonlinear EPR-steering inequality by several standard (HAMAMASTU, S3759SPL) of 0.01. The inaccuracy of Al-
deviations (0.042 ± 0.006). ice’s homodyne detection is mostly caused by phase fluctua-
tions when locking the relative phase between the signal beam
This EPR-steering experiment is a form of entangle- to be measured and the LO: the root-mean-square phase-lock
ment verification, for a single-photon mode-entangled fluctuations are ∆θ = 3.9◦ , estimated from the signal to noise
state, which does not require Bob to trust Alice’s ratio of the error signal feedback to the piezo electric trans-
devices, or her reported outcomes. It was possible only ducer used to lock the relative phase to zero. Other losses
because we used a high-fidelity single photon state in Alice’s apparatus are the propagation inefficiency from the
and very high efficiency homodyne measurements, to OPO to the homodyne detector of 0.015±0.003 and the losses
induced by the imperfections of the anti-reflection coating in
perform the steering measurements on Alice’s side and
the beam splitter which are 1.07×10−2 ±0.01×10−2 at worst;
the tomographic state reconstruction on Bob’s. Our these add up to an additional loss of lA = 0.025 ± 0.007.
results may open a way to new protocols for one-sided
device-independent quantum key distribution [9] based
on the DLCZ protocol employing single-rail qubits [45].

Electronic address: h.wiseman@griffith.edu.au

Electronic address: akiraf@ap.t.u-tokyo.ac.jp
[1] Bacciagaluppi, G. & Valentini, A. Quantum theory at the
METHODS
crossroads: Reconsidering the 1927 Solvay Conference.
Cambridge University Press, Cambridge (2009).
Experiment Here we present the experimental details [2] Born, M. Natural philosophy of cause and chance. Oxford
of the scheme depicted in Fig. 1. The source laser is a University Press, Oxford (1949), page 109.
continuous-wave Ti:sapphire laser of 860 nm. The heralded [3] Einstein, A., Podolsky, B. & Rosen, N. Can quantum-
single photons are conditionally produced based on the setup mechanical description of physical reality really be con-
presented in Ref. [53] at an average count rate of about 8000 sidered complete? Phys. Rev. 47, 777 (1935).
s−1 using a weakly pumped non-degenerate optical paramet- [4] Wiseman, H. M., Jones, S. J. & Doherty, A. C. Steering,
ric oscillator (OPO). The single photons are impinged on a entanglement, nonlocality, and the Einstein-Podolsky-
beam splitter characterized by reflectivity R, which was set Rosen paradox. Phys. Rev. Lett. 98, 140402 (2007).
to four different values R ∈ {0.08, 0.38, 0.50, 0.90}. [5] Cavalcanti, E. G., Jones, S. J., Wiseman, H. M. &
Alice and Bob perform homodyne measuments on the Reid, M. D. Experimental criteria for steering and
transmitted and reflected signals, respectively. A piezo elec- the Einstein-Podolsky-Rosen paradox. Phys. Rev. A 80,
tric transducer is used to control the relative optical path 032112 (2009).
length of the LO of about 10 mW and the signal field, which [6] Schrödinger, E. Discussion of probability relations be-
determines the relative phase. This relative phase is locked, tween separated systems. Proc. Cambridge Philos. Soc.
using an Integral (I) feedback control scheme, to one of the 31, 553 (1935).
six values θ ∈ {0, ±π/6, ±π/3, π/2} for Alice, and is scanned [7] Terhal, B. M. Bell inequalities and the separability crite-
across the full range φ ∈ [0, 2π) for Bob. 200,000 quadra- rion. Phys. Lett. A 271, 319–326 (2000).
tures, each with an integration time of 500 ns, were recorded [8] Bell, J. S. On the Einstein Podolsky Rosen paradox.
for each phase θ, resulting in a total of 1,200,000 quadratures Physics 1, 195–200 (1964).
for each data set; the entire experiment took about 20 min- [9] Branciard, C., Cavalcanti, E. G., Walborn, S. P., Scarani,
utes for each data set. Four data sets were measured for each V. & Wiseman, H. M. One-sided device-independent
beam splitter reflectivity R. quantum key distribution: Security, feasability and the
6

