This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

An Einstein-Podolsky-Rosen argument based on weak forms of local realism not falsifiable by GHZ or Bell experiments

Jesse Fulton, Run Yan Teh and M. D. Reid1 1 Centre for Quantum Science and Technology Theory, Swinburne University of Technology, Melbourne 3122, Australia
Abstract

The Einstein-Podolsky-Rosen (EPR) paradox gives an argument for the incompleteness of quantum mechanics based on the premises of local realism. A general view is that the argument is compromised, because EPR’s premises are falsified by Greenberger-Horne-Zeilinger (GHZ) and Bell experiments. In this paper, we present an EPR argument based on premises not falsifiable by these experiments. First, we propose macroscopic EPR and GHZ experiments using spins S^θ\hat{S}_{\theta} defined by two macroscopically distinct states. The analyzers that realize the unitary operations UθU_{\theta} determining the measurement settings θ\theta are nonlinear devices known to create macroscopic superposition states. We note two definitions of macroscopic realism (MR). For a system with two macroscopically distinct states available to it, MR posits a predetermined outcome for a measurement S^θ\hat{S}_{\theta} distinguishing between the states. Deterministic macroscopic realism (dMR) posits MR for the system defined prior to the interaction UθU_{\theta} being carried out. Weak macroscopic realism (wMR) posits MR for the system after UθU_{\theta}, at the time tft_{f} - when the system is prepared with respect to the measurement basis, ready for a final “pointer” measurement and readout. For this system, wMR posits that the outcome of S^θ\hat{S}_{\theta} is determined, and not changed by interactions that might occur at a remote system BB. The premise wMR also posits that if the outcome for S^θA\hat{S}_{\theta}^{A} of a system AA can be predicted with certainty by a “pointer” measurement on the system BB defined at time tfBt_{f}^{B} after the unitary interaction that fixes the setting at BB, then the outcome for S^θA\hat{S}_{\theta}^{A} is determined for AA at this time tfBt_{f}^{B}, regardless of whether the unitary interaction required to fix the setting as θ\theta has taken place at AA. We show that the GHZ predictions negate dMR but are consistent with wMR. Yet, an EPR paradox arises based on wMR. As considered by Schrödinger, it is possible to measure two complementary spins of system AA simultaneously, “one by direct, the other by indirect measurement”: If we assume wMR, then at time tf=tfBt_{f}=t_{f}^{B}, the outcomes of the two spins are both determined. We revisit the original EPR paradox and find a similar result: An EPR argument can be based on a weak contextual form of local realism (wLR) not falsifiable by GHZ or Bell tests.

I Introduction

In their argument of 1935, Einstein-Podolsky-Rosen (EPR) introduced premises, based on local realism, which if valid suggested quantum mechanics to be an incomplete description of physical reality epr . The argument considered two separated particles with correlated positions and anticorrelated momenta. The correlations imply that the outcome of a measurement of either position or momentum could be inferred with certainty for one particle, by an experimentalist making the appropriate measurement on the second particle. Assuming there is no disturbance to the first particle by the experimentalist’s actions, EPR argued from their premises that the position and momentum of the first particle are both simultaneously precisely determined prior to measurement, thereby creating an inconsistency with any quantum-state description for the localised particle.

Bell later proved that all local realistic theories could be falsified by quantum predictions bell-1969 ; bell-cs-review ; bell-1971 ; chsh ; bell-brunner-rmp . Moreover, Greenberger-Horne-Zeilinger (GHZ) ghz-1 ; mermin-ghz ; ghz-amjp ; clifton-ghz gave a falsification of EPR’s premises in an “all or nothing” situation. Bell and GHZ predictions have been experimentally verified ghz-pan ; ghz-exp-Bou . Consequently, the EPR paradox is most often regarded as an illustration of the incompatibility between local realism and quantum mechanics, rather than as a valid argument for the incompleteness of quantum mechanics epr-rmp ; mermin-ghz .

In this paper, we present a different perspective on the EPR paradox. We first propose a test of local realism versus quantum mechanics in a macroscopic GHZ set-up, where realism refers to a system being in one or other of two macroscopically distinct states. This provides a way to falsify local realism at a macroscopic level. Given such a falsification may raise questions for the interpretation of quantum measurement legggarg-1 ; s-cat-1935 , we then examine carefully the definitions of macroscopic realism, showing that a less restrictive definition of macroscopic realism is not falsified by GHZ or Bell experiments. This leads to a second conclusion: A modified EPR argument that quantum mechanics is incomplete can be given, based on an alternative and (arguably) nonfalsifiable premise.

Specifically, we show how the GHZ and EPR paradoxes can be realised for macroscopic qubits, where all relevant measurements S^θ\hat{S}_{\theta} distinguish between two macroscopically distinct states. The EPR paradox is presented as Bohm’s version Bohm ; bohm-aharonov ; aspect-Bohm which examines two spatially-separated entangled spin-1/21/2 systems. The macroscopic version is a direct mapping of the original paradox, where a spin ||\uparrow\rangle and ||\downarrow\rangle is realised as two macroscopically distinct states, such as coherent states |α|\alpha\rangle and |α|-\alpha\rangle, or collective multimode spin states k=1N|k\prod_{k=1}^{N}|\uparrow\rangle_{k} and k=1N|k\prod_{k=1}^{N}|\downarrow\rangle_{k}. The necessary unitary transformations UθiU_{\theta_{i}} which determine the measurements settings θi\theta_{i} for each system ii are realised by nonlinear interactions, or CNOT gates.

Leggett and Garg gave a definition of macroscopic realism (MR) for a system “with two or more macroscopically distinct states available to it”: MR asserts that the system “will at all times be in one or other of those states” legggarg-1 . Following previous work, we note different definitions are possible manushan-bell-cat-lg ; delayed-choice-cats . Deterministic macroscopic (local) realism (dMR) asserts there is a predetermined value λ\lambda for the outcome of a measurement SS that will distinguish between the macroscopically distinct states. Locality is implied, since it is assumed that this value is not affected by spacelike-separated interactions or events.

However, the EPR-Bohm, Bell and GHZ experiments require choices of measurement settings θi\theta_{i} at each site ii, the choice establishing which spin component SθiS_{\theta_{i}} will be measured. This leads to different definitions of MR. The measurement basis is determined by the setting of a physical device (analogous to a Stern-Gerlach analyzer) which realizes a unitary operation Uθi=eiHθit/U_{\theta_{i}}=e^{-iH_{\theta_{i}}t/\hbar} where HθiH_{\theta_{i}} is the interaction Hamiltonian. After the interaction UθiU_{\theta_{i}} at a site ii, there is a final stage of measurement that includes an irreversible coupling to an environment to give a readout of the spin SθiS_{\theta_{i}}. We refer to this final stage as the “pointer measurement”.

In the macroscopic experiments we propose, the system after the selected interactions UθiU_{\theta_{i}} is in a superposition of macroscopically distinct states which have definite values for the outcomes SθiS_{\theta_{i}}. Weak macroscopic realism (wMR) posits that each system ii prepared at a time tfit_{f_{i}} after the interaction UθiU_{\theta_{i}} can be ascribed a predetermined value λθi\lambda_{\theta_{i}} for the final pointer part of the measurement S^θi\hat{S}_{\theta_{i}}, which will distinguish between the macroscopic states.

The premise wMR posits not only a weak form of realism, but also a weak form of locality. Consider the local system ii prepared at the time tfit_{f_{i}} for the pointer measurement: There is no disturbance to the value λθi\lambda_{\theta_{i}} for the pointer measurement from interactions that might occur at a spacelike-separated site; nor from the pointer measurement (if it happened to be carried out) to a spacelike-separated system. In such a model, we will show that the nonlocal effects contributing to the GHZ and Bell contradictions with local realism emerge when there are further unitary interactions occurring at both sites.

The premise of deterministic local macroscopic realism (dMR) is similar to EPR and Bell’s form of local realism, because the premise applies to the system as it exists prior to the unitary interactions UθiU_{\theta_{i}}. The macroscopic GHZ set-up hence enables an “all or nothing” falsification of dMR, which supports previous work revealing dMR to be falsifiable by macroscopic Bell tests macro-bell-lg ; manushan-bell-cat-lg ; delayed-choice-cats ; macro-bell-jeong ; wigner-friend-macro ; bell-contextual .

The main result of this paper is that weak macroscopic realism (wMR) is not falsified by the GHZ or Bell set-ups. Yet, an EPR-type argument can be put forward based on wMR. The argument applies to the EPR setup considered by Schrödinger, where two noncommuting observables are measured simultaneously “one by a direct, the other by indirect measurementsch-epr-exp-atom ; s-cat-1935 . In our paper, the EPR-Bohm gedanken experiment is examined at the time tft_{f} after the unitary interactions UθiU_{\theta_{i}} have been carried out at each site, in order to measure spin SzS_{z} of one system and SyS_{y} of the other. The premise wMR posits a predetermined value for the pointer measurement of SzS_{z} of the first system, since it can be shown to have two “macroscopically distinct states available to it” (with respect to the measurement basis SzS_{z}). But also for the EPR-Bohm state, the outcome for SyS_{y} can be inferred for the first system by a pointer measurement on the other. Hence, wMR posits simultaneous precise values for both SzS_{z} and SyS_{y}. This is not consistent with a local quantum state, the Pauli spin variances being constrained by (Δσz)2+(Δσy)21(\Delta\sigma_{z})^{2}+(\Delta\sigma_{y})^{2}\geq 1 hofmann-take , which leads to the paradox. It is important to note that there is no actual violation of the uncertainty principle, because the quantum state defined at the time tft_{f} is different to the quantum state defined after the further interaction UθiAU_{\theta^{\prime}_{i}}^{A} necessary to change the measurement setting from zz to yy at AA.

The results motivate us to revisit the original microscopic EPR-Bohm paradox, and to demonstrate an EPR argument based on a weak contextual form of local realism (wLR), which we show is not falsified by GHZ or Bell set-ups. The definitions of wMR and wLR are both contextual, being defined for the system with a specified measurement basis. The weaker assumptions are motivated by Bohr’s criticism of EPR’s 1935 paper bell-cs-review ; bohr-epr . Clauser and Shimony state that “[Bohr’s] argument is that when the phrase ‘without in any way disturbing the system’ is properly understood, it is incorrect to say that system 2 is not disturbed by the experimentalist’s option to measure aa rather than aa^{\prime} on system 1.” This suggests that after the experimentalist’s option to measure say the spin component aa, there is reason to justify no disturbance, the nonlocality stemming from the unitary interactions giving the options. We explain how wLR may be implied by wMR, by considering the system at the time of the reversible coupling to a macroscopic meter.

The layout of the paper is as follows. In Section II, we outline the definitions of local realism and macroscopic realism used in this paper. In Section III, we review the original EPR-Bohm and GHZ paradoxes, giving in Section IV the modified EPR-Bohm paradox based on the premise of wLR. In Sections V and VII, we present the proposals for macroscopic EPR-Bohm and GHZ paradoxes using cat states. Conclusions drawn from these paradoxes are given in Sections VI and VIII. The macroscopic GHZ paradox falsifies dMR, but shows consistency with wMR. Similarly, the GHZ paradox is consistent with wLR. In Section IX, we demonstrate consistency of wLR (and wMR) with violations of Bell inequalities. Further tests of wMR and wLR are devised in Section X.

II Definitions

We formalize the definitions of local realism and macroscopic realism relevant to this paper. Several different definitions are introduced. The difference between the definitions is clarified, once we recognise that there are two stages to a spin measurement SθS_{\theta}: First, there is the reversible measurement-setting stage involving a unitary interaction UθU_{\theta} which determines the measurement setting θ\theta. Second, there is the stage that comes after, which includes a final irreversible readout of a meter. We refer to the later stage as the pointer [stage of] measurement.

Consider two separated spin-1/21/2 systems AA and BB prepared at time t0t_{0} in the state |ψ|\psi\rangle. Local unitary interactions UθAU_{\theta}^{A} and UϕBU_{\phi}^{B} prepare the systems for spin measurements SθAS_{\theta}^{A} and SϕBS_{\phi}^{B}. The UθU_{\theta} are realised as reversible interactions of the system with a real device, such as a Stern-Gerlach analyzer or polarizing beam splitter, and are represented by a Hamiltonian HθH_{\theta}, where Uθ=eiHθt/U_{\theta}=e^{-iH_{\theta}t/\hbar}. The UθU_{\theta} takes place over a time interval tt, and the states prior and after the interaction UθU_{\theta} may therefore be regarded as different, in that they define the system at a different time. The state after the interaction at time tft_{f} is

|ψ(tf)=eiHθtf/|ψ|\psi(t_{f})\rangle=e^{-iH_{\theta}t_{f}/\hbar}|\psi\rangle (1)

Specific examples are given in Sections IV and V, where we note that the “system” may include another (local) set of modes, or a meter that is originally decoupled to the spin system, in which case |ψ|\psi\rangle is suitably defined. After the interaction UU, the final irreversible stage of the measurement is made, which indicates the measurement outcome. This stage often involves a direct detection of a particle at a given location, as well as amplification and a coupling to a meter. The local system prepared after the interaction UU that fixes the measurement setting, but before the irreversible stage of the measurement, is considered to be prepared for the pointer measurement. The system is said to be prepared in the preferred basis, also referred to as the measurement, or pointer, basis.

A common realisation is the spin 1/21/2 system ||1,0|\uparrow\rangle\equiv|1,0\rangle and ||0,1|\downarrow\rangle\equiv|0,1\rangle defined for two orthogonally polarized modes a±a_{\pm}. Here, |n1,n2|n1+|n2|n_{1},n_{2}\rangle\equiv|n_{1}\rangle_{+}|n_{2}\rangle_{-} where |n±|n\rangle_{\pm} is a number state for the mode a±a_{\pm}. A transformation UθU_{\theta} can then be achieved with a polarizing beam splitter, with mode transformations (a^±\hat{a}_{\pm} are boson operators defining the modes)

c^+\displaystyle\hat{c}_{+} =\displaystyle= a^+cosθa^sinθ\displaystyle\hat{a}_{+}\cos\theta-\hat{a}_{-}\sin\theta
c^\displaystyle\hat{c}_{-} =\displaystyle= a^+sinθ+a^cosθ.\displaystyle\hat{a}_{+}\sin\theta+\hat{a}_{-}\cos\theta. (2)

The c^±\hat{c}_{\pm} are boson operators for the outgoing modes emerging from the beam splitter. The interaction is described by the Hamiltonian Hθ=ik(a^+a^a^+a^)H_{\theta}=i\hbar k(\hat{a}_{+}\hat{a}_{-}^{\dagger}-\hat{a}_{+}^{\dagger}\hat{a}_{-}) where θ=kt\theta=kt. The choice θ=0\theta=0 (θ=π/4)\theta=\pi/4) corresponds to a measurement of σz\sigma_{z} (σx\sigma_{x}). A single photon impinges on the beam splitter and is finally detected at one or other locations associated with the outgoing modes aspect-Bohm . The final detection and readout of the locations constitutes the pointer measurement.

More recently, an EPR experiment has been realised for pseudo-spin measurements in a BEC setting sch-epr-exp-atom . Here, the measurement setting is determined by an interaction of a field with a two-level atom, which forms the spin 1/21/2 system.

Refer to caption
Figure 1: The set-up for an EPR-Bohm paradox. Two separated systems AA and BB are prepared in an entangled state |ψ|\psi\rangle (Eq. (4)). A switch (red dashed arrow) gives the choice to measure either SzS_{z} or SyS_{y} for each of AA and BB, by passing through an appropriately orientated analyzer symbolised by UU at each site. Here, SzS_{z} and SyS_{y} correspond to Pauli spins σz\sigma_{z} and σy\sigma_{y}. EPR’s local realism asserts that because one can predict the result for SzS_{z} (or SyS_{y}) at AA by measuring SzS_{z} (or SyS_{y}) at BB, the outcomes for SzS_{z} and SyS_{y} at AA are both predetermined, at the time t0t_{0} (prior to the choices of measurement setting). The premises of macroscopic local realism and deterministic macroscopic realism apply when the outcomes ++ and - for both spins SzS_{z} and SyS_{y} are associated with macroscopically distinct states for the system at the time t0t_{0}.

II.1 Strong elements of reality: EPR’s Local realism and Macroscopic realism

II.1.1 EPR’s Local realism

The premises presented in the original 1935 argument given by EPR are based on the assumptions of local realism. The premises, which we refer to as EPR’s local realism (LR), are summarized as two Assertions for space-like separated systems, AA and BB.

EPR’s Assertion LR (1): Realism

EPR’s reality criterion is: “If, without in any way disturbing a system, we can predict with certainty the value of a physical quantity, then there exists an element of physical reality corresponding to this physical quantity epr .”

This is interpreted as follows: “The "element of physical reality" is that predictable value, and it ought to exist whether or not we actually carry out the procedure necessary for its prediction, since that procedure in no way disturbs it mermin-ghz .” Hence, EPR Assertion LR (1) reads: If one can predict with certainty the outcome of a measurement SS on system AA without disturbing that system, then the outcome of the measurement is predetermined. The system AA can be ascribed a variable λA\lambda^{A}, the value of which gives the outcome for SS mermin-ghz .

EPR Assertion LR (2): No disturbance (Locality)

There is no disturbance to system AA from a spacelike-separated interaction or event (e.g. a measurement on system BB).

The consequences of the two Assertions as applied to the EPR-Bohm set-up leads to the EPR-Bohm paradox (Section III). Consider the system of Figure 1: If the outcome of the measurement SθAS_{\theta}^{A} at AA can be predicted with certainty by a measurement at BB, then the EPR Assertions imply the system AA at time t0t_{0} can be ascribed a variable λθA\lambda_{\theta}^{A}, the value of which gives the outcome of the measurement SθAS_{\theta}^{A}. The assignment of the variable λθA\lambda_{\theta}^{A} can be made to the system AA regardless of the measurement device actually being prepared, either at AA or at BB. In the Figure 1, this allows the assignment of both variables λzA\lambda_{z}^{A} and λyA\lambda_{y}^{A} to system AA, at the time t0t_{0}, prior to the unitary interactions UθAU_{\theta}^{A} and UϕBU_{\phi}^{B} that determine the measurement settings.

II.1.2 Macroscopic local realism and deterministic macroscopic realism

The assertions defining macroscopic local realism (MLR) are identical to those of EPR’s local realism, except that the assertions are weaker, being restricted to apply only to the subset of systems where the outcomes for all relevant measurements, SθAS_{\theta}^{A} and SϕBS_{\phi}^{B}, correspond to macroscopically distinct states of the system. This means that the systems upon which those measurements are made can be viewed as having two (or more) macroscopically distinct states available to them, so that the Leggett-Garg definition of macroscopic realism legggarg-1 can be applied, to separately posit deterministic macroscopic realism (dMR), as below. It would be argued that the assumption of realism is more robustly justified for macroscopically distinct states s-cat-1935 .

Assertion MLR (1a): EPR’s realism

This reads as for EPR Assertion LR (1). “If, without in any way disturbing a system, we can predict with certainty the value of a physical quantity, then there exists an element of physical reality corresponding to this physical quantity.”

Assertion dMR (1b): Leggett-Garg’s criterion for Macroscopic realism - Deterministic macroscopic realism (dMR)

A system which has two or more macroscopically distinct states available to it can be ascribed a predetermined value λ\lambda for the measurement SS that will distinguish between these states. The predetermined value λ\lambda is not affected by spacelike-separated measurements (e.g. further unitary transformations UϕU_{\phi}) that may occur at site BB. Hence, locality is implied, which also follows from Assertion MLR(2) below.

Assertion MLR (2): No disturbance (Locality)

There is no macroscopic disturbance to system AA from a spacelike-separated interaction or event.

Similar to EPR’s local realism, deterministic macroscopic realism asserts there is a predetermined value λθ\lambda_{\theta} for the outcome of the measurement at the time t0t_{0}, for the system as it exists prior to, or irrespective of, the interaction UθU_{\theta} (Figure 1).

II.2 Weak elements of reality: weak macroscopic realism and weak local realism

II.2.1 Weak macroscopic realism

Weak macroscopic realism (wMR) involves weaker (i.e. less restrictive) assumptions than macroscopic local realism. Macroscopic local realism implies wMR, but the converse is not true. The assertions for wMR are modified so that EPR’s local realism applies to the systems after the selection of the measurement settings, at time tft_{f} in the Figure 2.

Refer to caption
Figure 2: The assumption of weak local realism (wLR) gives rise to an EPR paradox: The set-up is for an EPR-Bohm paradox as in Figure 1. At time tft_{f}, the measurement settings are set to SzAS_{z}^{A} and SyBS_{y}^{B}, as indicated by the positions of the red dashed arrows. Weak local realism asserts that for the system AA at time tft_{f}, after the unitary rotation UzAU_{z}^{A}, the outcome for the final pointer measurement SzS_{z} at AA is determined, given by the variable λzA\lambda_{z}^{A}. Weak local realism also asserts that because one can predict the result for SyS_{y} at AA by making a final pointer measurement SyS_{y} at BB, the outcome for SyS_{y} at AA is also determined at time tft_{f} (after the rotation UyBU_{y}^{B} at BB). This is despite that a further unitary evolution UAU^{A} would be required at AA to perform the measurement SyAS_{y}^{A}. This leads to the paradox. The premise of weak macroscopic realism applies similarly, assuming the outcomes ++ and - for the spins can be viewed as corresponding to macroscopically distinct states for the system at the time tft_{f}.
Assertion wMR (1a): EPR’s criterion for realism

This assertion reads as for Assertion MLR (1a). “If, without in any way disturbing a system, we can predict with certainty the value of a physical quantity, then there exists an element of physical reality corresponding to this physical quantity epr .”

The conclusions from Assertion wMR (1a) will now be different from those of EPR, due to the modification of Assertion (2a) below. Suppose system BB at time tft_{f}, after the unitary interaction UϕBU_{\phi}^{B}, is prepared for the pointer stage of measurement of spin SϕBS_{\phi}^{B}. Assertion wMR 2(a) asserts that there is no disturbance to system AA due to whether or not the pointer stage of measurement at BB actually takes place. The modification means that the prediction for the outcome of measurement SS at AA, as based on a measurement at BB, ensures a predetermination of the result at AA at the time tft_{f}, provided the unitary interaction UϕBU_{\phi}^{B} that fixes the measurement setting ϕ\phi at BB has taken place.

The Assertion wMR (1a) can hence be rephrased: Consider the system of Figure 2: If the outcome of a measurement SθAS_{\theta}^{A} at AA can be predicted with certainty by a pointer measurement on the system at time tft_{f} at BB, then there exists an element of physical reality corresponding to the outcome of SθAS_{\theta}^{A} at AA. Thus, the system AA can be ascribed a hidden variable λθA\lambda_{\theta}^{A} that determines the outcome for SθAS_{\theta}^{A}. This is true regardless of whether the pointer measurement at BB is actually carried out (because that would not disturb the system AA), and regardless of whether the interaction UAU^{A} at AA that fixes the measurement setting θ\theta has actually been carried out at AA (and regardless of future unitary interactions at AA). However, the predetermination is based on the system BB being prepared for a pointer measurement, and therefore only applies at the time tft_{f}, when no further unitary interactions that would cause a change of measurement setting at BB have taken place.

Assertion wMR (1b): Macroscopic realism for the system prepared for the pointer measurement

Suppose a system AA that is prepared for the pointer measurement of SS can be considered to have two or more macroscopically distinct states available to it, where each of those states has a definite outcome for the pointer measurement. Weak macroscopic realism asserts that such a system can be ascribed a predetermined value λA\lambda^{A} for the pointer measurement SS that will distinguish between these states.

The premise applies, so that the result of the pointer measurement for SθAS_{\theta}^{A} is predetermined at the time tft_{f}, once the interaction UθAU_{\theta}^{A} determining the measurement setting at AA has taken place (Figure 2).

Assertion wMR (2a): Pointer locality: No disturbance from a pointer measurement

The pointer stage of a measurement gives no disturbance to a spacelike-separated system i.e. there is no disturbance to system AA from a pointer measurement on system BB.

Assertion wMR (2b): Pointer locality: No disturbance to a pointer measurement

There is no disturbance to the predetermined value λA\lambda^{A} for the pointer measurement at AA (as described in Assertion wMR (1b)) by spacelike-separated interactions or events (e.g. further unitary transformations UϕBU_{\phi}^{B}) that may occur at BB.

The assertions when applied to EPR-Bohm set-up of Figure 2 will lead to an EPR-type paradox, as explained in Sections IV and V.

II.3 Weak local realism

The premise of weak local realism (wLR) is a weak form of local realism, which applies to the system prepared at time tft_{f} for the pointer measurement (Figure 2). This contrasts with definitions of local realism that apply to the system at time t0t_{0}, prior to the entire measurement process, and which can be falsified. It should be mentioned that weak local realism, as with weak macroscopic realism, does not exclude “nonlocality”, since, as we will see, these premises are consistent with the quantum predictions for Bell and GHZ experiments.

The assertions of wLR are as for weak macroscopic realism (wMR), except there is no longer the restriction that the outcomes correspond to macroscopically distinct states of the system being measured. A connection between wLR and wMR is given in the Section II.D. As with wMR, in this model, the pointer measurement constitutes a passive stage of the measurement.

We comment on the terminology local realism. The most general definition of local realism is given by Bell’s local hidden variable theories, also referred to as local realistic theories bell-cs-review . These theories allow for local interactions with local measurement devices, so that the value for the outcome of the measurement need not be predetermined (at time t0t_{0} or tft_{f}). This contrasts with the stricter definition, which we refer to as a non-contextual or deterministic local realism bell-cs-review , meaning the values for measurement outcomes are predetermined at the time t0t_{0}, prior to the entire measurement process including the unitary interactions UU. However, general local realistic hidden variable theories imply EPR’s local realism. Moreover, for the EPR-Bohm system where the correlations between certain spin measurements are maximum, the local realistic theories also imply the stricter deterministic local realism, for spin measurements bell-1969 .

Generally speaking, however, we cannot assume wLR to be a “weaker” assumption than local realism, in the sense that it not necessarily a subset of those assumptions, since local realistic theories may allow non-passive interactions with the pointer measurement apparatus. To avoid confusion, we emphasize that weak local realism refers to a weaker version of the non-contextual deterministic form of local realism. In later Sections, we show that while all local realistic theories are ruled out by GHZ and Bell experiments, wLR theories are not, which motivates the terminology “weak”.

Assertion wLR (1a): EPR’s realism

The assertion is as for EPR Assertion LR (1). As with wMR, the conclusions drawn from this Assertion are impacted by Assertion (2a). Weak local realism asserts: If the outcome of the measurement SθAS_{\theta}^{A} at AA can be predicted with certainty at time tft_{f} by the final pointer stage of measurement at BB (the local dynamics UU of the measurement setting being already performed at BB), then the system at AA at time tft_{f} can be ascribed a variable λθA\lambda_{\theta}^{A}, the value of which gives the outcome of the measurement SθAS_{\theta}^{A}. This is true regardless of whether the pointer measurement at BB is actually performed, and regardless of whether the unitary interaction UθAU_{\theta}^{A} has taken place at AA, and regardless of any further unitary interactions at AA (Figure 2).

Assertion wLR (1b): Realism for the system prepared for the pointer measurement

The result of the pointer measurement for SAS^{A} for system AA is predetermined (by a variable λA\lambda^{A}) once the local interaction UθAU_{\theta}^{A} for the measurement setting at AA has taken place.

Consider the system AA at time tft_{f}, after the unitary rotation UθAU_{\theta}^{A}. The system AA is prepared for the pointer stage of measurement of SθAS_{\theta}^{A}, without the need for a further unitary rotation UU. The premise wLR asserts that the system AA at time tft_{f} can be ascribed by a hidden variable λθA\lambda_{\theta}^{A} with value +1+1 or 1-1, that value determining the outcome of the pointer measurement for SθAS_{\theta}^{A} at AA, if that pointer stage of measurement were to be carried out on the prepared state. In accordance with Assertion (2b), the predetermined value λθA\lambda_{\theta}^{A} is not affected by spacelike-separated events (e.g. further unitary transformations UϕU_{\phi}) that may occur at BB.

Assertion wLR (2a): Pointer locality: No disturbance from a pointer measurement

The assertion reads as for Assertion wMR (2a).

Assertion wLR (2b): Pointer locality: No disturbance to a pointer measurement

The assertion reads as for Assertion wMR (2b).

The Assertion wLR (1b) which asserts realism is less convincing for a microscopic system than for a macroscopic system, and might seem to contradict the results of Bell’s theorem. We later show that wLR is not contradicted by the Bell or GHZ predictions. The assertions when applied to the set-up of Figure 2 will nonetheless lead to an EPR-type paradox, as explained in Section IV.

II.4 Link between wLR and wMR

At first glance, weak local realism (wLR) is seen to be a stronger assumption than weak macroscopic realism (wMR), meaning it is a more restrictive (less convincing) assumption. However, if the time tft_{f} is carefully specified, we show that wLR can be justified by wMR.

Consider the system at time tft_{f} after the unitary rotations UθAU_{\theta}^{A} and UϕBU_{\phi}^{B} that determine the measurement settings, say θ\theta and ϕ\phi, respectively at AA and BB. At this stage, or later, in the measurement process, there is a coupling of each local system to a macroscopic meter, via an interaction HMH_{M}. The final state after coupling is of the form

|ψM\displaystyle|\psi_{M}\rangle =\displaystyle= c1|p+A|θ|pB|ϕ+c2|pA|θ|p+B|ϕ\displaystyle c_{1}|p_{+}\rangle_{A}|\uparrow\rangle_{\theta}|p_{-}\rangle_{B}|\downarrow\rangle_{\phi}+c_{2}|p_{-}\rangle_{A}|\downarrow\rangle_{\theta}|p_{+}\rangle_{B}|\uparrow\rangle_{\phi}
+c3|p+A|θ|p+B|ϕ+c4|pA|θ|pB|ϕ\displaystyle+c_{3}|p_{+}\rangle_{A}|\uparrow\rangle_{\theta}|p_{+}\rangle_{B}|\uparrow\rangle_{\phi}+c_{4}|p_{-}\rangle_{A}|\downarrow\rangle_{\theta}|p_{-}\rangle_{B}|\downarrow\rangle_{\phi}

where cic_{i} are probability amplitudes, and |p+A/B|p_{+}\rangle_{A/B} and |pA/B|p_{-}\rangle_{A/B} are macroscopic states for the pointer of the meter, indicating Pauli spin outcomes of +1+1 and 1-1 respectively, at AA and BB. We see that |ψM|\psi_{M}\rangle is a macroscopic superposition state. Weak macroscopic realism implies predetermined values λMA\lambda_{M}^{A} and λMB\lambda_{M}^{B} for the outcomes of measurements on the meter systems - the pointers are already in some kind of definite state that will indicate the result of the measurement to be either “spin up” or “spin down”.

In view of the correlation, it can then be argued that the systems AA and BB (which may be microscopic) are similarly specified to be in states with a definite outcome for the final measurement of spin components θ\theta and ϕ\phi, respectively. The interaction HMH_{M} is reversible, and hence the definition of wLR can be rephrased to apply to the system at the time tft_{f} where it is assumed the stage of the measurement that couples each system to a meter has already occurred, just after or in association with the unitary interactions UθAU_{\theta}^{A} and UϕBU_{\phi}^{B}. Due to the reversibility of HMH_{M}, this does not change the results of the paper.

III EPR-Bohm and GHZ paradoxes

III.1 Bohm’s version of the EPR paradox

Bohm generalized the EPR paradox to spin measurements by considering two spatially separated spin 1/21/2 particles prepared in the Bell state Bohm ; bell-1969

|ψB\displaystyle|\psi_{B}\rangle =\displaystyle= 12(|z|z|z|z).\displaystyle\frac{1}{\sqrt{2}}(|\uparrow\rangle_{z}|\downarrow\rangle_{z}-|\downarrow\rangle_{z}|\uparrow\rangle_{z}). (4)

The particles and their respective sites are denoted by AA and BB. Here |z|\uparrow\rangle_{z} and |z|\downarrow\rangle_{z} are the eigenstates of the zz component σz\sigma_{z} of the Pauli spin σ=(σx,σy,σz)\overrightarrow{\sigma}=(\sigma_{x},\sigma_{y},\sigma_{z}), with eigenvalues +1+1 and 1-1 respectively. We use the standard notation, where the first and second states of the product |||\uparrow\rangle|\downarrow\rangle refer to the states of particle AA and particle BB respectively. The spin operators for the two particles are distinguished by superscripts e.g. σzA\sigma_{z}^{A} and σzB\sigma_{z}^{B}.

III.1.1 A two-spin version

From (4), it is clear that the outcomes of spin-zz measurements on each particle are anticorrelated. Similarly, we may measure the component σy\sigma_{y} of each particle. To predict the outcomes, we transform the state into the yy basis, noting the transformation

|y\displaystyle|\uparrow\rangle_{y} =\displaystyle= 12(|z+i|z)\displaystyle\frac{1}{\sqrt{2}}(|\uparrow\rangle_{z}+i|\downarrow\rangle_{z})
|y\displaystyle|\downarrow\rangle_{y} =\displaystyle= 12(|zi|z)\displaystyle\frac{1}{\sqrt{2}}(|\uparrow\rangle_{z}-i|\downarrow\rangle_{z}) (5)

where |y|\uparrow\rangle_{y} and |y|\downarrow\rangle_{y} are the eigenstates of σy\sigma_{y}, with respective eigenvalues +1+1 and 1-1. Hence we find |z=(|y+|y)/2|\uparrow\rangle_{z}=(|\uparrow\rangle_{y}+|\downarrow\rangle_{y})/\sqrt{2} and |z=i(|y|y)/2|\downarrow\rangle_{z}=-i(|\uparrow\rangle_{y}-|\downarrow\rangle_{y})/\sqrt{2}. The state becomes in the new basis

|ψB\displaystyle|\psi_{B}\rangle =\displaystyle= i2(|y|y|y|y).\displaystyle\frac{i}{\sqrt{2}}(|\uparrow\rangle_{y}|\downarrow\rangle_{y}-|\downarrow\rangle_{y}|\uparrow\rangle_{y}). (6)

Denoting the respective measurements at each site by σyA\sigma_{y}^{A} and σyB\sigma_{y}^{B}, we see that the spin-yy outcomes at AA and BB are also anticorrelated.

An EPR-Bohm paradox follows from the following argument (Figure 1). By making a measurement of σzB\sigma_{z}^{B} on particle BB, the outcome for the measurement σzA\sigma_{z}^{A} on particle AA is known with certainty. EPR present their Assertions of local realism (LR) epr , summarized in Section II.A.1. Invoking EPR Assertion LR (2) that there is no disturbance to system AA due to the measurement at BB, EPR’s Assertion LR (1) therefore implies system AA can be ascribed a hidden variable λzA\lambda_{z}^{A} mermin-ghz , which predetermines the outcome for the measurement σzA\sigma_{z}^{A} should that measurement be performed. However, the outcomes at AA and BB are also anticorrelated for measurements of σy\sigma_{y} at both sites. The assumption of EPR’s premises therefore ascribes two hidden variables λzA\lambda_{z}^{A} and λyA\lambda_{y}^{A} to the system AA, which simultaneously predetermine the outcome of either σz\sigma_{z} or σy\sigma_{y} at AA, should either measurement be performed. This description is not compatible with any quantum wavefunction |ψ|\psi\rangle for the spin 1/21/2 system AA. The conclusion is that if EPR’s local realism is valid, then quantum mechanics gives an incomplete description of physical reality.

The above conclusions draw on the assumption that the subsystem AA is described quantum mechanically as a spin 1/21/2 system. For such a system, the Pauli spin variances defined by (Δσi)2=σi2σi2(\Delta\sigma_{i})^{2}=\langle\sigma_{i}^{2}\rangle-\langle\sigma_{i}\rangle^{2} satisfy the uncertainty relation (Δσx)2+(Δσy)2+(Δσz)22(\Delta\sigma_{x})^{2}+(\Delta\sigma_{y})^{2}+(\Delta\sigma_{z})^{2}\geq 2 hofmann-take . Since (Δσz)21(\Delta\sigma_{z})^{2}\leq 1, this implies hofmann-take

(Δσy)2+(Δσz)21.(\Delta\sigma_{y})^{2}+(\Delta\sigma_{z})^{2}\geq 1. (7)

For a quantum state description of the system AA, the values of σy\sigma_{y} and σz\sigma_{z} cannot be simultaneously precisely defined. A realisation has been given for two spin 1/21/2 particles (photons) which showed near-perfect correlation for both of two orthogonal spins (orthogonal linear polarizations), for a subensemble where both photons are detected aspect-Bohm .

III.1.2 Three-spin version

A stricter argument not dependent on the assumption of a spin 1/21/2 system is possible, if the experimentalist can measure the correlation of all three spin components Bohm ; epr-rmp ; bohm-test-uncertainty . Consider the spin-xx measurements σxA\sigma_{x}^{A} and σxB\sigma_{x}^{B}. The eigenstates of σx\sigma_{x} are

|x\displaystyle|\uparrow\rangle_{x} =\displaystyle= 12(|z+|z)\displaystyle\frac{1}{\sqrt{2}}(|\uparrow\rangle_{z}+|\downarrow\rangle_{z})
|x\displaystyle|\downarrow\rangle_{x} =\displaystyle= 12(|z|z).\displaystyle\frac{1}{\sqrt{2}}(|\uparrow\rangle_{z}-|\downarrow\rangle_{z}). (8)

The state (4) becomes in the spin-xx basis

|ψB\displaystyle|\psi_{B}\rangle =\displaystyle= 12(|x|x|x|x).\displaystyle\frac{1}{\sqrt{2}}(|\downarrow\rangle_{x}|\uparrow\rangle_{x}-|\uparrow\rangle_{x}|\downarrow\rangle_{x}). (9)

The spin-xx outcomes at AA and BB are also anticorrelated. According to the EPR premises, it is therefore possible to assign a hidden variable λxA\lambda_{x}^{A} to the subsystem AA that predetermines the outcome of the measurement σxA\sigma_{x}^{A}. Hence, local realism implies that the system AA would at any time be described by three precise values, λxA\lambda_{x}^{A}, λyA\lambda_{y}^{A} and λzA\lambda_{z}^{A}, which predetermine the outcomes of measurements σx\sigma_{x}, σy\sigma_{y} and σz\sigma_{z} respectively. Each of λx\lambda_{x}, λy\lambda_{y} and λz\lambda_{z} has the value +1+1 or 1-1 . Since always |λzA|=1|\lambda_{z}^{A}|=1, such a hidden variable description cannot be given by a local quantum state |ψ|\psi\rangle of AA, as this would be a violation of the quantum uncertainty relation

ΔσxΔσy|σz|,\Delta\sigma_{x}\Delta\sigma_{y}\geq|\langle\sigma_{z}\rangle|, (10)

which applies to all quantum states. Hence, we arrive at an EPR paradox, where local realism implies an inconsistency with the completeness of quantum mechanics.