connection with steering. Phys. Rev. A 85, 010301 entanglement with local homodyne measurements. Phys.
(2012). Rev. Lett. 110, 130401 (2013).
[10] Smith, D. H. et al. Conclusive quantum steering with [30] Tan, S. M., Walls, D. F. & Collet, M. J. Nonlocality of a
superconducting transition-edge sensors. Nat. Commun. single photon. Phys. Rev. Lett. 66, 252 (1991).
3:625 doi: 10.1038/ncomms1628 (2012). [31] Hardy, L. Nonlocality of a single photon revisited. Phys.
[11] Wittmann, B. et al. Loophole-free Einstein-Podolsky- Rev. Lett. 73, 2279 (1994).
Rosen experiment via quantum steering. New J. Phys. [32] Lee, H.-W. & Kim, J. Quantum teleportation and Bell’s
14, 053030 (2012). inequality using single-particle entanglement. Phys. Rev.
[12] Bennet, A. J. et al. Arbitrarily loss-tolerant Einstein- A 63, 012305 (2000).
Podolsky-Rosen steering allowing a demonstration over [33] Björk, G., Jonsson, P. & Sánchez-Soto, L. L. Single-
1 km of optical fiber with no detection loophole. Phys. particle nonlocality and entanglement with the vacuum.
Rev. X doi:10.1103/PhysRevX.2.031003 (2012). Phys. Rev. A 64, 042106 (2001).
[13] Evans, D. A., Cavalcanti, E. G., Wiseman, H. M. Loss- [34] Wiseman, H. M. & Vaccaro, J. A. Entanglement of indis-
tolerant tests of Einstein-Podolsky-Rosen steering. Phys. tinguishable particles shared between two parties. Phys.
Rev. A 88, 022106 (2013). Rev. Lett. 91, 097902 (2003).
[14] Reid, M. D. et al. Colloquium: The Einstein-Podolsky- [35] van Enk, S. J. Single-particle entanglement. Phys. Rev.
Rosen paradox: From concepts to applications. Rev. A 71, 032339 (2005).
Mod. Phys. 81, 1727 (2009). [36] Dunningham, J. & Vedral, V. Nonlocality of a single par-
[15] Carvalho, M. A. D. et al. Experimental observation of ticle. Phys. Rev. Lett. 99, 180404 (2007).
quantum correlations in modular variables. Phys. Rev. A [37] Quintino, M. T., Aráujo, M., Cavalcanti, D., Santos, M.
86, 032332 (2012). F. & Cunha, M. T., Maximal violations and efficiency
[16] Saunders, D. J., Jones, S. J., Wiseman, H. M. & Pryde, requirements for Bell tests with photodetection and ho-
G. J. Experimental EPR-steering using Bell-local states. modyne measurements. J. Phys. A 45, 215308 (2012).
Nat. Phys. 6, 845 (2010). [38] Jones, S. J. & Wiseman, H. M. Nonlocality of a single
[17] Saunders, D. J. et al. The simplest demonstrations of photon: Paths to an Einstein-Podolsky-Rosen-steering
quantum nonlocality. New J. Phys. 14, 113020 (2012). experiment. Phys. Rev. A 84, 012110 (2011).
[18] Ou, Z. Y., Pereira, S. F., Kimble, H. J. & Peng, K. C. [39] Brask, J. B., Chaves, R., & Brunner, N. Testing nonlocal-
Realization of the Einstein-Podolsky-Rosen paradox for ity of a single photon without a shared reference frame.
continuous variables. Phys. Rev. Lett. 68, 3663 (1992). Phys. Rev. A 88, 012111 (2013).
[19] Bowen, W. P., Schnabel, R., Lam, P. K. and Ralph, T. C. [40] Babichev, S. A., Appel, J. & Lvovsky, A. I. Remote
Experimental investigation of criteria for continuous vari- preparation of a single-mode photonic qubit by measur-
able entanglement. Phys. Rev. Lett. 90, 043601 (2003). ing field quadrature noise. Phys. Rev. Lett. 92, 047903
[20] Bowen, W. P., Treps, N., Schnabel, R., Ralph, T. C. (2004).
& Lam, P. K. Continuous variable polarization entangle- [41] Guerreiro, T., Sanguinetti, B., Zbinden, H., Gisin, N. &
ment, experiment and analysis. J. Opt. B 5, S467 (2003). Suarez, A. Single-photon space-like antibunching. Phys.
[21] Howell, J. C., Bennink, R. S., Bentley, S. J. & Boyd, Lett. A 376, 2174–2177 (2012).
R. W. Realization of the Einstein-Podolsky-Rosen para- [42] Lombardi, E., Sciarrino, F., Popescu, S. & De Martini,
dox using momentum- and position-entangled photons F. Teleportation of a vacuum–one-photon qubit. Phys.
from spontaneous parametric down conversion. Phys. Rev. Lett. 88, 070402 (2002).
Rev. Lett. 92, 210403 (2004). [43] Sciarrino, F., Lombardi, E., Milani, G. & De Martini,
[22] Eberle, T. et al. Strong Einstein-Podolsky-Rosen entan- F. Delayed-choice entanglement swapping with vacuum–
glement from a single squeezed light source. Phys. Rev. one-photon quantum states. Phys. Rev. A 66, 024309
A 83, 052329 (2011). (2002).
[23] Walborn, S. P., Salles, A., Gomes, R. M., Toscano, [44] Salart, D. et al. Purification of single-photon entangle-
F. & Souto Ribeiro, P. H. Revealing hidden Einstein- ment. Phys. Rev. Lett. 104, 180504 (2010).
Podolsky-Rosen nonlocality. Phys. Rev. Lett. 106, [45] Duan, L.-M., Lukin, M. D., Cirac, J. I. & Zoller, P. Long-
130402 (2011). distance quantum communication with atomic ensembles
[24] Händchen, V. et al. Observation of one-way Einstein- and linear optics. Nature 414, 413 (2001).
Podolsky-Rosen steering. Nat. Photon. 6, 596 (2012). [46] Sangouard, N., Simon, C., de Riedmatten, H. & Gisin,
[25] Steinlechner, S., Bauchrowitz, J., Eberle, T. & Schn- N. Quantum repeaters based on atomic ensembles and
abel, R. Strong Einstein-Podolsky-Rosen steering with linear optics. Rev. Mod. Phys. 83, 33 (2011).
unconditional entangled states. Phys. Rev. A 87, 022104 [47] Knill, E., Laflamme, R. & Milburn, G. J. A scheme for
(2013). effcient quantum computation with linear optics. Nature
[26] Schneeloch, J., Dixon, P. B., Howland, G. A., Broadbent, 396, 46 (2001).
C. J. & Howell, J. C. Violation of continuous-variable [48] Illuminati, F. Quantum optics: Light does matter. Nat.
Einstein-Podolsky-Rosen steering with discrete measure- Phys. 2, 803 (2006).
ments. Phys. Rev. Lett. 110, 130407 (2013). [49] Born, M. The Born-Einstein letters. Macmillan Press,
[27] Babichev, S. A., Appel, J. & Lvovsky, A. I. Homodyne London (1971).
tomography characterization and nonlocality of a dual- [50] Vogel, K. & Risken, H. Determination of quasiprobability
mode optical qubit. Phys. Rev. Lett. 92, 193601 (2004). distributions in terms of probability distributions for the
[28] Hessmo, B., Usachev, P., Heydari, H. & Björk, G. Ex- rotated quadrature phase. Phys. Rev. A 40, 2847 (1989).
perimental demonstration of single photon nonlocality. [51] Lvovsky, A. I. & Raymer, M. G. Continuous-variable op-
Phys. Rev. Lett. 92, 180401 (2004). tical quantum-state tomography. Rev. Mod. Phys. 81,
[29] Morin, O. et al. Witnessing trustworthy single-photon 299 (2009).
7