III.2 GHZ paradox

The Greenberger-Horne-Zeilinger (GHZ) argument shows that local realism can be falsified, if quantum mechanics is correct ghz-1 ; mermin-ghz ; ghz-amjp ; clifton-ghz . The GHZ state

|ψGHZ\displaystyle|\psi_{GHZ}\rangle =\displaystyle= 12(|z|z|z|z|z|z)\displaystyle\frac{1}{\sqrt{2}}(|\uparrow\rangle_{z}|\uparrow\rangle_{z}|\uparrow\rangle_{z}-|\downarrow\rangle_{z}|\downarrow\rangle_{z}|\downarrow\rangle_{z}) (11)

involves three spatially separated spin 1/21/2 particles, AA, BB and CC. We denote the Pauli spin measurement σθ\sigma_{\theta} at the site J{A,B,C}J\in\{A,B,C\} by σθJ\sigma_{\theta}^{J}. Consider measurements of σx\sigma_{x} at each site. To obtain the predicted outcomes, we rewrite in the spin-xx basis. The GHZ state becomes

|ψGHZ\displaystyle|\psi_{GHZ}\rangle =\displaystyle= 12(|x|x|x+|x|x|x\displaystyle\frac{1}{\sqrt{2}}(|\downarrow\rangle_{x}|\uparrow\rangle_{x}|\uparrow\rangle_{x}+|\uparrow\rangle_{x}|\downarrow\rangle_{x}|\uparrow\rangle_{x} (12)
+|x|x|x+|z|z|z).\displaystyle+|\uparrow\rangle_{x}|\uparrow\rangle_{x}|\downarrow\rangle_{x}+|\downarrow\rangle_{z}|\downarrow\rangle_{z}|\downarrow\rangle_{z}).

From this we see σxAσxBσxC=1.\langle\sigma_{x}^{A}\sigma_{x}^{B}\sigma_{x}^{C}\rangle=-1. Now we also consider the measurement σxAσyBσyC\sigma_{x}^{A}\sigma_{y}^{B}\sigma_{y}^{C} on the system in the GHZ state. The GHZ state in the spin-yy basis is:

|ψGHZ\displaystyle|\psi_{GHZ}\rangle =\displaystyle= 12(|x|y|y+|x|y|y\displaystyle\frac{1}{\sqrt{2}}(|\uparrow\rangle_{x}|\uparrow\rangle_{y}|\uparrow\rangle_{y}+|\downarrow\rangle_{x}|\downarrow\rangle_{y}|\uparrow\rangle_{y} (13)
+|x|y|y+|x|y|y).\displaystyle+|\downarrow\rangle_{x}|\uparrow\rangle_{y}|\downarrow\rangle_{y}+|\uparrow\rangle_{x}|\downarrow\rangle_{y}|\downarrow\rangle_{y}).

This shows σxAσyBσyC=1.\langle\sigma_{x}^{A}\sigma_{y}^{B}\sigma_{y}^{C}\rangle=1. Similarly, σyAσxBσyC=1\langle\sigma_{y}^{A}\sigma_{x}^{B}\sigma_{y}^{C}\rangle=1 and σyAσyBσxC=1\langle\sigma_{y}^{A}\sigma_{y}^{B}\sigma_{x}^{C}\rangle=1.

The GHZ argument is well known. The outcome for σx\sigma_{x} at AA can be predicted with certainty by performing measurements σxB\sigma_{x}^{B} and σxC\sigma_{x}^{C}. The measurements do not disturb the system AA because the measurements at AA and those at BB and CC are spacelike-separated events. Similarly, the outcome for σyA\sigma_{y}^{A} can be predicted, without disturbing the system AA, by measurements of σyB\sigma_{y}^{B} and σxC\sigma_{x}^{C}. Hence, there exist hidden variables λxA\lambda_{x}^{A} and λyA\lambda_{y}^{A} that can be simultaneously ascribed to system AA, these variables predetermining the outcome for measurements σxA\sigma_{x}^{A} and σyA\sigma_{y}^{A} at AA. The variables assume the values of +1+1 or 1-1. A similar argument can be made for particles BB and CC. The contradiction with EPR’s local realism arises because the product λxAλxBλxC\lambda_{x}^{A}\lambda_{x}^{B}\lambda_{x}^{C} must equal 1-1, in order that the prediction σxAσxBσxC=1\langle\sigma_{x}^{A}\sigma_{x}^{B}\sigma_{x}^{C}\rangle=-1 holds. Similarly, λxAλyBλyC=λyAλxBλyC=λyAλyBλxC=1\lambda_{x}^{A}\lambda_{y}^{B}\lambda_{y}^{C}=\lambda_{y}^{A}\lambda_{x}^{B}\lambda_{y}^{C}=\lambda_{y}^{A}\lambda_{y}^{B}\lambda_{x}^{C}=1 in order that the prediction σxAσyBσyC=σyAσxBσyC=σyAσyBσxC=1\langle\sigma_{x}^{A}\sigma_{y}^{B}\sigma_{y}^{C}\rangle=\langle\sigma_{y}^{A}\sigma_{x}^{B}\sigma_{y}^{C}\rangle=\langle\sigma_{y}^{A}\sigma_{y}^{B}\sigma_{x}^{C}\rangle=1 holds. Yet, we see algebraically that λxAλxBλxC=λxAλxBλxC(λyB)2(λyA)2(λyC)2\lambda_{x}^{A}\lambda_{x}^{B}\lambda_{x}^{C}=\lambda_{x}^{A}\lambda_{x}^{B}\lambda_{x}^{C}(\lambda_{y}^{B})^{2}(\lambda_{y}^{A})^{2}(\lambda_{y}^{C})^{2}, and hence

λxAλxBλxC\displaystyle\lambda_{x}^{A}\lambda_{x}^{B}\lambda_{x}^{C} =\displaystyle= (λxAλyBλyC)(λxBλyAλyC)(λxCλyBλyA)\displaystyle(\lambda_{x}^{A}\lambda_{y}^{B}\lambda_{y}^{C})(\lambda_{x}^{B}\lambda_{y}^{A}\lambda_{y}^{C})(\lambda_{x}^{C}\lambda_{y}^{B}\lambda_{y}^{A}) (14)
=\displaystyle= 1\displaystyle 1

which gives a complete “all or nothing” contradiction. The conclusion is that local realism does not hold.

IV An EPR paradox based on weak local realism

An argument can be formulated that quantum mechanics is incomplete, based on the wLR premise. The argument follows along similar lines to the original EPR argument, and is depicted in Figure 2. The argument applies to the set-up considered by Schrödinger, involving simultaneous measurements, one direct and the other indirect sch-epr-exp-atom ; s-cat-1935 .

The system at time t0t_{0} is prepared in the Bell state |ψB|\psi_{B}\rangle. Let us choose the measurement setting ϕy\phi\equiv y such that the system BB is prepared for the pointer measurement of σyB\sigma_{y}^{B} (denoted SyBS_{y}^{B} on the diagram). From the anticorrelation of state (6), the outcome for σyA\sigma_{y}^{A} (i.e. SyAS_{y}^{A}) can be predicted with certainty, by measurement on system BB. This constitutes Schrödinger’s “indirect measurement”, of σyA\sigma_{y}^{A} (i.e. SyAS_{y}^{A}). Therefore, by Assertions wLR (1a) and (2a), the system AA at time tfB=tft_{f_{B}}=t_{f}, after UyBU_{y}^{B} has been performed, can be ascribed a definite value for the variable λyA\lambda_{y}^{A}. We note that a further unitary interaction UAU^{A} is required at AA after time tft_{f} so that the system is prepared for a pointer measurement σyA\sigma_{y}^{A}. Regardless, the final outcome is already determined at time tfBt_{f_{B}} by the value of λyA\lambda_{y}^{A}. This inferred variable is depicted as λyA\lambda_{y}^{A} in black in Figure 2.

However, at the time tft_{f}, the system AA is itself prepared for a pointer measurement of σzA\sigma_{z}^{A} (i.e. SzAS_{z}^{A}). Hence, by Assertion wLR (1b) and (2b), there is a hidden variable λzA\lambda_{z}^{A} that predetermines the value for the measurement σzA\sigma_{z}^{A}, should it be performed. This constitutes Schrödinger’s “direct measurement”, of σzA\sigma_{z}^{A} (i.e. SzAS_{z}^{A}). This variable is depicted as λzA\lambda_{z}^{A} in red in Figure 2. According to the premises, the system AA at the time tft_{f} therefore can be ascribed two definite spin values, λzA\lambda_{z}^{A} and λyA\lambda_{y}^{A}. This assignment cannot be given by any localised quantum state for a spin 1/21/2 system, and hence the argument can be put forward similarly to the original argument that quantum mechanics is incomplete.

In an experiment, it would be demonstrated that the result of SyAS_{y}^{A} can be inferred from the measurement at BB with certainty. It would also be established that system AA is given quantum mechanically as a spin 1/21/2 system. A description for a realistic experiment is given in Appendix D (see also Ref. sch-epr-exp-atom ).

Comment:

The above argument is based on a two-spin version of the EPR-Bohm argument. The three-spin version could not be formulated using wLR, because this would require preparation of three pointer measurements, which is not possible for the bipartite system. The original three-spin EPR-Bohm paradox requires the assumption of EPR’s LR, which can be falsified. The two-spin version is based on wLR which has not been falsified. On the other hand, the two-spin version of the EPR-Bohm paradox allows a counterargument against the incompleteness of quantum mechanics: It could be proposed that a local quantum state description is possible for AA, but that this description is a complex one, not describing a spin 1/21/2 particle.

V Macroscopic EPR-Bohm paradox using cat states

In this section, we strengthen the EPR argument by presenting cases where one may invoke macroscopic local realism. We do this by demonstrating an EPR-Bohm paradox which uses two macroscopically distinct states. First, we consider where the distinct states are coherent states. In the second example, the distinct states are a collection of multi-mode spin states with spins either all “up’, or all “down”. The unitary operations UθU_{\theta} that fix the measurement settings are chosen to preserve the macroscopic two-state nature of the system, and are realised by Kerr interactions and CNOT gates.

V.1 Two-spin EPR-Bohm paradox with coherent states

We first consider a realisation of the two-spin EPR paradox described Section III.A.1 using coherent and cat states. This requires macroscopic spins defined in terms of the two macroscopically distinct coherent states.

V.1.1 The initial state, unitary rotations, and definition of macroscopic spins

We consider the system to be prepared at time t1t_{1} in the entangled cat state cat-bell-wang-1 ; cat-det-map

|ψBell=𝒩(|α|β|α|β).|\psi_{Bell}\rangle=\mathcal{N}(|\alpha\rangle|-\beta\rangle-|-\alpha\rangle|\beta\rangle). (15)

Here |α|\alpha\rangle and |β|\beta\rangle are coherent states for single-mode fields AA and BB, and we take α\alpha and β\beta to be real, positive and large. 𝒩=12{1exp(2|α|22|β|2)}1/2\mathcal{N}=\frac{1}{\sqrt{2}}\{1-\exp(-2\left|\alpha\right|^{2}-2\left|\beta\right|^{2})\}^{-1/2} is the normalization constant. The phase of the coherent amplitudes α\alpha and β\beta are defined as real relative to an fixed axis, which is usually defined by a phase specified in the preparation process. For example, this is may be fixed by the phase of a pump field, as in the coherent state superpositions generated by nonlinear dispersion yurke-stoler-1 .

For each system AA and BB, one may measure the field quadrature phase amplitudes X^A=12(a^+a^)\hat{X}_{A}={\color[rgb]{1,0,0}{\color[rgb]{0,0,1}{\color[rgb]{0,0,0}\frac{1}{\sqrt{2}}}}}(\hat{a}+\hat{a}^{\dagger}), P^A=1i2(a^a^)\hat{P}_{A}={\color[rgb]{1,0,0}{\color[rgb]{0,0,1}{\color[rgb]{0,0,0}\frac{1}{i\sqrt{2}}}}}(\hat{a}-\hat{a}^{\dagger}), X^B=12(b^+b^)\hat{X}_{B}=\frac{1}{\sqrt{2}}(\hat{b}+\hat{b}^{\dagger}) and P^A=1i2(a^a^)\hat{P}_{A}={\color[rgb]{1,0,0}{\color[rgb]{0,0,1}{\color[rgb]{0,0,0}\frac{1}{i\sqrt{2}}}}}(\hat{a}-\hat{a}^{\dagger}), which are defined in a rotating frame yurke-stoler-1 . The boson destruction mode operators for modes AA and BB are denoted by a^\hat{a} and b^\hat{b}. As α\alpha \rightarrow\infty, the probability distribution P(XA)P(X_{A}) for the outcome XAX_{A} of the measurement X^A\hat{X}_{A} consists of two well-separated Gaussians which can be associated with the distributions for the coherent states |α|\alpha\rangle and |α|-\alpha\rangle. (Any central component due to interference vanishes for large α\alpha, β\beta). Hence, the outcome XAX_{A} distinguishes between the states |α|\alpha\rangle and |α|-\alpha\rangle. Similarly, X^B\hat{X}_{B} distinguishes between the states |β|\beta\rangle and |β|-\beta\rangle.

We define the outcome of the “spin” measurement S^A\hat{S}^{A} to be SA=+1S^{A}=+1 if XA0X_{A}\geq 0, and 1-1 otherwise. Similarly, the outcome of the measurement S^B\hat{S}^{B} is SB=+1S^{B}=+1 if XB0X_{B}\geq 0, and 1-1 otherwise. The result is identified as the spin of the system i.e. the qubit value. For each system, the coherent states become orthogonal in the limit of large α\alpha and β\beta, in which case the superposition maps to the two-qubit Bell state |ψBell=12(|+a|b|a|+b)|\psi_{Bell}\rangle=\frac{1}{\sqrt{2}}(|+\rangle_{a}|-\rangle_{b}-|-\rangle_{a}|+\rangle_{b}), given by (4). At time t1t_{1}, the outcomes SAS^{A} and SBS^{B} are anticorrelated.

Refer to caption
Figure 3: Macroscopic version of the EPR-Bohm paradox: The system is prepared in a cat state, e.g. (15), for which the final outcomes ++ and - correspond to the macroscopically distinct amplitudes α\alpha and α-\alpha. At each site AA and BB, a switch (dashed arrow) allows the independent and random choice to evolve the systems by UyU_{y}, or not. With no evolution, SzS_{z} is measured. If the rotation UyU_{y} takes place, SyS_{y} is measured. The outcomes for SzAS_{z}^{A} and SzBS_{z}^{B} (and SyAS_{y}^{A} and SyBS_{y}^{B}) are anticorrelated.

In order to realise the EPR-Bohm paradox, it is necessary to identify the noncommuting spin observables and the appropriate unitary rotations UU at each site required to measure these. For this purpose, we examine the systems AA and BB as they evolve independently according to local transformations UA(ta)U_{A}(t_{a}) and UB(tb)U_{B}(t_{b}), defined as

UA(ta)=eiHNLAta/,UB(tb)=eiHNLBtb/U_{A}(t_{a})=e^{-iH_{NL}^{A}t_{a}/\hbar},\ \ U_{B}(t_{b})=e^{-iH_{NL}^{B}t_{b}/\hbar} (16)

where

HNLA=Ωn^ak,HNLB=Ωn^bk.H_{NL}^{A}=\Omega\hat{n}_{a}^{k},\ \ H_{NL}^{B}=\Omega\hat{n}_{b}^{k}. (17)

Here, tat_{a} and tbt_{b} are the times of evolution at each site, n^a=a^a^\hat{n}_{a}=\hat{a}^{\dagger}\hat{a} and n^b=b^b^\hat{n}_{b}=\hat{b}^{\dagger}\hat{b}, and Ω\Omega is a constant. We consider k=2k=2, noting that k=4k=4 allows a Bell test manushan-bell-cat-lg . The dynamics of this evolution is well known yurke-stoler-1 ; collapse-revival-bec-2 ; collapse-revival-super-circuit-1 ; wright-walls-gar-1 . If the system AA is prepared in a coherent state |α|\alpha\rangle, then after a time ta=π/2Ωt_{a}=\pi/2\Omega the state of the system AA becomes manushan-cat-lg ; manushan-bell-cat-lg ; macro-bell-lg ; yurke-stoler-1 ; cat-states-super-cond ; cat-states-mirrahimi

Uπ/4A|α\displaystyle U_{\pi/4}^{A}|\alpha\rangle =\displaystyle= eiπ/4(|α+i|α)/2.\displaystyle e^{-i\pi/4}(|\alpha\rangle+i|-\alpha\rangle)/\sqrt{2}. (18)

Here we define Uπ/4A=UA(π/2Ω)U_{\pi/4}^{A}=U_{A}(\pi/2\Omega). A similar transformation Uπ/4BU_{\pi/4}^{B} is defined at BB for tb=π/2Ωt_{b}=\pi/2\Omega. This state is a superposition of two macroscopically distinct states, and is referred to as a cat state after Schrödinger’s paradox s-cat-1935 ; scat-rmp-frowis . Further interaction for the whole period ta=2π/Ωt_{a}=2\pi/\Omega returns the system to the coherent state |α|\alpha\rangle.

The macroscopic version of the EPR-Bohm paradox is depicted in Figure 3. We consider the spin-1/21/2 observables S^z=|++|||\hat{S}_{z}=|+\rangle\langle+|-|-\rangle\langle-|, S^x=|+|+|+|\hat{S}_{x}=|+\rangle\langle-|+|-\rangle\langle+| and S^y=1i(|+||+|)\hat{S}_{y}=\frac{1}{i}(|+\rangle\langle-|-|-\rangle\langle+|), defined for orthogonal states |±|\pm\rangle of a two-level system, which we also denote by |z|\uparrow\rangle_{z} and |z|\downarrow\rangle_{z}. Here, we identify the eigenstates |±|\pm\rangle of S^zA\hat{S}_{z}^{A} (S^zB\hat{S}_{z}^{B}) as the coherent states |±α|\pm\alpha\rangle (|±β|\pm\beta\rangle) respectively, with α\alpha and β\beta real, and in the limit of large α\alpha and β\beta where orthogonality is justified. In this limit, we define

S^zA\displaystyle\hat{S}_{z}^{A} =\displaystyle= |αα||αα|\displaystyle|\alpha\rangle\langle\alpha|-|-\alpha\rangle\langle-\alpha|
S^xA\displaystyle\hat{S}_{x}^{A} =\displaystyle= |αα|+|αα|\displaystyle|\alpha\rangle\langle-\alpha|+|-\alpha\rangle\langle\alpha|
S^yA\displaystyle\hat{S}_{y}^{A} =\displaystyle= 1i(|αα||αα|)\displaystyle\frac{1}{i}(|\alpha\rangle\langle-\alpha|-|-\alpha\rangle\langle\alpha|) (19)

for system AA. The scaling corresponds to Pauli spins σ=(σx,σy,σz)\overrightarrow{\sigma}=(\sigma_{x},\sigma_{y},\sigma_{z}). The spins S^zB\hat{S}_{z}^{B}, S^xB\hat{S}_{x}^{B} and S^yB\hat{S}_{y}^{B} for system BB are defined in identical fashion on replacing α\alpha with β\beta. We omit operator “hats”, where the meaning is clear.

V.1.2 Performing the measurement of spins SzS_{z} and SyS_{y}

The EPR-Bohm paradox requires measurement of SzAS_{z}^{A} and SzBS_{z}^{B}. The system in the state (15) is prepared for the pointer stage of the measurements of SzS_{z}. This is because for this system, a measurement of (the sign of) X^A\hat{X}_{A} and X^B\hat{X}_{B} is all that is required to complete the SzAS_{z}^{A} and SzBS_{z}^{B} measurement. The local measurement constitutes an optical homodyne, in which the fields are combined with a strong field across a beamsplitter with a relative phase shift ϑ\vartheta, followed by direct detection in the arms of the beam splitter yurke-stoler-1 . Here, ϑ\vartheta is chosen to measure X^A\hat{X}_{A} (X^B\hat{X}_{B}), the axis so that α\alpha (β\beta) as real. The ϑ\vartheta is defined by the preparation process, usually involving a pump field.

The EPR-Bohm argument also requires measurements of SyAS_{y}^{A} and SyBS_{y}^{B} on the Bell state (15) prepared at time t1t_{1} (Figure 3). Here, it is required to adjust the measurement-setting by applying a local unitary transformation UyU_{y} at each site.

The eigenstates of S^y\hat{S}_{y} are often written in the form |y=(|z+i|z)/2|\uparrow\rangle_{y}=(|\uparrow\rangle_{z}+i|\downarrow\rangle_{z})/\sqrt{2} and |y=(|zi|z)/2|\downarrow\rangle_{y}=(|\uparrow\rangle_{z}-i|\downarrow\rangle_{z})/\sqrt{2}, but the normalization can vary by a phase factor. We can abbreviate as |±y=12(|±+i|)|\pm\rangle_{y}=\frac{1}{\sqrt{2}}(|\pm\rangle+i|\mp\rangle), denoting ||\uparrow\rangle as |+|+\rangle, and ||\downarrow\rangle as ||-\rangle, interchangeably. We choose

|y\displaystyle|\uparrow\rangle_{y} =\displaystyle= eiπ/42(|z+i|z)\displaystyle\frac{e^{-i\pi/4}}{\sqrt{2}}(|\uparrow\rangle_{z}+i|\downarrow\rangle_{z})
|y\displaystyle|\downarrow\rangle_{y} =\displaystyle= eiπ/42(|z+i|z)=eiπ/42(|zi|z),\displaystyle\frac{e^{-i\pi/4}}{\sqrt{2}}(|\downarrow\rangle_{z}+i|\uparrow\rangle_{z})=\frac{e^{i\pi/4}}{\sqrt{2}}(|\uparrow\rangle_{z}-i|\downarrow\rangle_{z}),

and will denote the eigenstates at different sites by a subscript. We have temporarily dropped for convenience the superscripts and subscripts indicating the AA and BB, since the transformations are local and apply independently to both sites. It is readily verified that the S^y|y=|y\hat{S}_{y}|\uparrow\rangle_{y}=|\uparrow\rangle_{y} and S^y|y=|y\hat{S}_{y}|\downarrow\rangle_{y}=|\downarrow\rangle_{y} i.e. S^yA|±y,A=±|±y,A\hat{S}_{y}^{A}|\pm\rangle_{y,A}=\pm|\pm\rangle_{y,A} and S^yB|±y,B=±|±y,B\hat{S}_{y}^{B}|\pm\rangle_{y,B}=\pm|\pm\rangle_{y,B}.

Now we consider how to perform the measurement of S^y\hat{S}_{y}. As explained in Section II, the first stage of measurement involves a unitary operation UyU_{y}, giving a transformation to the measurement basis, so that the system is then prepared for the second stage of measurement, which is the “pointer [stage of] measurement” of S^y\hat{S}_{y}. The pointer stage constitutes a measurement of the sign S^\hat{S} of X^\hat{X}, which for large α\alpha (β\beta) will (after UyU_{y} has been applied) directly yield the outcome ±1\pm 1 for the system prepared in |±y|\pm\rangle_{y}. To establish UyU_{y}, following the procedure of Eqs. (4-6) and (11-13), any state

|ψ=c+|z+c||\psi\rangle=c_{+}|\uparrow\rangle_{z}+c_{-}|\downarrow\rangle (21)

written in the zz basis can be transformed into the yy basis, by substituting

|z\displaystyle|\uparrow\rangle_{z} \displaystyle\rightarrow (eiπ/4|y+eiπ/4|y)/2\displaystyle(e^{i\pi/4}|\uparrow\rangle_{y}+e^{-i\pi/4}|\downarrow\rangle_{y})/\sqrt{2}
|z\displaystyle|\downarrow\rangle_{z} \displaystyle\rightarrow i(eiπ/4|yeiπ/4|y)/2,\displaystyle-i(e^{i\pi/4}|\uparrow\rangle_{y}-e^{-i\pi/4}|\downarrow\rangle_{y})/\sqrt{2}, (22)

This gives

|ψ=d+|+y+d|y|\psi\rangle=d_{+}|+\rangle_{y}+d_{-}|-\rangle_{y} (23)

where d±=(c±ic)e±iπ/4d_{\pm}=(c_{\pm}\mp ic_{-})e^{\pm i\pi/4}. To obtain the transformed state (23), ready for the pointer stage of measurement of SyS_{y}, the system is thus evolved according to

Uy|ψBellU_{y}|\psi_{Bell}\rangle (24)

where UyUπ/41=U1(π/2Ω)U_{y}\equiv U_{\pi/4}^{-1}=U^{-1}(\pi/2\Omega). We explain this result further in the Appendix A for the purpose of clarity.

The UyU_{y} is the inverse of the transformation eiHNLt/e^{-iH_{NL}t/\hbar} where t=π/2Ωt=\pi/2\Omega, given by (18). The UyU_{y} is achieved by evolving the local system for a time t=π/2Ω3π/2Ωt=-\pi/2\Omega\equiv 3\pi/2\Omega, since the solutions are periodic.

Comment

The states |+y|+\rangle_{y} and |y|-\rangle_{y} refer to the macroscopically distinct coherent states |α|\alpha\rangle and |α|-\alpha\rangle defined at the time tt after the local unitary rotation UyU_{y} has taken place. This is important in identifying the macroscopic nature of the paradox. We then see that the premises of weak macroscopic realism defined in Section II.A.2 will apply (Figure 2).

V.1.3 EPR-Bohm argument

The EPR-Bohm argument is as follows (Figure 3). Consider the system prepared in the Bell state (15) at time t1t_{1} (α\alpha, β\beta\rightarrow\infty). One first measures SzAS_{z}^{A} and SzBS_{z}^{B} for this state at time t1t_{1}. The anticorrelation of the Bell state means that the result for SzAS_{z}^{A} at AA can be predicted with certainty by the measurement at BB.

The EPR-Bohm argument continues, by considering measurements of SyAS_{y}^{A} and SyBS_{y}^{B} on the Bell state (15) prepared at time t1t_{1} (Figure 3). The measurement of SyAS_{y}^{A} (SyBS_{y}^{B}) is thus made by applying the local unitary rotation UyAU_{y}^{A} (UyBU_{y}^{B}) to the Bell state prepared at t1t_{1}, followed by a measurement of the sign of XAX_{A} (XBX_{B}). The state of the system prepared after the unitary rotations UyAU_{y}^{A} and UyBU_{y}^{B} is also given as the Bell state (15), which we write as

|ψBelly,y=12(|y|+y|+y|y)|\psi_{Bell}\rangle_{y,y}=\frac{1}{\sqrt{2}}(|-\rangle_{y}|+\rangle_{y}-|+\rangle_{y}|-\rangle_{y}) (25)

where we have taken α\alpha, β\beta large. As with the original paradox given in Section III.A.1, as seen by Eq. (6), the final measurements of XAX_{A} and XBX_{B} therefore reveal an anticorrelation between SyAS_{y}^{A} and SyBS_{y}^{B}, and the result for SyAS_{y}^{A} can be revealed, with certainty, by measurement of SyBS_{y}^{B} at site BB. Here, we note the Comment in the above section, that |±y|\pm\rangle_{y} are the macroscopically distinct coherent states |α|\alpha\rangle and |α|-\alpha\rangle (or |β|\beta\rangle and |β|-\beta\rangle) that are realised at the time tft_{f} in Figure 3, which corresponds to the time after the transformation UyU_{y} has been carried out at each location.

The Bohm-EPR argument continues. The correlation between the spins enables an experimentalist at BB to determine with certainty either SzAS_{z}^{A} or SyAS_{y}^{A}, for the system prepared at t1t_{1}, by choosing the suitable measurement at the site BB. Assuming EPR’s local realism, this implies that both the spins of system AA are predetermined with certainty. Following along the lines of the two-spin paradox of Section III.A.1, this constitutes Bohm’s EPR paradox, because it is not possible to define a local quantum state φ\varphi for the spin 1/21/2 system AA at the time t1t_{1} with simultaneously specified values for both S^zA\hat{S}_{z}^{A} and S^yA\hat{S}_{y}^{A}. The paradox is the inconsistency between macroscopic local realism and the completeness of quantum mechanics.

In the gedanken experiment, it is assumed that the system AA is described quantum mechanics as a spin 1/21/2 system, which is valid as α\alpha\rightarrow\infty, where the two coherent states |α|\alpha\rangle and |α|-\alpha\rangle are orthogonal. In the same limit, the spin outcomes for S^zA\hat{S}_{z}^{A} and S^zB\hat{S}_{z}^{B}, and also for S^yA\hat{S}_{y}^{A} and S^yB\hat{S}_{y}^{B}, are perfectly anticorrelated, so that this realisation of the EPR-Bohm paradox strictly follows in the macroscopic limit, where α\alpha\rightarrow\infty. Proposals for finite α\alpha that also account for imperfect anticorrelation of the spins are presented in Appendices C and D.

V.2 Two- and three-spin paradox with spins and CNOT gates

A useful realization uses multimode spin states and CNOT gates. This allows a realisation of both types of EPR-Bohm paradox presented in Section III.A, the two- and three-spin versions, at an increasingly macroscopic level depending on the number of modes. Here, because the spin qubits correspond to macroscopically distinct states, the paradoxes will reveal an inconsistency between macroscopic local realism and the completeness of quantum mechanics.

By analogy with the microscopic example of Section III.A.2, the three-spin paradox requires a transformation UxU_{x} at each site, where (apart from phase factors)

Ux1|\displaystyle U_{x}^{-1}|\uparrow\rangle \displaystyle\rightarrow 12(|+|)\displaystyle\frac{1}{\sqrt{2}}(|\uparrow\rangle+|\downarrow\rangle)
Ux1|\displaystyle U_{x}^{-1}|\downarrow\rangle \displaystyle\rightarrow 12(||),\displaystyle\frac{1}{\sqrt{2}}(|\uparrow\rangle-|\downarrow\rangle), (26)

as well as that for UyU_{y}, given as

Uy1|z\displaystyle U_{y}^{-1}|\uparrow\rangle_{z} \displaystyle\rightarrow 12(|+i|)\displaystyle\frac{1}{\sqrt{2}}(|\uparrow\rangle+i|\downarrow\rangle)
Uy1|z\displaystyle U_{y}^{-1}|\downarrow\rangle_{z} \displaystyle\rightarrow 12(|i|).\displaystyle\frac{1}{\sqrt{2}}(|\uparrow\rangle-i|\downarrow\rangle). (27)

The important step is to find a Hamiltonian that gives a realisation of UxU_{x} and UyU_{y}. In the previous section, for the cat-states involving the coherent states, a transformation UyU_{y} was specified but not for UxU_{x}. We note that cat-state superpositions |α±|α|\alpha\rangle\pm|-\alpha\rangle can be created using conditional measurements cat-state-phil ; cat-states-super-cond ; cat-det-map , and open dissipative systems cat-states-wc ; cat-even-odd-transient ; transient-cat-states-leo ; cat-dynamics-ry ; cats-hach ; cat-states-mirrahimi ; cat-det-map ; cat-states-super-cond . However, we prefer to use simple unitary transformations. A realisation based on NOON states is given in Appendix B.

A realisation can be achieved using an array of spins. The qubits of (27) become the macroscopically distinct states ||N|\uparrow\rangle\equiv|\uparrow\rangle^{\otimes N} and ||N|\downarrow\rangle\equiv|\downarrow\rangle^{\otimes N}, for large NN, so that the initial Bell state (4) becomes the two-site GHZ state

|ψBellz,z=12(|z,AN|z,BN|z,AN|,BN).|\psi_{Bell}\rangle_{z,z}=\frac{1}{\sqrt{2}}(|\uparrow\rangle_{z,A}^{\otimes N}|\uparrow\rangle_{z,B}^{\otimes N}-|\downarrow\rangle_{z,A}^{\otimes N}|\downarrow\rangle_{,B}^{\otimes N}). (28)

The premises of macroscopic realism can be applied to the macroscopically distinct states. Here, |z,JN=k=1N|J,k|\uparrow\rangle_{z,J}^{\otimes N}=\prod_{k=1}^{N}|\uparrow\rangle_{J,k} where |J,k|\uparrow\rangle_{J,k} is the eigenstate of the Pauli spin σzk\sigma_{z}^{k} for the mode labelled kk at site JJ, the collection of modes k=1,..Nk=1,..N forming the system labelled JJ. The |JN|\uparrow\rangle_{J}^{\otimes N} and |JN|\downarrow\rangle_{J}^{\otimes N} represent macroscopically distinct states, with collective Pauli spin values of NN or N-N, and are eigenstates of the spin product SzJ=k=1NσzkS_{z}^{J}=\prod_{k=1}^{N}\sigma_{z}^{k}.

In order to realise the paradox, the transformations UU needed at each site JJ are, for UxU_{x} and UyU_{y}, of the form (26)-(27), but where we replace ||N|\uparrow\rangle\equiv|\uparrow\rangle^{\otimes N} and ||N|\downarrow\rangle\equiv|\downarrow\rangle^{\otimes N}. Generally, one can first consider how to achieve

|Ncosθ2|N+eiϑsinθ2|N.|\uparrow\rangle^{\otimes N}\rightarrow\cos\frac{\theta}{2}|\uparrow\rangle^{\otimes N}+e^{i\vartheta}\sin\frac{\theta}{2}|\downarrow\rangle^{\otimes N}. (29)

Following the experiment described in IBM-macrorealism-1 , the unitary transformations UxU_{x} and UyU_{y} are made in two steps.

The first step is a rotation on the single-mode spin |1(10)|\uparrow\rangle_{1}\equiv\begin{pmatrix}1\\ 0\end{pmatrix}, |1(01)|\downarrow\rangle_{1}\equiv\begin{pmatrix}0\\ 1\end{pmatrix}, given by the unitary matrix Uθ,ϑ=(cosθ2sinθ2eiϑsinθ2eiϑcosθ2)U_{\theta,\vartheta}=\begin{pmatrix}\cos\frac{\theta}{2}&-\sin\frac{\theta}{2}\\ e^{i\vartheta}\sin\frac{\theta}{2}&e^{i\vartheta}\cos\frac{\theta}{2}\end{pmatrix} where ϑ=0\vartheta=0 or π/2\pi/2, which transforms the spin as

|1\displaystyle|\uparrow\rangle_{1} \displaystyle\rightarrow Uθ,ϑ|1=cosθ2|1+eiϑsinθ2|1\displaystyle U_{\theta,\vartheta}|\uparrow\rangle_{1}=\cos\frac{\theta}{2}|\uparrow\rangle_{1}+e^{i\vartheta}\sin\frac{\theta}{2}|\downarrow\rangle_{1}
|1\displaystyle|\downarrow\rangle_{1} \displaystyle\rightarrow Uθ,ϑ|1=sinθ2|1+eiϑcosθ2|1.\displaystyle U_{\theta,\vartheta}|\downarrow\rangle_{1}=-\sin\frac{\theta}{2}|\uparrow\rangle_{1}+e^{i\vartheta}\cos\frac{\theta}{2}|\downarrow\rangle_{1}. (30)

Here, we drop the subscript JJ representing the site, for notational simplicity. Choosing θ=π/2\theta=\pi/2 gives the starting point for the transformation UxU_{x} or UyU_{y} at each site, with ϑ=0\vartheta=0 or π/2\pi/2 respectively.

A common physical realisation of the spin qubit involves two polarisation modes: ||1,0|\uparrow\rangle\equiv|1,0\rangle and ||0,1|\downarrow\rangle\equiv|0,1\rangle defined for two modes a±a_{\pm} as in Appendix B. The transformation Uθ,ϑU_{\theta,\vartheta} can then be achieved with a polarizing beam splitter, with mode transformations (a^±\hat{a}_{\pm} are boson operators defining the modes)

c^+\displaystyle\hat{c}_{+} =\displaystyle= a^+cosθa^sinθ\displaystyle\hat{a}_{+}\cos\theta-\hat{a}_{-}\sin\theta
eiϑc^\displaystyle e^{i\vartheta}\hat{c}_{-} =\displaystyle= a^+sinθ+a^cosθ.\displaystyle\hat{a}_{+}\sin\theta+\hat{a}_{-}\cos\theta. (31)

The c^±\hat{c}_{\pm} are boson operators for the outgoing modes emerging from the beam splitter. The interaction is described by the Hamiltonian H=ik(a^+a^a^+a^)H=i\hbar k(\hat{a}_{+}\hat{a}_{-}^{\dagger}-\hat{a}_{+}^{\dagger}\hat{a}_{-}) where θ=kt\theta=kt, for ϑ=0\vartheta=0. The addition of a ϑ=π/2\vartheta=\pi/2 phase shift (or not) relative to the two outputs gives the mode transformations with the dependence on ϑ=0\vartheta=0 or π/2\pi/2. If the input is ||\uparrow\rangle, the output state is

|1,0in\displaystyle|1,0\rangle_{in} =\displaystyle= a^+|0\displaystyle\hat{a}_{+}^{\dagger}|0\rangle (32)
=\displaystyle= cosθ|1,0out+eiϑsinθ|0,1out.\displaystyle\cos\theta|1,0\rangle_{out}+e^{i\vartheta}\sin\theta|0,1\rangle_{out}.

If the input is ||\downarrow\rangle, the output is found according to

|0,1in\displaystyle|0,1\rangle_{in} =\displaystyle= a^|0\displaystyle\hat{a}_{-}^{\dagger}|0\rangle (33)
=\displaystyle= sinθ|1,0out+eiϑcosθ|0,1out\displaystyle-\sin\theta|1,0\rangle_{out}+e^{i\vartheta}\cos\theta|0,1\rangle_{out}

which gives a starting point for the transformation UxU_{x} (where ϑ=0\vartheta=0) and UyU_{y} (where ϑ=π/2\vartheta=\pi/2) at each site JJ.