[52] Leonhardt, U. Measuring the quantum state of light. prior to the sharing of the single photon [56], it does not add
Cambridge University Press, Cambridge (2005), page 43. any additional assumptions to our experiment.
[53] Takeda, S. et al. Generation and eight-port homo- In our EPR-steering test, Alice performs homodyne mea-
dyne characterization of time-bin qubits for continuous- surements on her part of the mixed state ρ̂AB using inefficient
variable quantum information processing. Phys. Rev. A photoreceivers. The inefficiency of her photoreceivers is mod-
87, 043803 (2013). eled by introducing a photon loss channel, which is then fol-
[54] Efron B. & Tibshirani, R. J. An introduction to the boot- lowed by a fully efficient homodyne measurement. The quan-
strap. Champman & Hall/CRC, London (1994). tum state shared by Alice and Bob subjected to the photon
[55] Jones, S. J., Wiseman, H. M., Bartlett, S. D., Vaccaro, loss process on Alice’s side can be written as
J. A. & Pope, D. T. Entanglement and symmetry: A
case study in superselection rules, reference frames, and 2
X
beyond. Phys. Rev. A 74, 062313 (2006). ρ̂loss
AB = K̂n ρ̂AB K̂n† , (8)
[56] Wiseman, H.M. Defending Continuous Variable Telepor- n=0
tation: Why a laser is a clock, not a quantum channel.
J. Opt. B: Quant. Semiclass. Opt. 6, S849 (2004). where K̂n ’s are three Kraus operators that describe the pho-
This work was partly supported by the SCOPE program of ton loss channel for ρ̂AB . These operators are given by
the MIC of Japan, PDIS, GIA, G-COE, and APSA commis-
sioned by the MEXT of Japan, FIRST initiated by the CSTP √
of Japan, ASCR-JSPS, and the Australian Research Coun- K̂0 = |0iA h0| + ηA |1iA h1| + ηA |2iA h2|, (9)
p p
cil Centre of Excellence CE110001027. M.F. and S.T. ac- K̂1 = 1 − ηA |0iA h1| + 2ηA (1 − ηA )|1iA h2|, (10)
knowledges financial support from ALPS. M.Z. would like to
K̂2 = (1 − ηA )|0iA h2| (11)
acknowledge financial support from the European Union Sev-
enth Framework Programme (FP7/2007-2013) under grant
agreement n◦ [316244]. and correspond to three distinct possibilities of not losing any
photons, losing a single photon and losing two photons, re-
Author Contributions H.M.W. and A.F. supervised the spectively. Here ηA = ηh (1−lA ), where ηh = 0.96±0.01 is the
project. H.M.W and M.Z. developed the theory and defined efficiency of Alice’s photoreceivers and lA = 0.025 ± 0.007 de-
scientific goals. M.F. designed, performed the experiment, ac- notes other losses in Alice’s apparatus. The theoretical fully
quired and analyzed data with the help of S.T. M.F., H.M.W. efficient homodyne measurement, which follows the photon
and M.Z. wrote the manuscript with assistance from all other loss process is modelled by projecting ρ̂loss
AB onto the quadra-
co-authors. ture eigenstate |xθA i.
We also allow for the fact that Alice’s homodyning is not
perfectly accurate by including in our analysis the long-time
SUPPLEMENTARY MATERIAL phase-lock fluctuations, modeled by a Gaussian distribution
centered around the phase lock point θ. This gives us the fol-
lowing expression for Bob’s (normalized) local quantum state
Here we assess the influence of the experimental imper- conditioned on Alice’s local oscillator phase θ and a coarse-
fections on the violation of the EPR-steering inequality pre-
sented in the main text. We consider the experimental im-
perfections associated with the single-photon source, and the
inefficiency and inaccuracy of Alice’s homodyne detection.
State preparation Heraled single photons (modeled by
Eq. 6) are impinged on the beam splitter, generating a mixed
quantum state shared by Alice and Bob, which is given by