The second step of the transformations UxU_{x} and UyU_{y} involves a sequence of CNOT gates. Consider the example of two qubits, with the initial state |00|||00\rangle\equiv|\uparrow\rangle|\uparrow\rangle. The transformation Uθ,ϑU_{\theta,\vartheta} on the first qubit evolves the state into:

Uθ,ϑ||\displaystyle U_{\theta,\vartheta}|\uparrow\rangle|\uparrow\rangle =cosθ2||+eiϑsinθ2||.\displaystyle=\cos\frac{\theta}{2}|\uparrow\rangle|\uparrow\rangle+e^{i\vartheta}\sin\frac{\theta}{2}|\downarrow\rangle|\uparrow\rangle. (34)

The subsequent CNOT gate then flips the second (target) qubit to |1||1\rangle\equiv|\downarrow\rangle if the first (control) qubit is |1|1\rangle. For n>2n>2, the CNOT gates will be performed between the first qubit and all other qubits. This gives

Uθ,ϑ|N\displaystyle U_{\theta,\vartheta}|\uparrow\rangle^{\otimes N} =cosθ2|N+eiϑsinθ2|N.\displaystyle=\cos\frac{\theta}{2}|\uparrow\rangle^{\otimes N}+e^{i\vartheta}\sin\frac{\theta}{2}|\downarrow\rangle^{\otimes N}. (35)

In this way, the transformations (26)-(27) for UxU_{x} and UyU_{y} can be achieved macroscopically (for large NN) for each site.

In the two-spin experiment, either UyU_{y} or UzU_{z} is selected at each site, in order to measure SyJS_{y}^{J} or SzJS_{z}^{J}. We specify that the initial state |ψBellz,z|\psi_{Bell}\rangle_{z,z} (Eq. (28)) has been prepared for the pointer measurement of SzJS_{z}^{J}. This means that a direct detection of the qubit value (such as a direct detection of a photon in the mode a+a_{+} or aa_{-}) is all that is required to complete the measurement of SzJS_{z}^{J}.

The experiment of IBM-macrorealism-1 used the IBM quantum computer to perform the operations with N=26N=2-6, enabling a test of macrorealism. In a macroscopic realisation, similar operations have been performed using Rydberg atoms, for N20N\sim 20 omran-cats .

The analysis given in Section V.A above follows for this example, on replacing the macroscopically distinct states |α|\alpha\rangle and |α|-\alpha\rangle with |N|\uparrow\rangle^{\otimes N} and |N|\downarrow\rangle^{\otimes N}. One can define the macroscopic spins and the eigenstates |±y|\pm\rangle_{y} and |±x|\pm\rangle_{x} of SyS_{y} and SxS_{x} similarly. Following the Comment in Section V.A, the states after the transformations UyU_{y} and UxU_{x} are thus superpositions of the two macroscopically distinct states |N|\uparrow\rangle^{\otimes N} and |N|\downarrow\rangle^{\otimes N} (for large NN), which are prepared for the pointer measurement. Hence, the premises of weak macroscopic realism defined in Section II.A.2 apply (Figure 2). The application of the premises macroscopic local realism and deterministic macroscopic realism is explained in the next section. The macroscopic paradoxes map onto the microscopic ones discussed in Section III, and the predictions for the correlations follow accordingly.

VI Conclusions from the macroscopic EPR-Bohm paradox

The EPR-Bohm paradoxes of Section V involving cat states give a stronger version of the EPR argument. The predetermined values for the spins are macroscopically distinct, being the amplitudes α\alpha and α-\alpha, or else the collective Pauli spin values of NN and N-N. Two types of EPR paradox based on macroscopic realism can be put forward. The first is based on MLR (or deterministic macroscopic realism), which can be falsified. The second is based on weak macroscopic realism.

VI.1 EPR paradox based on deterministic macroscopic realism

Macroscopic local realism (MLR) is EPR’s local realism when applied to the system of Figure 3, where the outcomes ++ and - imply macroscopically distinct states for the system defined at time t1t_{1} (Section II.A.2). The Locality Assertion LR (2) becomes more convincing, since any disturbance of AA due to the measurement at BB would then require a macroscopic change of the state at AA. The application of EPR’s local realism to the system for both measurements SzS_{z} and SyS_{y} leads to the conclusion that system AA is described simultaneously by both hidden variables λzA\lambda_{z}^{A} and λyA\lambda_{y}^{A} at the time t1t_{1} (and for the three-spin paradox, similarly for SxS_{x}). The macroscopic paradox therefore indicates inconsistency between MLR and the completeness of quantum mechanics. It is important however, that we justify the application of MLR to both measurements.

According to the definition given by Leggett and Garg, the premises of MLR and dMR require identification of two macroscopically distinct states that the system at that time “has available to it”. At the time t1t_{1}, the systems of Section V.A and V.B are superpositions of two states |z|\uparrow\rangle_{z} and |z|\downarrow\rangle_{z} (|α|\alpha\rangle and |α|-\alpha\rangle, or |N|\uparrow\rangle^{\otimes N} or |N|\downarrow\rangle^{\otimes N}) that can be regarded as macroscopically distinct (for large α\alpha and NN). These states are prepared for the pointer measurement SzS_{z}.

The application of the premises to the set-up also requires that the states |y|\uparrow\rangle_{y} and |y|\downarrow\rangle_{y} distinguished by the measurement S^y\hat{S}_{y} be regarded as macroscopically distinct at this time t1t_{1}. The eigenstates can be represented as superpositions e.g. |z±|z|\uparrow\rangle_{z}\pm|\downarrow\rangle_{z} of the macroscopically distinct states say, |α|\alpha\rangle and |α|-\alpha\rangle, at this time. It is argued that the superpositions represented by the different probability amplitudes are macroscopically distinct, because two basis states are. The distinction can be made macroscopic in terms of the pointer basis, by applying a unitary transformation UyU_{y} which does not involve amplification.

The macroscopic versions of the EPR-Bohm paradox can be based on deterministic macroscopic realism alone, defined in Section II.A.2. The term “deterministic” is used, because in the context of the EPR-Bohm setup, the premise implies that the system (at the time t1t_{1}) is simultaneously specified by both hidden variables, λzA\lambda_{z}^{A} and λyA\lambda_{y}^{A}. The outcome for measurement of either SzAS_{z}^{A} or SyAS_{y}^{A} at AA is considered predetermined, without regard to the measurement apparatus, as in classical mechanics.

We now argue that the assumption of deterministic macroscopic realism (dMR) is equivalent to that of macroscopic local realism (MLR) for the macroscopic EPR set-up. In the EPR set-up, for any macroscopic spin SθAS_{\theta}^{A} (θz,y\theta\equiv z,y), one may determine which of two macroscopically distinct states the macroscopic system AA is in, without disturbing system AA, by performing a spacelike separated measurement on BB. Thus, for the Bohm example where we realize anticorrelated outcomes between AA and BB, dMR is implied by MLR. The converse is also true. The premise of dMR is that system AA already be in a state with predetermined value λ\lambda for the spin SAS^{A} (whether SzAS_{z}^{A} or SyAS_{y}^{A}), prior to the measurement being performed. The locality assumption at a macroscopic level is naturally part of the definition of dMR: The value of λA\lambda^{A} cannot be affected by measurements performed on a spacelike separated system BB. The anticorrelation allows determination of the predetermined value for AA, given a measurement at BB. Thus, it follows that dMR implies MLR. Hence, we use the terms dMR and MLR interchangeably in this paper.

We mention that Leggett and Garg motivated tests of macroscopic realism legggarg-1 . However, in order to establish a test, the additional assumption of noninvasive measurability was introduced for single systems. Therefore, reports of violations of Leggett-Garg inequalities (e.g. asadian-lg ; emary-review ; IBM-macrorealism-1 ; leggett-garg-uola ; NSTmunro-1-1 ; manushan-cat-lg ) do not imply falsification of macroscopic realism, but rather of the combined premises of macrorealism.

The EPR-Bohm paradox for cat states thus illustrates inconsistency between dMR (or MLR) and the notion that quantum mechanics is a complete theory. However, dMR (and MLR) can be falsified by violations of Bell inequalities for cat states manushan-bell-cat-lg ; manushan-cat-lg ; macro-bell-lg ; macro-bell-jeong . We show in Section VII that dMR (and MLR) can also be falsified in a macroscopic GHZ set-up.

VI.2 An EPR paradox based on weak macroscopic realism

It is also possible to make an argument for the incompleteness of quantum mechanics, based on the premise of weak macroscopic realism (wMR) (Figure 4). The macroscopic paradox follows along the same lines as that for weak local realism, given in Section IV, except that the outcomes ++ and - for the spins SzS_{z} and SyS_{y} can be shown to correspond to macroscopically distinct states for the system measured at time tft_{f}. The EPR-Bohm paradox based on wMR is stronger than that based on deterministic macroscopic realism, or macroscopic local realism, because the assumption of wMR is weaker and is not falsified by the GHZ or Bell predictions.

Refer to caption
Figure 4: A macroscopic EPR-Bohm paradox based on weak macroscopic realism: The system is as described for Figure 3. At the time tft_{f}, the sketch depicts the choice to measure SzAS_{z}^{A} and SyBS_{y}^{B}. Hence, wMR ascribes to system AA at time tft_{f} a predetermined value λzA\lambda_{z}^{A} for the outcome for SzAS_{z}^{A}, since the system has been prepared for a pointer measurement of SzAS_{z}^{A}. The final outcomes for SzAS_{z}^{A} and SzBS_{z}^{B} (and SyAS_{y}^{A} and SyBS_{y}^{B}) are anticorrelated. Hence, wMR also ascribes a hidden variable λyA\lambda_{y}^{A} for system AA that predetermines the outcome SyAS_{y}^{A}.

The argument for the inconsistency between wMR and the notion of the completeness of quantum mechanics is illustrated in Figure 4. At time tft_{f} the system BB has undergone the evolution UyBU_{y}^{B} to prepare the system BB for the pointer measurement of SyBS_{y}^{B}, whereas system AA is prepared for the pointer measurement of SzAS_{z}^{A}. The premise wMR asserts by Assertion (1b) that at time tft_{f}, a value λzA\lambda_{z}^{A} predetermines the outcome for SzAS_{z}^{A} at AA, and similarly, a value λyB\lambda_{y}^{B} predetermines the outcome for SyBS_{y}^{B} at BB. By Assertion (1a), because of the anticorrelation between AA and BB, the value of λyA=λyB\lambda_{y}^{A}=-\lambda_{y}^{B} also predetermines the outcome for SyAS_{y}^{A} at AA, at the time tft_{f} (even though a further unitary rotation at AA would be necessary to carry out the measurement). Thus, wMR asserts that the system AA at the time tft_{f} can be simultaneously assigned values λzA\lambda_{z}^{A} and λyA\lambda_{y}^{A} predetermining the results of measurements SzAS_{z}^{A} and SyAS_{y}^{A}. Hence, there is an EPR paradox.

The quantum correlations of the macroscopic EPR-Bohm, Bell and GHZ paradoxes are consistent with wMR, because the systems are prepared for a pointer measurement of SzS_{z} at one time t1t_{1}, and then can be prepared for a pointer measurement of SyS_{y} at a later time tft_{f}, after further unitary rotations UAU_{A} or UBU_{B}. The hidden variables for the EPR-Bohm paradox are tracked in Figure 4. We note wMR does not assert that at the time tft_{f}, the value of an arbitrary third measurement SθS_{\theta} is predetermined prior to the unitary rotation, since that rotation UU has not been performed at site AA or BB.

For a bipartite system, it is the introduction of a third measurement setting that leads to the falsification of dMR, as evident by the Bell tests which require three or more different measurement settings bell-1969 ; manushan-bell-cat-lg . A falsification is possible for dMR, because the premise dMR asserts that the system AA (or BB) has simultaneously predetermined values for the outcomes of all pointer measurements, at the time t1t_{1}, prior to unitary dynamics UU that finalizes the choice of measurement setting.

The ideal experiment realizing the paradox based on wMR would establish that the outcome for SyAS_{y}^{A} can be inferred with certainty from the measurement at BB. It would also establish that systems AA and BB are in a superposition (or mixture) of the two relevant macroscopically distinct states. It is also necessary to demonstrate that the system AA is a spin 1/21/2 system, as in Eqs. (19), e.g. demonstrating both measurements SzAS_{z}^{A} and SyAS_{y}^{A} and the relation (7). Conditions for a realistic experiment are given in Appendix D.

VII GHZ cat gedanken experiment

The GHZ argument outlined in Section III.B becomes macroscopic when the spins ||\uparrow\rangle and ||\downarrow\rangle correspond to macroscopically distinct states. The macroscopic set-up begins with the preparation at time t1t_{1} of the GHZ state

|ψGHZ\displaystyle|\psi_{GHZ}\rangle =\displaystyle= 12(|z,A|z,B|z,C|z,A|z,B|z,C)\displaystyle\frac{1}{\sqrt{2}}(|\uparrow\rangle_{z,A}|\uparrow\rangle_{z,B}|\uparrow\rangle_{z,C}-|\downarrow\rangle_{z,A}|\downarrow\rangle_{z,B}|\downarrow\rangle_{z,C})

where |z,J|z,JN|\uparrow\rangle_{z,J}\equiv|\uparrow\rangle_{z,J}^{\otimes N} and |z,J|z,JN|\downarrow\rangle_{z,J}\equiv|\downarrow\rangle_{z,J}^{\otimes N}, defined in Section V.B, are eigenstates of SzJ=k=1NσzkS_{z}^{J}=\prod_{k=1}^{N}\sigma_{z}^{k} with eigenvalues 11 and 1-1 respectively. Here, JA,B,CJ\equiv A,B,C denotes the site. As explained in Section V.B, the system is prepared at t1t_{1} for a pointer measurement of SzASzBSzCS_{z}^{A}S_{z}^{B}S_{z}^{C}. One then considers the measurements of SxASxBSxCS_{x}^{A}S_{x}^{B}S_{x}^{C} and SxASyBSyCS_{x}^{A}S_{y}^{B}S_{y}^{C}. By analogy with the microscopic example, this involves applying the transformations UxU_{x} or UyU_{y} given by (26)-(27) at each site. After the interactions UxAU_{x}^{A}, UxBU_{x}^{B} and UxCU_{x}^{C}, the system is prepared for the pointer measurement of SxASxBSxCS_{x}^{A}S_{x}^{B}S_{x}^{C}. The state in the new basis is

|ψGHZ\displaystyle|\psi_{GHZ}\rangle =\displaystyle= 12(|x,A|x,B|x,C+|x,A|x,B|x,C\displaystyle\frac{1}{2}(|\downarrow\rangle_{x,A}|\uparrow\rangle_{x,B}|\uparrow\rangle_{x,C}+|\uparrow\rangle_{x,A}|\downarrow\rangle_{x,B}|\uparrow\rangle_{x,C}
+|x,A|x,B|x,C+|x.A|x.B|x,C)\displaystyle+|\uparrow\rangle_{x,A}|\uparrow\rangle_{x,B}|\downarrow\rangle_{x,C}+|\downarrow\rangle_{x.A}|\downarrow\rangle_{x.B}|\downarrow\rangle_{x,C})

The product of the spins is SxASxBSxC=1S_{x}^{A}S_{x}^{B}S_{x}^{C}=-1.

If we evolve the state (LABEL:eq:ghz-cat) with UxAU_{x}^{A}, UyBU_{y}^{B} and UyCU_{y}^{C}, the system is prepared for a pointer measurement of SxASyBSyCS_{x}^{A}S_{y}^{B}S_{y}^{C}. In the new basis,

|ψGHZ\displaystyle|\psi_{GHZ}\rangle =\displaystyle= 14(|y,A|y,B|y,C+|y,A|y,B|y,C\displaystyle\frac{1}{4}(|\uparrow\rangle_{y,A}|\uparrow\rangle_{y,B}|\uparrow\rangle_{y,C}+|\downarrow\rangle_{y,A}|\downarrow\rangle_{y,B}|\uparrow\rangle_{y,C}
+|y,A|y,B|y,C+|y,A|y,B|y,C).\displaystyle+|\downarrow\rangle_{y,A}|\uparrow\rangle_{y,B}|\downarrow\rangle_{y,C}+|\uparrow\rangle_{y,A}|\downarrow\rangle_{y,B}|\downarrow\rangle_{y,C}).

Always, SxASyBSyC=1S_{x}^{A}S_{y}^{B}S_{y}^{C}=1. Similarly, we consider SyASyBSxCS_{y}^{A}S_{y}^{B}S_{x}^{C} and SyASxBSyCS_{y}^{A}S_{x}^{B}S_{y}^{C}, and arrive at the GHZ contradiction, as for the microscopic case. The unitary interactions UxU_{x} and UyU_{y} were shown possible using CNOT gates in Section V.B.

Refer to caption
Figure 5: Set-up for the GHZ paradox with cat states. The outcome of SxS_{x} (and SyS_{y}) at each of the sites AA, BB and CC can be predicted with certainty, by choosing certain spacelike separated measurements at the other two sites. EPR’s (macroscopic) local realism implies the outcomes are predetermined by variables λx\lambda_{x} and λy\lambda_{y} at the time t1t_{1}, as indicated on the diagram, which gives the GHZ contradiction. The hidden variables can also be deduced from the premise of deterministic macroscopic realism. The GHZ contradiction is a falsification of deterministic macroscopic realism.

VIII Conclusions from the GHZ-cat gedanken experiment

The macroscopic GHZ set-up enables a falsification of the MLR premises, and hence is a stronger version of the GHZ experiment. This is because the states |z,J|z,JN|\uparrow\rangle_{z,J}\equiv|\uparrow\rangle_{z,J}^{\otimes N} and |z,J|z,JN|\downarrow\rangle_{z,J}\equiv|\downarrow\rangle_{z,J}^{\otimes N} are macroscopically distinct for large NN, and the transformations UxU_{x} and UyU_{y} given by (26)-(27) create superpositions of the macroscopically distinct states. Applying the justification given in Section VI.A that the eigenstates |y,J|\uparrow\rangle_{y,J} and |y,J|\downarrow\rangle_{y,J} (and |x,J|\uparrow\rangle_{x,J} and |x,J|\downarrow\rangle_{x,J}) are also macroscopically distinct, the hidden variables λxA\lambda_{x}^{A}, λyA\lambda_{y}^{A}, λxB\lambda_{x}^{B}, λyB\lambda_{y}^{B}, λxC\lambda_{x}^{C}, λyC\lambda_{y}^{C} defined for the GHZ system in Section III.B are deduced based on MLR. MLR asserts that the spacelike measurement at BB or CC cannot induce a macroscopic change to the system AA. This is a weaker assumption than local realism, which rules out all changes. The GHZ contradiction explained in Section VI.A falsifies MLR.

VIII.1 Falsification of deterministic macroscopic realism

The GHZ paradox as applied to the cat states is also a falsification of deterministic macroscopic realism (dMR). We may present the GHZ paradox directly from the premise of dMR. The premise dMR asserts that the system AA (as it exists at time t1t_{1}) can be ascribed a hidden variable λθ\lambda_{\theta}, the value of which gives the outcome of the macroscopic spin SθAS_{\theta}^{A}, should that measurement be performed, because the eigenstates of SθAS_{\theta}^{A} are assumed macroscopically distinct. The premise dMR asserts that the value of λθ\lambda_{\theta} is not affected by measurements on spacelike separated systems. One may determine which of the two macroscopically distinct states (given by λθ=1\lambda_{\theta}=1 or 1-1) the system AA is in, by the measurements on BB and CC. Deterministic macroscopic realism asserts that the hidden variable λθ\lambda_{\theta} applies to the system AA, prior to the selection of the measurement settings at BB and CC. The set-up is as in Figure 5, where a switch controls whether SxAS_{x}^{A} or SyAS_{y}^{A} will be inferred at AA, by measuring either SxBSxCS_{x}^{B}S_{x}^{C}, or SxBSyCS_{x}^{B}S_{y}^{C}. The argument is that the measurement set-up at BB and CC does not disturb the outcome at AA, and hence both values, λxA\lambda_{x}^{A} and λyA\lambda_{y}^{A}, are simultaneously determined at AA, at the time t1t_{1}.