ρ̂AB = ] p0 |00i
√ √ √
] p1 ( R|01i − 1 − R|10i)
√ p
] p2 (R|02i − 2R(1 − R)|11i + (1 − R)|20i) ,(7)

where R denotes the beam splitter reflectivity and ] is defined


as ]α|ai ≡ +|α|2 |aiha| [55]. A typical two-mode tomographic
reconstruction of this state with R = 0.50 is presented in
Fig. 4. In the figure, the reconstructed state is truncated to
the {|0i , |1i} subspace on Bob’s side because, as stated in the
main text, we consider in this work a nonlinear EPR-steering
inequality for qubits.
Homodyne detection The local oscillators of Alice and
Bob are synchronized by sourcing them from the same laser, FIG. 4: A typical two-mode tomographic reconstruction of
as a continuous process during the experiment. This phase the absolute values of the density matrix elements associated
locking is necessary for our experiment to succeed, because with the quantum state given in Eq. (7) with R = 0.50, trun-
if Alice’s LO is not well phase-locked to Bob’s, she will not cated to the qubit subspace on Bob’s side. The blue, magenta,
be able to obtain good correlations, and therefore will not be and turquoise boxes show the density matrix elements of the
able to demonstrate EPR-steering. Note that since this phase vacuum, the single-photon, the two-photon and the higher-
locking can in principle be done using only a classical channel, order subspaces, respectively.
8