The GHZ paradox thus demonstrates that dMR will fail, assuming the paradox can be experimentally realised in agreement with quantum predictions. This is a strong result, giving an “all or nothing” contradiction with dMR. Other macroscopic versions of the GHZ paradox ghz-macro-1 ; ghz-macro-2 ; ghz-macro-3 ; mermin-inequality refer to multidimensional systems, and usually do not address the macroscopic distinction between the spin states. The falsification of MLR and dMR undermines the macroscopic EPR-Bohm argument for the incompleteness of quantum mechanics, given in Section V, which is based on the assumption that these premises are valid.

VIII.2 Consistency of GHZ quantum predictions with wMR and wLR

The conclusion that macroscopic realism does not hold would be a startling one. This motivates consideration of the less restrictive definition of weak macroscopic realism (wMR), defined in Section II.B. The premise wMR can be applied to the GHZ set-up to show there is no inconsistency of wMR with the quantum predictions. This is in agreement with previous work manushan-bell-cat-lg ; delayed-choice-cats , where consistency with wMR was shown for Bell violations using cat states.

To demonstrate the consistency with wMR, we consider state (LABEL:eq:ghz-cat) at time t1t_{1}, and then suppose the systems BB and CC are prepared so that pointer measurements of SxBS_{x}^{B} and SxCS_{x}^{C}, at the time tft_{f} will yield the outcomes of SxBS_{x}^{B} and SxCS_{x}^{C} (as in Figure 6). Weak macroscopic realism asserts that the systems are each assigned a predetermined value λxB\lambda_{x}^{B} and λxC\lambda_{x}^{C} respectively for the outcomes of those pointer measurements, at the time tft_{f}. The premise of wMR also assigns an inferred value

λxAλx,infA=λxBλxC\lambda_{x}^{A}\equiv\lambda_{x,inf}^{A}=\lambda_{x}^{B}\lambda_{x}^{C} (39)

to the system AA, since the values λxB\lambda_{x}^{B} and λxC\lambda_{x}^{C} enable a prediction with certainty for the outcome of the measurement SxAS_{x}^{A}, if performed at AA.

Refer to caption
Figure 6: Weak macroscopic realism (wMR) applied to the GHZ experiment. The premise wMR asserts validity of hidden variables for systems at time tft_{f} prepared for the pointer measurements. Sketched is the set-up where SyS_{y} is measured at AA, and SxS_{x} at BB and SxS_{x} at CC. The wMR premise asserts hidden variables for the system AA at the time tft_{f} that predetermine the final outcome of SyS_{y} at AA, and also predetermine the outcome of SxS_{x} at AA. This is because the prediction for SxS_{x} at AA can be given with certainty by the pointer measurements at BB and CC. Similar logic implies hidden variables that predetermine the outcomes for both SxS_{x} and SyS_{y} for sites BB and CC. However, there is no contradiction with wMR. This is because the hidden variables λxA\lambda_{x}^{A}, λyB\lambda_{y}^{B} and λyC\lambda_{y}^{C} give the outcomes of pointer measurements to be made after a further local unitary interaction UU, assuming there are no further unitary interactions at the other sites.

It is also the case however that wMR applies directly to AA. If the system AA undergoes rotation UyU_{y}, as depicted in Figure 6, then it is prepared in a pointer superposition with respect to SyAS_{y}^{A}. Hence the system AA is ascribed a hidden variable λyA\lambda_{y}^{A} to predetermine the outcome SyAS_{y}^{A} based on the pointer preparation of the system A itself.

At first glance, this seems to suggest a GHZ contradiction for wMR. Suppose one prepares the systems BB and CC for the pointer measurements of SxBS_{x}^{B} and SxCS_{x}^{C}, at time tft_{f} (Figure 6). Hence, for the systems BB and CC at time tft_{f}, the outcomes for SxAS_{x}^{A}, SxBS_{x}^{B} and SxCS_{x}^{C} are all predetermined, and given by variables λxAλx,infA\lambda_{x}^{A}\equiv\lambda_{x,inf}^{A}, λxB\lambda_{x}^{B} and λxC\lambda_{x}^{C} respectively. Additionally, one can prepare system AA in pointer measurement for yy, and the outcome for SyAS_{y}^{A} is also determined (Figure 6). Then, one can infer the values for the outcomes of measurements SyBS_{y}^{B} and SyCS_{y}^{C}, should they be performed by carrying out the appropriate unitary interaction at BB and CC. We have for the inferred values:

λx,infA\displaystyle\lambda_{x,inf}^{A} =\displaystyle= λxBλxC\displaystyle-\lambda_{x}^{B}\lambda_{x}^{C}
λy,infC\displaystyle\lambda_{y,inf}^{C} =\displaystyle= λxBλyA\displaystyle\lambda_{x}^{B}\lambda_{y}^{A}
λy,infB\displaystyle\lambda_{y,inf}^{B} =\displaystyle= λxCλyA.\displaystyle\lambda_{x}^{C}\lambda_{y}^{A}. (40)

For each system, the value of either SxS_{x} or SyS_{y} is determined (by the pointer preparation), and the value of the other measurement is determined, by inference of the other (pointer) values. Hence, it appears that there is the GHZ contradiction, because it is as though the outcomes of both SxS_{x} and SyS_{y} are determined at each site (at the same time), and these outcomes are either +1+1 or 1-1, hence creating the contradiction of Eq. (14).

However, there is no contradiction with wMR. The value for either SxS_{x} or SyS_{y} (the one that is inferred at each site) will require a local unitary rotation UU (a change of measurement setting) before the final read-out given by a pointer measurement. The unitary interaction UU occurs over a time interval. The unitary rotation means that the value λ\lambda that predetermines the outcome of the pointer measurement at time tft_{f} no longer (necessarily) applies at the later time, tmt_{m}, after the interaction UU. The system at tmt_{m} is prepared with respect to a different pointer measurement. Hence, at time tmt_{m}, the earlier predictions of the inferred values λinf\lambda_{inf} for the other sites no longer apply. The paradox as arising from Eq. (14) assumes all values of λ\lambda apply, to the state at time tft_{f}.

Refer to caption
Figure 7: Tracking the hidden variables as given by the premise of weak macroscopic realism (wMR). The dynamics is entirely consistent with the predictions of wMR, despite there being a GHZ paradox. These variables λ\lambda give the outcomes for the appropriate measurement if performed. The variables in bold red are implied by Assertion (1b) of wMR, that the system is prepared at that time tit_{i} at that site, with respect to the pointer basis. The variables indicated by the subscript infinf are deduced by Assertion (1a), that the outcome for the measurement can be predicted by the pointer measurements at the other sites, at that time tit_{i}. The hidden variables that are implied by wMR at the times tit_{i} are indicated beside the dashed vertical line labelled tit_{i}.

In Figure 7, we give more details of the way in which the hidden variables implied by wMR can be tracked and found consistent with the predictions of quantum mechanics. Suppose the system is prepared ready for pointer measurements SyAS_{y}^{A}, SxBS_{x}^{B} and SxCS_{x}^{C} at the time tkt_{k}, and the hidden variables λyA\lambda_{y}^{A}, λxB\lambda_{x}^{B} and λxC\lambda_{x}^{C} (in bold red) determine those pointer outcomes. The decision is then made to measure instead SyAS_{y}^{A}, SyBS_{y}^{B} and SxCS_{x}^{C}. This requires a further unitary rotation 𝐔yB=UyUx1\mathbf{U}_{y}^{B}=U_{y}U_{x}^{-1} at site BB. At time tmt_{m}, after 𝐔yB\mathbf{U}_{y}^{B} has taken place, the system is described by a different set of hidden variables, λyA\lambda_{y}^{A}, λyB\lambda_{y}^{B} and λxC\lambda_{x}^{C}. The outcome of the measurement of SyS_{y} is however determined with certainty by the pointer measurements for AA and CC, as λyB=λy,infB=λyAλxC\lambda_{y}^{B}=\lambda_{y,inf}^{B}=\lambda_{y}^{A}\lambda_{x}^{C}. At time tmt_{m}, we then see that system BB is no longer prepared in a pointer state for SxS_{x}. Hence, at time tmt_{m}, the earlier value of the inferred result λx,infA\lambda_{x,inf}^{A} at AA (which depended on λxB\lambda_{x}^{B}) is not relevant. A further unitary rotation 𝐔xA=UxUy1\mathbf{U}_{x}^{A}=U_{x}U_{y}^{-1} at AA that prepares the system AA for a final pointer measurement SxS_{x} will not (necessarily) give the results that applied at time tkt_{k} (which was prior to the UyU_{y} at BB taking place). Consider the hidden variables that are defined (based on the premise of wMR) at this time, t4t_{4}, after the evolution 𝐔xA\mathbf{U}_{x}^{A}. At the time t4t_{4}, the system is ascribed the variables λxA\lambda_{x}^{A}, λyB\lambda_{y}^{B} and λxC\lambda_{x}^{C}, with λyA\lambda_{y}^{A} also determined, for a future single unitary transformation at AA. The outcome for SyS_{y} at AA is predetermined (by the pointer outcomes at BB and CC) according to wMR, given by

λy,inf,4A\displaystyle\lambda_{y,inf,4}^{A} =\displaystyle= λyBλxC=λy,infBλxC\displaystyle\lambda_{y}^{B}\lambda_{x}^{C}=\lambda_{y,inf}^{B}\lambda_{x}^{C} (41)
=\displaystyle= λxCλyAλxC\displaystyle\lambda_{x}^{C}\lambda_{y}^{A}\lambda_{x}^{C}
=\displaystyle= λyA,\displaystyle\lambda_{y}^{A},

which gives consistency with the earlier value at tkt_{k}. However, the outcome for SxS_{x} at BB is inferred from the pointer values at time t4t_{4}:

λx,inf,4B\displaystyle\lambda_{x,inf,4}^{B} =\displaystyle= λxAλxC.\displaystyle-\lambda_{x}^{A}\lambda_{x}^{C}. (42)

For consistency with the values defined at tkt_{k}, we could propose λxA=λx,infA=λxBλxC\lambda_{x}^{A}=\lambda_{x,inf}^{A}=-\lambda_{x}^{B}\lambda_{x}^{C}, in which case we would obtain λx,inf,4B=λxAλxC=λxB(λxC)2=λxB\lambda_{x,inf,4}^{B}=-\lambda_{x}^{A}\lambda_{x}^{C}=\lambda_{x}^{B}(\lambda_{x}^{C})^{2}=\lambda_{x}^{B}, giving an apparent consistency with the earlier value. However, the value of SyS_{y} at CC at time t4t_{4} is inferred to be

λy,inf,4C\displaystyle\lambda_{y,inf,4}^{C} =\displaystyle= λyBλxA=λy,infBλxA\displaystyle\lambda_{y}^{B}\lambda_{x}^{A}=\lambda_{y,inf}^{B}\lambda_{x}^{A} (43)
=\displaystyle= (λxCλyA)λxA.\displaystyle(\lambda_{x}^{C}\lambda_{y}^{A})\lambda_{x}^{A}.

Now if we propose λxA=λx,infA=λxBλxC\lambda_{x}^{A}=\lambda_{x,inf}^{A}=-\lambda_{x}^{B}\lambda_{x}^{C}, we obtain λy,inf,4C=(λxCλyA)(λxBλxC)=λyAλxB\lambda_{y,inf,4}^{C}=(\lambda_{x}^{C}\lambda_{y}^{A})(-\lambda_{x}^{B}\lambda_{x}^{C})=-\lambda_{y}^{A}\lambda_{x}^{B}. We see here that this is different to the earlier value, λy,infC=λyAλxB\lambda_{y,inf}^{C}=\lambda_{y}^{A}\lambda_{x}^{B}. Hence, it is not possible to gain consistency between wMR and the values λ\lambda asserted by the premise of dMR. While dMR is falsified by the GHZ paradox, we see that the GHZ contradiction does not apply to wMR.

We note that according to wMR, the value λyA\lambda_{y}^{A} for system AA prepared for the pointer measurement SyAS_{y}^{A}, for example, is not changed by unitary rotations that may take place at BB or CC (Figure 7, at time tmt_{m}). However, if there is a further unitary rotation at AA, and also at BB (i.e. two unitary rotations, at different sites) the value λx,infA\lambda_{x,inf}^{A} for SxAS_{x}^{A} can change (Figure 7, at time t4t_{4}).

IX Consistency of weak local realism with Bell violations

It has been shown possible to falsify deterministic macroscopic realism for the cat-state system described in Section V.A manushan-bell-cat-lg . This was demonstrated by a violation of Bell inequalities constructed for the macroscopic spins, |α|\alpha\rangle and |α|-\alpha\rangle. Similarly, in Ref. manushan-bell-cat-lg , it was shown that wMR was not falsified by the Bell violations. This proof was expanded in wigner-friend-macro . Essentially, the same proof holds to show consistency of wLR with Bell violations wigner-friend-macro . For the sake of completeness, we present below the proof demonstrating consistency of wLR with Bell violations. This is relevant, since we note that an identical proof holds to show consistency of wMR with macroscopic Bell violations, where the spins are realised by the macroscopically distinct states |J|JN|\uparrow\rangle_{J}\equiv|\uparrow\rangle_{J}^{\otimes N} and |z,J|z,JN|\downarrow\rangle_{z,J}\equiv|\downarrow\rangle_{z,J}^{\otimes N} and the unitary operations of the analyzer are realised by the CNOT gates as proposed in Section V.B.

The Bell test involves the EPR-Bohm system. The Pauli spin components defined as

S^θA\displaystyle\hat{S}_{\theta}^{A} =\displaystyle= S^xAsinθ+S^zAcosθ\displaystyle\hat{S}_{x}^{A}\sin\theta+\hat{S}_{z}^{A}\cos\theta
S^ϕB\displaystyle\hat{S}_{\phi}^{B} =\displaystyle= S^xBsinϕ+S^zBcosϕ\displaystyle\hat{S}_{x}^{B}\sin\phi+\hat{S}_{z}^{B}\cos\phi (44)

can be measured by adjusting the analyzer (Stern-Gerlach apparatus or polarizing beam splitter) and the expectation value given by E(θ,ϕ)=S^θAS^ϕBE(\theta,\phi)=\langle\hat{S}_{\theta}^{A}\hat{S}_{\phi}^{B}\rangle measured. According to the EPR-Bohm argument based on EPR’s local realism, each spin component S^θA\hat{S}_{\theta}^{A} and S^ϕB\hat{S}_{\phi}^{B} is represented by a hidden variable (λθA\lambda_{\theta}^{A} and λϕB\lambda_{\phi}^{B}), because the value can be predicted with certainty by a spacelike-separated measurement bell-1971 . This leads to the constraint 2S2-2\leq S\leq 2 where S=E(θ,ϕ)E(θ,ϕ)+E(θ,ϕ)+E(θ,ϕ)S=E(\theta,\phi)-E(\theta,\phi^{\prime})+E(\theta^{\prime},\phi)+E(\theta^{\prime},\phi^{\prime}), known as the Clauser-Horne-Shimony-Holt-(CHSH) Bell inequality, which is violated for the Bell state (4) bell-1971 ; chsh ; bell-cs-review . The violation therefore falsifies EPR’s premises based on local realism. More generally, the violation shows failure of all local realistic theories defined as those satisfying Bell’s local realism assumptions bell-1971 ; chsh .

Refer to caption
Refer to caption
Figure 8: Tracking the hidden variables according to weak local realism (wLR) through the dynamics of the Bell test, which shows violation of the Bell inequality. First, the moment SθASϕB\langle S_{\theta}^{A}S_{\phi}^{B}\rangle is measured. We take the initial time t1t_{1} to be after the passage through the analysers set at θ\theta and ϕ\phi, so that measurement settings θ\theta and ϕ\phi have been determined and the system prepared for the final pointer measurements SθAS_{\theta}^{A} and SϕBS_{\phi}^{B}. At each time tit_{i}, wLR implies that certain hidden variables λ\lambda are valid, depending on the preparation of the system at that time. The hidden variables are depicted in the brackets. Those in red bold are implied by wLR Assertion 1b. Those in black (and not bold) are implied by Assertion 1a.

In the Figure 8, we track the hidden variables that predetermine the values of the spin measurements at each time, based on wLR. We illustrate without loss of generality with one possible time sequence, based on the preparation at the initial time for pointer measurements in the directions θ\theta and ϕ\phi. Assuming wLR, the values of SθAS_{\theta}^{A} and SϕBS_{\phi}^{B} that are realised by the pointer stage of measurement (if made at that time) are predetermined, given by λθA\lambda_{\theta}^{A} and λϕB\lambda_{\phi}^{B} at the time t1t_{1}. Hence

E(θ,ϕ)=λθAλϕB.E(\theta,\phi)=\langle\lambda_{\theta}^{A}\lambda_{\phi}^{B}\rangle. (45)

To measure E(θ,ϕ)E(\theta^{\prime},\phi), there is a further rotation UθAU_{\theta^{\prime}}^{A} at AA. At time t2t_{2}, the state is prepared for the pointer measurements of SθAS_{\theta^{\prime}}^{A} and SϕBS_{\phi}^{B}. The hidden variables λθA\lambda_{\theta^{\prime}}^{A} and λϕB\lambda_{\phi}^{B} specify the outcomes for those pointer measurements should they be performed. Based on the premise wLR Assertion (1b), these variables are assigned to describe the state of the system at the time t2t_{2}. We note that also because of the anticorrelation evident for spins prepared in the Bell state (4), the wLR Assertion (1a) implies that the hidden variable λϕB\lambda_{\phi}^{B} also specifies the outcome of a measurement SϕAS_{\phi}^{A}, if performed on the system defined at time t2t_{2}. The prediction for wLR is

E(θ,ϕ)=λθAλϕB.E(\theta^{\prime},\phi)=\langle\lambda_{\theta^{\prime}}^{A}\lambda_{\phi}^{B}\rangle. (46)

Similarly, the measurements of SθAS_{\theta}^{A} and SϕBS_{\phi^{\prime}}^{B} require a further rotation UϕBU_{\phi^{\prime}}^{B} at BB, after the initial preparation at t1t_{1}, with no rotation at AA. A variable λϕB\lambda_{\phi^{\prime}}^{B} is defined to give the outcome for SϕBS_{\phi^{\prime}}^{B}, if that measurement were to be performed at t2t_{2} after the rotation UϕBU_{\phi^{\prime}}^{B}. Hence, wLR implies

E(θ,ϕ)=λθAλϕB.E(\theta,\phi^{\prime})=\langle\lambda_{\theta}^{A}\lambda_{\phi^{\prime}}^{B}\rangle. (47)
Refer to caption
Figure 9: Tracking the hidden variables through the dynamics of the Bell test, which shows violation of the Bell inequality. The description is as for Figure 8, but here the moment SθASϕB\langle S_{\theta^{\prime}}^{A}S_{\phi^{\prime}}^{B}\rangle is measured. We note that the conditioning for λϕ\lambda_{\phi^{\prime}} necessitates that a rotation UϕBU_{\phi^{\prime}}^{B} has also occurred at BB, as well as UθAU_{\theta^{\prime}}^{A}. The nonlocality emerges only after rotations at both sites.

The difference between Bell’s local hidden variable theories and the assertions of wLR are evident when considering measurement of SθAS_{\theta^{\prime}}^{A} and SϕBS_{\phi^{\prime}}^{B}. This measurement requires two further rotations after the preparation at time t1t_{1}. A possible sequence is given in Figure 9. Suppose the rotation UθAU_{\theta^{\prime}}^{A} is performed first, and at time t2t_{2}, the hidden variables defining the state are λθA\lambda_{\theta^{\prime}}^{A} and λϕB\lambda_{\phi}^{B}. The pointer measurements are not actually performed, and a rotation UϕBU_{\phi^{\prime}}^{B} is then made at BB. The final state at time t3t_{3} is given by hidden variables λθA\lambda_{\theta^{\prime}}^{A} and λϕ|θB\lambda_{\phi^{\prime}|\theta^{\prime}}^{B}. Here, we use subscripts |θ|\theta^{\prime} to specify that λϕ|θB\lambda_{\phi^{\prime}|\theta^{\prime}}^{B} is the variable defined for the state specified at the time t3t_{3}, conditioned on the rotation UθAU_{\theta^{\prime}}^{A} at AA. This is necessary in the context of a wLR model, because the premise of wLR specifies the necessity of locality for the hidden values of the pointer measurements only. The value λθA\lambda_{\theta^{\prime}}^{A} is defined for the pointer measurement of SθAS_{\theta^{\prime}}^{A}, and is independent of the choice for ϕ\phi^{\prime}, since this value λθA\lambda_{\theta^{\prime}}^{A} is (according to wLR) not affected by the unitary rotation UϕBU_{\phi^{\prime}}^{B} at the other site BB. However, we cannot conclude from the wLR assertions that the value of λϕB\lambda_{\phi^{\prime}}^{B} defined for the measurement SϕBS_{\phi^{\prime}}^{B} on the state after the rotation UθAU_{\theta^{\prime}}^{A} is the same as that defined for pointer measurement SϕBS_{\phi^{\prime}}^{B} in Figure 8, where there was no rotation at AA. Hence we write

E(θ,ϕ)=λθAλϕ|θB.E(\theta^{\prime},\phi^{\prime})=\langle\lambda_{\theta^{\prime}}^{A}\lambda_{\phi^{\prime}|\theta^{\prime}}^{B}\rangle. (48)

We see that wLR does not imply the CHSH-Bell inequality, which is derived based on the full Bell locality assumption that λϕB\lambda_{\phi^{\prime}}^{B} is independent of the value of θ\theta^{\prime} i.e. is independent of whether the rotation UθAU_{\theta^{\prime}}^{A} has been performed at AA or not.

It is well known that the where the values for λθ\lambda_{\theta}, λϕ\lambda_{\phi}, λθ\lambda_{\theta^{\prime}}, and λϕ\lambda_{\phi^{\prime}} are either +1+1 or 1-1, and if Bell locality is assumed so that λϕ|θB=λϕB\lambda_{\phi^{\prime}|\theta}^{B}=\lambda_{\phi^{\prime}}^{B}, then the value of SS is bounded by 2-2 and 22, leading to the CHSH-Bell inequality bell-cs-review . However, where we consider λϕ|θB\lambda_{\phi^{\prime}|\theta}^{B} to be an independent variable, +1+1 or 1-1, the bound for SS becomes the algebraic bound of 44. Hence, wLR does not constrain SS to be bounded by the Bell inequality.

Nonlocality and deeper models

The wLR and wMR premises allow for nonlocal effects, as evident by the violation of the Bell inequality. This is the meaning of “weak”, that the premises do not encompass the full local realism assumptions of EPR.

The nonlocal effect arises in the above analysis because it cannot be assumed that the value λϕB\lambda_{\phi^{\prime}}^{B} is independent of the value θ\theta^{\prime}, which determines the unitary rotation at AA. The λθA\lambda_{\theta^{\prime}}^{A} is independent of ϕ\phi^{\prime}, because the setting θ\theta^{\prime} is fixed (the unitary rotation UθAU_{\theta^{\prime}}^{A} has occurred before UϕBU_{\phi^{\prime}}^{B} at AA), but it cannot be excluded that the λϕB\lambda_{\phi^{\prime}}^{B} can depend on θ\theta^{\prime} because it occurred earlier. It seems to matter which order the unitary rotations are taken, despite that the two rotations occur at spatially separated locations. The predictions will not however depend on the order of rotation. The joint distribution for values λθA\lambda_{\theta^{\prime}}^{A}, λϕB\lambda_{\phi^{\prime}}^{B} depends on both θ\theta^{\prime} and ϕ\phi^{\prime}, the final settings. We note that the conditioning for λϕ\lambda_{\phi^{\prime}} necessitates that a rotation UϕBU_{\phi^{\prime}}^{B} has also occurred at BB, since time t1t_{1}. Rotations at both sites are required for the nonlocality to emerge. This feature of wMR and wLR is proved for the GHZ set-up in Section X. We make two comments.

First, the wLR and wMR premises remove the possibility of a strong sort of nonlocality. In these models, the choice to measure ϕ\phi^{\prime} instead of ϕ\phi at BB does not change the value of λθA\lambda_{\theta}^{A} at AA once the unitary rotation has occurred at AA to fix the measurement setting as θ\theta at AA. There is hence no instantaneous nonlocal effect. Similarly, for the system prepared in the Bell state, the value for SϕAS_{\phi}^{A} at AA is fixed once the rotation UϕBU_{\phi}^{B} has occurred at BB, because the value for SϕAS_{\phi}^{A} can be predicted with certainty, even when the unitary rotation UϕAU_{\phi}^{A} at AA has not occurred. While this gives a nonlocal effect, a further local interaction UϕAU_{\phi}^{A} is required at AA for the nonlocality to be confirmed.

Second, there is motivation to examine deeper models and tests of quantum mechanics q-contextual-1 ; bohm-hv ; Maroney ; maroney-timpson ; hall-cworlds ; griffiths-histories ; grangier-auffeves-context ; brukner-wigner ; spekkens-toy-model ; q-measurement ; simon-q ; objective-fields-entropy-1 ; castagnoli-2021 ; sabine-retro-toy ; griffiths-nonlocal-not ; roman-schn ; philippe-grang-context-1 ; fr-paradox ; wigner-weak ; bohmian-fr ; losada-wigner-friend for consistency with wLR and wMR. Maroney and Timpson proposed models for macroscopic realism that allow violation of Leggett-Garg inequalities Maroney ; maroney-timpson . In their “supra eigenstate support MR model”, for which there is a predetermination of the outcome of the measurement that distinguishes between two macroscopic distinct states, it is explained that the “state” of the system cannot be an operational eigenstate, meaning it cannot be a preparable state for the system. It is clear from their context that the system is considered to be prepared for a pointer measurement, an example being the observation of a ball in a box. This would give consistency with wMR. The authors gave the de Broglie-Bohm theory as an example of such a model. The de Broglie-Bohm theory is a nonlocal theory for quantum mechanics bohm-hv .

Another model of quantum mechanics that appears consistent with wMR is the objective field theory, motivated by solutions from quantum field theory and the Q function q-contextual-1 ; q-measurement ; simon-q ; objective-fields-entropy-1 . Solutions have been given where the second stage of a measurement is modeled dynamically as amplification of field amplitudes. Here, there is no direct nonlocal mechanism, but rather a retrocausality based on future boundary conditions, which leads to hidden causal loops castagnoli-2021 . The joint distribution for values λθA\lambda_{\theta^{\prime}}^{A}, λϕB\lambda_{\phi^{\prime}}^{B} is shown to depend on θ\theta^{\prime} and ϕ\phi^{\prime}, the final settings, and Bell’s local hidden variable model does not apply. In recent work, it is reported how, for EPR and Bell correlations based on continuous-variable measurements, the premise of wMR is upheld q-measurement . The premise of wMR does not allow retrocausality at a macroscopic level because the hidden variable λ\lambda is fixed at the given time, being independent of any future event.

X Further predictions for wMR/ wLR

We present further predictions for wMR. These provide a means to experimentally test wMR. The predictions are identical to those of quantum mechanics. However, wMR differs from standard quantum mechanics which does not account for predetermined values of a measurement. The analyses apply in identical fashion to wLR.

X.1 Moments involving a further single rotation are consistent with local realism

Prediction of wMR: We consider the entangled cat-state GHZ system |ψ|\psi\rangle (Eq. (LABEL:eq:ghz-cat)) which is then prepared at time tkt_{k} for pointer measurements SyAS_{y}^{A}, SxBS_{x}^{B} and SxCS_{x}^{C} at the respective sites (as in Figure 10). The GHZ contradiction with local realism is realised by first further changing the measurement settings, to measure SxAS_{x}^{A}, SxBS_{x}^{B} and SxCS_{x}^{C}, which involves one unitary rotation 𝐔xA\mathbf{U}_{x}^{A}. Also required are measurements SxAS_{x}^{A}, SyBS_{y}^{B} and SyCS_{y}^{C}, which involve two further rotations, one at each site BB and CC, as well as 𝐔xA\mathbf{U}_{x}^{A} (Figure 11). The prediction is that results violating local realism do not arise from the correlations involving only one unitary UxAU_{x}^{A} after the preparation at tkt_{k}. The violations arise from the correlations involving the two further rotations. A similar result was proved for Bell violations manushan-bell-cat-lg .

Proof: We denote the state prepared at the time tkt_{k} by |ψy,x,x|\psi\rangle_{y,x,x}. A final pointer measurement if conducted at time tkt_{k} at AA gives the outcome for SyAS_{y}^{A}. According to wMR, at time tkt_{k}, the values for SxAS_{x}^{A}, SyAS_{y}^{A}, SxBS_{x}^{B} and SxCS_{x}^{C} are each predetermined, being given by the variables λinf,xA\lambda_{inf,x}^{A}, λyA\lambda_{y}^{A}, λxB\lambda_{x}^{B} and λxC\lambda_{x}^{C}. A single unitary rotation 𝐔xA=Uy1Ux\mathbf{U}_{x}^{A}=U_{y}^{-1}U_{x} at AA will enable the result λinf,xA\lambda_{inf,x}^{A} to be revealed. According to wMR, the prediction for SxASxBSxCS_{x}^{A}S_{x}^{B}S_{x}^{C} is predetermined at time tkt_{k}, for the system prepared in the state |ψy,x,x|\psi\rangle_{y,x,x}. There are hidden variables for each measurement, defined for the system at tkt_{k}. Therefore, if wMR is valid, the prediction for SxASxBSxCS_{x}^{A}S_{x}^{B}S_{x}^{C}, conditioned on the initial state |ψy,x,x|\psi\rangle_{y,x,x}, must be entirely consistent with local realism.\square

Refer to caption
Figure 10: Predictions of weak macroscopic realism (wMR) are consistent with local realism for the single rotation 𝐔𝐱𝐀\mathbf{U_{x}^{A}} after preparation at time tkt_{k}. The notation is as for Figure 7. The system at time tkt_{k} is prepared such that pointer measurements will yield outcomes for measurements of SyAS_{y}^{A}, SxBS_{x}^{B} and SxCS_{x}^{C}. The prediction for SxAS_{x}^{A} can be inferred from the pointer measurements at BB and CC and hence is also predetermined at the time tkt_{k}, according to wMR. This measurement requires a further unitary interaction UxAU_{x}^{A}. Final pointer measurements at the time tmt_{m} will yield the result for measurement SxASxBSxCS_{x}^{A}S_{x}^{B}S_{x}^{C}, which by wMR is consistent with local realism. Quantum mechanics also predicts consistency with local realism for the single rotation (see text).

Proof of agreement with quantum prediction: Here, we prove that the prediction of quantum mechanics also shows consistency with local realism for the set-up of just one rotation after preparation (Figure 10). To do this, we compare the predictions of quantum mechanics for the prepared state |ψy,x,x|\psi\rangle_{y,x,x} against those of a mixed state ρmixABC\rho_{mix}^{A-BC}. The state |ψy,x,x|\psi\rangle_{y,x,x} prepared at time tkt_{k} in the basis for SyAS_{y}^{A}, SxBS_{x}^{B} and SxBS_{x}^{B} is

|ψy,x,x\displaystyle|\psi\rangle_{y,x,x} =\displaystyle= 14{(|y+|y)(|x+|x)(|x+|x)\displaystyle\frac{1}{4}\{(|\uparrow\rangle_{y}+|\downarrow\rangle_{y})(|\uparrow\rangle_{x}+|\downarrow\rangle_{x})(|\uparrow\rangle_{x}+|\downarrow\rangle_{x}) (49)
+i(|y|y)(|x|x)(|x|x)}\displaystyle+i(|\uparrow\rangle_{y}-|\downarrow\rangle_{y})(|\uparrow\rangle_{x}-|\downarrow\rangle_{x})(|\uparrow\rangle_{x}-|\downarrow\rangle_{x})\}
=\displaystyle= 12(|ψA|ψ+BC+|ψ+A|ψBC)\displaystyle\frac{1}{\sqrt{2}}(|\psi_{-}^{A}\rangle|\psi_{+}^{BC}\rangle+|\psi_{+}^{A}\rangle|\psi_{-}^{BC}\rangle)

where

|ψ+BC\displaystyle|\psi_{+}^{BC}\rangle =\displaystyle= (|x|x+|x|x)/2\displaystyle(|\uparrow\rangle_{x}|\uparrow\rangle_{x}+|\downarrow\rangle_{x}|\downarrow\rangle_{x})/\sqrt{2}
|ψBC\displaystyle|\psi_{-}^{BC}\rangle =\displaystyle= (|x|x+|x|x)/2\displaystyle(|\downarrow\rangle_{x}|\uparrow\rangle_{x}+|\uparrow\rangle_{x}|\downarrow\rangle_{x})/\sqrt{2}

and

|ψ+A\displaystyle|\psi_{+}^{A}\rangle =\displaystyle= {(1i)|y+(1+i)|y}/2\displaystyle\{(1-i)|\uparrow\rangle_{y}+(1+i)|\downarrow\rangle_{y}\}/2
|ψA\displaystyle|\psi_{-}^{A}\rangle =\displaystyle= {(1+i)|y+(1i)|y}/2.\displaystyle\{(1+i)|\uparrow\rangle_{y}+(1-i)|\downarrow\rangle_{y}\}/2.

The state |ψy,x,x|\psi\rangle_{y,x,x} is a superposition involving entanglement between the system AA and the system denoted BCBC, which comprises the systems BB and CC. If the unitary rotation 𝐔xA\mathbf{U}_{x}^{A} is performed at AA, then the prediction for the pointer measurements is SxASxBSxC=1S_{x}^{A}S_{x}^{B}S_{x}^{C}=-1. Now we compare with the system initially prepared in the mixture

ρmixABC=|ψAψA|ρ+BC+|ψ+Aψ+A|ρBC.\rho_{mix}^{A-BC}=|\psi_{-}^{A}\rangle\langle\psi_{-}^{A}|\rho_{+}^{BC}+|\psi_{+}^{A}\rangle\langle\psi_{+}^{A}|\rho_{-}^{BC}. (50)

Here, ρ+BC=|ψ+BCψ+BC|\rho_{+}^{BC}=|\psi_{+}^{BC}\rangle\langle\psi_{+}^{BC}| and ρBC=|ψBCψBC|\rho_{-}^{BC}=|\psi_{-}^{BC}\rangle\langle\psi_{-}^{BC}|. This mixture has no entanglement between the system AA and the combined systems BB and CC i.e. it is fully separable with respect to the bipartition that we denote by ABCA-BC. If we transform to the xx basis at AA, then we write

ρmixABC=|x|ρ+BCx+|x|ρBCx.\rho_{mix}^{A-BC}=|\downarrow\rangle_{x}\langle\downarrow|{}_{x}\rho_{+}^{BC}+|\uparrow\rangle_{x}\langle\uparrow|{}_{x}\rho_{-}^{BC}. (51)

The prediction is SxASxBSxC=1S_{x}^{A}S_{x}^{B}S_{x}^{C}=-1, which is identical to the prediction for the system prepared in |ψy,x,x|\psi\rangle_{y,x,x}. The quantum prediction for the single unitary interaction 𝐔xA\mathbf{U}_{x}^{A} on |ψy,x,x|\psi\rangle_{y,x,x} is therefore consistent with local realism - since the prediction for ρmixABC\rho_{mix}^{A-BC} is fully local with respect to AA, arising from a local interaction at AA.\square

The GHZ test showing violation of local realism indeed requires two further rotations, 𝐔yB\mathbf{U}_{y}^{B} and 𝐔yC\mathbf{U}_{y}^{C} at the sites BB and CC, which allows measurement of SxASyBSyCS_{x}^{A}S_{y}^{B}S_{y}^{C} (Figure 11). This is because wMR does not (necessarily) predict for the system at time tkt_{k}, a predetermination of the outcomes for both SxAS_{x}^{A} and SyBS_{y}^{B}. Hence, there is no contradiction between the predictions of quantum mechanics and wMR. The hidden variables that are predicted by wMR are tracked in the Figure 11. An experiment could be performed, by comparing the observed moments for the GHZ state with those generated by the mixed states.

Refer to caption
Figure 11: Predictions of weak macroscopic realism (wMR) are consistent with those of quantum mechanics for multiple rotations. The notation is as for Figure 7. Here, further unitary interactions 𝐔yB\mathbf{U}_{y}^{B} and 𝐔yC\mathbf{U}_{y}^{C} prepare the system for measurement of SxASyBSyCS_{x}^{A}S_{y}^{B}S_{y}^{C} at the time t5t_{5}. However, at the time tmt_{m}, there is no longer the pointer preparation for SyAS_{y}^{A}, and the values that were inferred for measurements SyBS_{y}^{B} and SyCS_{y}^{C} no longer apply. The predictions lead to the GHZ paradox, but are consistent with wMR. The premise wMR also predicts that the results for SxASyBSyCS_{x}^{A}S_{y}^{B}S_{y}^{C} are independent of when the final irreversible stage of the pointer measurement for SxAS_{x}^{A} is performed, relative to the unitary transformations at BB and CC. The different timings are indicated by the PP symbol.

X.2 The timing of the pointer stage of measurement

Prediction of wMR: Consider the system of Figure 11, prepared at time tkt_{k} so that pointer measurements at AA, BB and CC will give the outcomes for SyAS_{y}^{A}, SxBS_{x}^{B} and SxCS_{x}^{C}. At AA, the system is then prepared for a pointer measurement of SxS_{x}. At BB, a unitary rotation then prepares system BB for a pointer measurement of SyBS_{y}^{B}, and then similarly at CC. If wMR is valid, then the predictions for the correlations are not dependent on whether the final pointer stages PP of the measurement for SxAS_{x}^{A} at AA occur before or after the unitary rotations at BB and CC. Here, the final pointer stages of the measurement (denoted PP in the Figure) involve a coupling to an environment, whereby the measurement becomes irreversible.