grained result s line in order to coarse-grain on the sign s of the quadrature



result xθ̃A .
(θ̃ − θ)2
Z  
00 dθ̃
ρ̂θs = exp −
2(∆θ)2
p
−∞ 2π(∆θ)2
Bob’s conditioned state Finally, we restrict Bob’s lo-
Z
× dxθ̃A hxθ̃A |ρ̂loss θ̃
AB |xA i/P (s|θ). (12) cal conditioned quantum state to the qubit subspace: ρ̂θs =
Rs 0 0
ρ̂θs /N , where ρ̂θs is the (unnormalized) restricted conditioned
◦ 00
Here ∆θ = 3.9 is the measured value for Alice’s root-mean- state obtained by discarding from ρ̂θs all terms that contain
square phase-lock fluctuations and P (s|θ) = 0.5. Note that we two-photon contributions and N = 1 − p2 R2 is the renormal-
had to integrate the coefficients of the conditioned state with ization factor. We can explicitly write Bob’s local conditioned
respect to xθ̃A over the negative (R− ) and positive (R+ ) real quantum state as

ρ̂θs = {[p1 R + 2p2 R(1 − R)] |1i h1| + [p0 + p1 (1 − R) + p2 (1 − R)2 ] |0i h0|
2
− s ηA R(1 − R)2/π[p1 + p2 (1 − R)(2 − ηA )]e−(∆θ) /2 (eiθ |1i h0| + e−iθ |0i h1|)}/(1 − p2 R2 ).
p

(13)

Expected violation of the EPR-steering inequality For the right-hand side of the EPR-steering inequality we also
Given this state we can evaluate the expected value for the restrict Bob’s unconditioned state to the qubit subspace:
left-hand side of the EPR-steering inequality (5) defined in
the main text: ρ̂ = {[p1 R + 2p2 R(1 − R)] |1i h1|
n + [p0 + p1 (1 − R) + p2 (1 − R)2 ] |0i h0|}/(1 − p2 R2 ).
1 XX θ θ
P (s|θj ) s Tr[σ̂Bj ρsj ] (15)
n j=1 s
2 p1 + p2 (1 − R)(2 − ηA ) The right-hand side of the EPR-steering inequality (5) thus
= 2e−(∆θ) /2
p
ηA R(1 − R)2/π . evaluates to
1 − p2 R 2
q
(14) z
f (n) 1 − Tr [σ̂B ρ̂]2
s
[p0 + p1 (1 − 2R) + p2 (1 − R)(1 − 3R)]2
= f (n) 1 − .
0.05 (1 − p2 R2 )2
(16)
0.04
Expected violation

We plot the expected violation of the EPR-steering inequal-


0.03 ity in Fig. 5. This manifests an asymmetry in the inequality
non-violation
region violation in terms of the beam splitter reflectivity R. If there
0.02 were no imperfections in the state preparation then the degree
of violation would actually be independent of R. The chief
0.01 imperfection causing this asymmetry is the less than unit ef-
ficiency p1 = 0.857 ± 0.008 < 1 of single-photon generation,
0 as the value of the left-hand side decreases linearly with p1 ,
0 0.2 0.4 0.6 0.8 1 at least for small p2 , while the right-hand side takes its max-
Beam splitter reflectivity
imum value when p1 R = 0.5. Consequently, there exists a
maximum reflectivity Rmax ≈ 0.68 for which the inequality is
FIG. 5: The expected violation of the EPR-steering inequality violated when 0 < R ≤ Rmax .
plotted as a function of the beam splitter reflectivity R. The There is no minimum reflectivity for which the inequal-
solid line shows the theoretically predicted violation, which ity is violated, however, there exists an optimum reflectiv-
takes into the account the experimental imperfections. Four ity Ropt < 0.50 (in our experiment this optimal value is
sets of data were taken for each R to confirm reproducibility Ropt ≈ 0.28) for which the degree of violation is greatest.
of the experimental results. To assess the tomographic er- This can be understood as follows. Alice is the party that
rors, error bars were calculated using the bootstrap method. is supposed to prove that she can steer Bob’s local quantum
The discrepancies in the data sets can be attributed to the state. Therefore, it is the efficiency and accuracy of Alice’s
inconsistencies in the setup preparation, most significantly, homodyne measurements that matters the most for a suc-
the varying single-photon generation probability p1 . The ob- cessful demonstration of EPR steering. As a result, in an
served violation is slightly smaller than the theoretically pre- EPR-steering experiment we have to compensate the experi-
dicted value due to the imperfections in the phase estimation mental imperfections on Alice’s side by increasing the portion
procedure during Bob’s tomography. of the single photon that is send to her.

You might also like