Proof: The premise wMR asserts that the value λ\lambda for the outcome of the pointer measurement is fixed locally for the appropriately prepared system, provided there is no further unitary UU on that system which changes the measurement setting. This prediction agrees with that of quantum mechanics. Quantum calculations do not distinguish the timing of the measurement stage PP.

XI Conclusion

The main conclusion of this paper is that the negation of local realism, as evidenced by a Bell or GHZ experiment, does not appear to fully resolve the EPR paradox. In summary, we have proposed how EPR-Bohm and GHZ experiments may be realised in mesoscopic and macroscopic regimes, using cat states and suitable unitary interactions. These are significant tests in a setting where all relevant measurements are coarse-grained, distinguishing only between two macroscopically distinct states. The macroscopic EPR-Bohm test illustrates an incompatibility between the assumptions of deterministic macroscopic realism (dMR) and the notion that quantum mechanics is a complete description of physical reality. We explain it is also possible to consider the weaker assumption, weak macroscopic realism (wMR), and to demonstrate a similar inconsistency with the notion that quantum mechanics is a complete theory, using a two-spin version of the EPR-Bohm argument. Yet, while dMR can be falsified by the GHZ experiment, the predictions of the GHZ test agree with those of wMR. Similarly, there is no incompatibility between wMR and violations of macroscopic Bell inequalities. In defining wMR, it is necessary to consider that the measurement occurs in two stages, a reversible stage establishing the measurement setting, and an irreversible stage referred to as the pointer stage of measurement.

Similar conclusions can be drawn for the original EPR and GHZ paradoxes. This paper motivates consideration of a weaker assumption, weak local realism (wLR), in the setting of the original paradox. The EPR argument can be modified to show inconsistency between wLR and the notion that quantum mechanics is a complete description of physical reality. Yet, we show that the predictions of quantum mechanics for the GHZ and Bell experiments are consistent with those of wLR. The definitions of wMR and wLR apply to systems after the choice of measurement basis, and hence do not conflict with the contextuality of quantum mechanics kochen-spekker . Our work may be seen as a supplement to other arguments presented for the incompleteness of quantum mechanics bell-against ; bell-speakable ; s-cat-1935 ; harrigan-spekkens ; delayed-choice-cats ; philippe-grang-context-1 , and may motivate a study of alternative models for quantum mechanics.

In addition to the cat-state experiments, we propose further tests of wLR and wMR. These tests examine correlations after single unitary rotations, and adjust the timing of the unitary interactions that lead to the GHZ contradiction. The predictions of wMR and wLR agree with those of quantum mechanics. The EPR and GHZ paradoxes apply where one can predict with certainty the outcome of a measurement, given measurements at spacelike-separated sites. Experimental factors may prevent the realisation of predictions that are certain. The tests can nonetheless be carried out using inequalities mermin-inequality ; ardehli ; belinski-klyshko ; epr-rmp ; bohm-test-uncertainty ; epr-r2 . Proposals for realistic tests are given in the Appendix.

The proposed experiments could be realised in the microscopic regime using standard techniques where the unitary interactions are performed with polarizing beam splitters. Potential macroscopic realizations are given in Sections V and VII. The two-mode cat states involving coherent states have been generated in cavities cat-det-map ; cat-bell-wang-1 , and GHZ states generated for N20N\sim 20 omran-cats . Mesoscopic realisations of the unitary transformations are in principle feasible using dynamical interactions involving a nonlinear medium, or else CNOT gates.

Acknowledgements.
This work was funded through the Australian Research Council Discovery Project scheme under Grants DP180102470 and DP190101480. The authors thank NTT Research for their financial and technical support.

Appendix

XI.1 The unitary operation UyU_{y} for measurement of SyS_{y}

Consider the system AA originally in the eigenstate for Sy:S_{y}:

|y=eiπ/42(|z+i|z),|\uparrow\rangle_{y}=\frac{e^{-i\pi/4}}{\sqrt{2}}(|\uparrow\rangle_{z}+i|\downarrow\rangle_{z}), (52)

which is

|yeiπ/42(|αz+i|αz)|\uparrow\rangle_{y}\equiv\frac{e^{-i\pi/4}}{\sqrt{2}}(|\alpha\rangle_{z}+i|-\alpha\rangle_{z}) (53)

in our realisation. The state after the operation UyU_{y} is |αy|\alpha\rangle_{y}, since we see from (18) that

Uy|y\displaystyle U_{y}|\uparrow\rangle_{y} =\displaystyle= Uπ/41eiπ/42(|z+i|z)\displaystyle U_{\pi/4}^{-1}\frac{e^{-i\pi/4}}{\sqrt{2}}(|\uparrow\rangle_{z}+i|\downarrow\rangle_{z}) (54)
=\displaystyle= |α\displaystyle|\alpha\rangle

The pointer measurement S^\hat{S} on this state (for large α\alpha) gives +1+1, corresponding to the outcome required for the eigenstate |y|\uparrow\rangle_{y}. Similarly, consider the system prepared in |y|\downarrow\rangle_{y}

|y=eiπ/42(|z+i|z|\downarrow\rangle_{y}=\frac{e^{-i\pi/4}}{\sqrt{2}}(|\downarrow\rangle_{z}+i|\uparrow\rangle_{z} (55)

which is

|yeiπ/42(|αz+i|αz)|\downarrow\rangle_{y}\equiv\frac{e^{-i\pi/4}}{\sqrt{2}}(|-\alpha\rangle_{z}+i|\alpha\rangle_{z}) (56)

The state after the operation UyU_{y} is |αy|-\alpha\rangle_{y}, since from (18), we see that

Uy|y\displaystyle U_{y}|\downarrow\rangle_{y} =\displaystyle= |α\displaystyle|-\alpha\rangle (57)

for which the pointer measurement XX gives the outcome 1-1, as required for this eigenstate. Hence, the system that is originally in the linear superposition (23) transforms after UyU_{y} to

Uy|ψ\displaystyle U_{y}|\psi\rangle =\displaystyle= d+Uy|+y+dUy|y\displaystyle d_{+}U_{y}|+\rangle_{y}+d_{-}U_{y}|-\rangle_{y} (58)
\displaystyle\rightarrow d+|α+d|α\displaystyle d_{+}|\alpha\rangle+d_{-}|-\alpha\rangle

As α\alpha\rightarrow\infty, the probability of an outcome +1+1 (1-1) for the measurement S^\hat{S} of the sign of X^A\hat{X}_{A} is |d+|2|d_{+}|^{2} (|d|2|d_{-}|^{2}) respectively, as required.

XI.2 Example of mesoscopic qubits: NOON states

We may also consider where the macroscopic spins are two-mode number states |N|0|N\rangle|0\rangle and |0|N|0\rangle|N\rangle, for NN large. We denote two distinct modes by symbols ++ and -, and simplify the notation so that |N|0|N,0|N\rangle|0\rangle\equiv|N,0\rangle and |0|N|0,N|0\rangle|N\rangle\equiv|0,N\rangle. The macroscopic qubits become ||N,0|\uparrow\rangle\rightarrow|N,0\rangle and ||0,N|\downarrow\rangle\rightarrow|0,N\rangle. For the GHZ paradoxes, we will consider three sites, labelled AA, BB and CC. There are two modes (labelled J+J+ and JJ-) identified for each site JAJ\equiv A, BB, CC. The initial state would be of the form (LABEL:eq:ghz-cat). For each site, we use the transformation macro-bell-lg

(UyJ)1|N,0J\displaystyle(U_{y}^{J})^{-1}|N,0\rangle_{J} =\displaystyle= eiφ(cosθ|N,0J+isinθ|0,NJ)\displaystyle e^{i\varphi}(\cos\theta|N,0\rangle_{J}+i\sin\theta|0,N\rangle_{J})
(UyJ)1|0,NJ\displaystyle(U_{y}^{J})^{-1}|0,N\rangle_{J} =\displaystyle= ieiφ(sinθ|N,|0Jicosθ|0,NJ)\displaystyle ie^{i\varphi}(\sin\theta|N,|0\rangle_{J}-i\cos\theta|0,N\rangle_{J})

where |N,0J|N,0\rangle_{J} and |N,0J|N,0\rangle_{J} are the two-mode number states at site JJ, and φ\varphi is a phase-shift. The transformation has been realised to an excellent approximation for N100N\lesssim 100 macro-bell-lg , using the interaction josHam-collett-steel-2-1 ; nonlinear-Ham-1

HnlJ=κ(a^J+a^J+a^J+a^J)+ga^J+2a^J+2+ga^J2a^J2H_{nl}^{J}=\kappa(\hat{a}_{J+}^{\dagger}\hat{a}_{J-}+\hat{a}_{J+}\hat{a}_{J-}^{\dagger})+g\hat{a}_{J+}^{\dagger 2}\hat{a}_{J+}^{2}+g\hat{a}_{J-}^{\dagger 2}\hat{a}_{J-}^{2} (60)

so that UyJ=eiHnlJt/U_{y}^{J}=e^{-iH_{nl}^{J}t/\hbar}. Here, a^J+\hat{a}_{J+}, a^J\hat{a}_{J-} are the boson destruction operators for the field modes J+J+ and JJ-, and κ\kappa and gg are the interaction constants. The θ\theta is a function of the interaction time tt and can be selected so that θ=π/4\theta=\pi/4. To realize UxJU_{x}^{J} at each site JA,BJ\equiv A,B, CC, we suppose the field modes J+J+ and JJ- are spatially separated at the site JJ, so that a phase shift θp\theta_{p} can be applied along one arm, that of mode JJ-, as used in the detection of NOON states noon-mitchell ; noon-dowling ; noon-afek ; herald-noon-1-2 . For a suitable choice of θp\theta_{p}, this induces an overall relative phase shift between the modes, allowing realisation of the final transformation

(UxJ)1|N,0Jcosθ|N,0J+sinθ|0,NJ.(U_{x}^{J})^{-1}|N,0\rangle_{J}\rightarrow\cos\theta|N,0\rangle_{J}+\sin\theta|0,N\rangle_{J}. (61)

XI.3 Considerations for a realistic test of the EPR-Bohm paradox

The EPR-Bohm paradox for the two-spin set-up of Section III.A.1 can be signified when

(Δinfσ^yA)2+(Δinfσ^zA)2<1(\Delta_{inf}\hat{\sigma}_{y}^{A})^{2}+(\Delta_{inf}\hat{\sigma}_{z}^{A})^{2}<1 (62)

where (Δinfσ^θA)2(\Delta_{inf}\hat{\sigma}_{\theta}^{A})^{2} is the variance associated with the estimate inferred for the outcome of σ^θA\hat{\sigma}_{\theta}^{A} given a result for a measurement of σ^ϕB\hat{\sigma}_{\phi}^{B} on system BB. The value of ϕ\phi is chosen optimally to minimize the error epr-r2 ; epr-rmp . Hence, ϕ=θ\phi=\theta. A sufficient condition that the inequality be satisfied is that

(Δ(σ^yA+σ^yB))2+(Δ(σ^zA+σ^zB))2<1.(\Delta(\hat{\sigma}_{y}^{A}+\hat{\sigma}_{y}^{B}))^{2}+(\Delta(\hat{\sigma}_{z}^{A}+\hat{\sigma}_{z}^{B}))^{2}<1. (63)

Then the estimate of the outcomes for σ^θA\hat{\sigma}_{\theta}^{A} is taken to be σθB-\sigma_{\theta}^{B}, where σθB\sigma_{\theta}^{B} is the outcome of the measurement σ^θB\hat{\sigma}_{\theta}^{B}, for θ=x,y\theta=x,y. The bound of 11 is half that given by Hofmann and Takeuchi for the entanglement criterion, Eq (24) of their paper hofmann-take , as expected for an EPR-steering inequality epr-steer ; epr-rmp . Clearly, for the EPR-Bohm test given in Sections III and V.B, the inequality is satisfied since there is a perfect anticorrelation between the outcomes, implying the left-side has a value of zero. Similar inequalities can be derived for the three-spin set-up epr-rmp .

For the EPR-Bohm test of Section V.A, the inequality is also satisfied in the limit of α\alpha large. However, for finite α\alpha, the states |α|\alpha\rangle and |α|-\alpha\rangle are not truly orthogonal. We propose a realistic experiment as follows. Two orthogonal states |+|+\rangle and ||-\rangle are defined. The spin operators are σ^z=|++|||\hat{\sigma}_{z}=|+\rangle\langle+|-|-\rangle\langle-| and σ^y=(|+||+|)/i\hat{\sigma}_{y}=(|+\rangle\langle-|-|-\rangle\langle+|)/i, where ++ indicates a state with an outcome for X^\hat{X} that is non-negative, x0x\geq 0, and - indicates a state with an outcome for X^\hat{X} that is negative, x<0x<0. The notation x+x\in+ implies x0x\geq 0; the notation xx\in- implies x<0x<0. The state being measured is |ψBell|\psi_{Bell}\rangle, which can be expressed as

c1|α+c2|α\displaystyle c_{1}|\alpha\rangle+c_{2}|-\alpha\rangle =\displaystyle= c+|+z+c|z\displaystyle c_{+}|+\rangle_{z}+c_{-}|-\rangle_{z} (64)

where

|±=x±(c1x|α+c2x|α)[x±|c1x|α+c2x|α|2]1/2|x|\pm\rangle=\frac{\sum_{x\in\pm}(c_{1}\langle x|\alpha\rangle+c_{2}\langle x|-\alpha\rangle)}{[\sum_{x\in\pm}|c_{1}\langle x|\alpha\rangle+c_{2}\langle x|-\alpha\rangle|^{2}]^{1/2}}|x\rangle (65)

and c±=[x±|c1x|α+c2x|α|2]1/2c_{\pm}=[\sum_{x\in\pm}|c_{1}\langle x|\alpha\rangle+c_{2}\langle x|-\alpha\rangle|^{2}]^{1/2}. The measurement of SzS_{z} corresponds to first determining whether the outcome xx of X^\hat{X} is non-negative or negative.

P(±)\displaystyle P(\pm) =\displaystyle= |c±|2=x±|c1x|α+c2x|α|.2\displaystyle|c_{\pm}|^{2}=\sum_{x\in\pm}|c_{1}\langle x|\alpha\rangle+c_{2}\langle x|-\alpha\rangle|.^{2} (66)

The overlap function for x^=(a^+a^)/2\hat{x}=(\hat{a}+\hat{a}^{\dagger})/\sqrt{2} where α\alpha is real and positive is x|αe(x2α)2/2/π1/4\langle x|\alpha\rangle\sim e^{-(x-\sqrt{2}\alpha)^{2}/2}/\pi^{1/4} yurke-stoler-1 . Hence,

P(+)\displaystyle P(+) =\displaystyle= x0|c1x|α+c2x|α|2\displaystyle\sum_{x\geq 0}|c_{1}\langle x|\alpha\rangle+c_{2}\langle x|-\alpha\rangle|^{2}
=\displaystyle= (x0|c1|2|x|α|2+|c2|2|x|α|2\displaystyle(\sum_{x\geq 0}|c_{1}|^{2}|\langle x|\alpha\rangle|^{2}+|c_{2}|^{2}|\langle x|-\alpha\rangle|^{2}
+c1c2x|αα|x+c1c2x|αα|x).\displaystyle+c_{1}c_{2}^{*}\langle x|\alpha\rangle\langle-\alpha|x\rangle+c_{1}^{*}c_{2}\langle x|-\alpha\rangle\langle\alpha|x\rangle).

The leading term is P(+)=x0|c1|2|x|α|2P(+)=\sum_{x\geq 0}|c_{1}|^{2}|\langle x|\alpha\rangle|^{2}. The second term is an integral in the tail of the Gaussian, e(x2α)2/π1/2e^{-(x-\sqrt{2}\alpha)^{2}}/\pi^{1/2}, which has mean μ=2α\mu=\sqrt{2}\alpha and standard deviation σ=1/2\sigma=1/\sqrt{2}. Taking a conservative value of α>2\alpha>2, the probability of the tail is much less than 0.030.03. The third and fourth terms are damped by a term of order eα2e^{-\alpha^{2}}, and are also negligible for α>2\alpha>2. A similar result holds for P()P(-). The error in assuming x|α=0\langle x|-\alpha\rangle=0 for x0x\geq 0 and x|α=0\langle x|\alpha\rangle=0 for x<0x<0 introduces errors of much less than 1010% in the variances of (63). Since the uncertainty bound for the inequality is of order 11, the errors are negligible for α>2\alpha>2.

The measurement of SyS_{y} requires the rotation

U(c+|+z+c|z)\displaystyle U(c_{+}|+\rangle_{z}+c_{-}|-\rangle_{z}) =\displaystyle= 12(c+ic)eiπ/4|+y\displaystyle\frac{1}{\sqrt{2}}(c_{+}-ic_{-})e^{i\pi/4}|+\rangle_{y}
+12(c++ic)eiπ/4|y.\displaystyle+\frac{1}{\sqrt{2}}(c_{+}+ic_{-})e^{-i\pi/4}|-\rangle_{y}.

so that the probability for an outcome x+x\in+ is

|c+ic|2\displaystyle|c_{+}-ic_{-}|^{2} =\displaystyle= |c+|2+|c|2\displaystyle|c_{+}|^{2}+|c_{-}|^{2}
=\displaystyle= x+|c1x|α+c2x|α|2\displaystyle\sum_{x\in+}|c_{1}\langle x|\alpha\rangle+c_{2}\langle x|-\alpha\rangle|^{2}
+x|c1x|α+c2x|α|2.\displaystyle+\sum_{x\in-}|c_{1}\langle x|\alpha\rangle+c_{2}\langle x|-\alpha\rangle|^{2}.

As above, the leading terms are

P(+)=x+|c1|2|x|α|2+x|c2|2|x|α|2,P(+)=\sum_{x\in+}|c_{1}|^{2}|\langle x|\alpha\rangle|^{2}+\sum_{x\in-}|c_{2}|^{2}|\langle x|-\alpha\rangle|^{2}, (69)

the remaining terms being negligible for α>2\alpha>2. The actual measurement is approximate, being the application of UU where:

U(c1|α+c2|α)\displaystyle U(c_{1}|\alpha\rangle+c_{2}|-\alpha\rangle) =\displaystyle= c12(eiπ/4|α+eiπ/4|α)\displaystyle\frac{c_{1}}{\sqrt{2}}(e^{i\pi/4}|\alpha\rangle+e^{-i\pi/4}|-\alpha\rangle)
ic22(eiπ/4|α+eiπ/4|α).\displaystyle-\frac{ic_{2}}{\sqrt{2}}(e^{i\pi/4}|\alpha\rangle+e^{-i\pi/4}|-\alpha\rangle).

This gives

x0{(c1ic2)2eiπ/4x|α+(c1+ic2)2eiπ/4x|α}|x\displaystyle\sum_{x\geq 0}\{\frac{(c_{1}-ic_{2})}{\sqrt{2}}e^{i\pi/4}\langle x|\alpha\rangle+\frac{(c_{1}+ic_{2})}{\sqrt{2}}e^{-i\pi/4}\langle x|-\alpha\rangle\}|x\rangle
+x<0{(c1ic2)2eiπ/4x|α+(c1+ic2)2eiπ/4x|α}|x.\displaystyle+\sum_{x<0}\{\frac{(c_{1}-ic_{2})}{\sqrt{2}}e^{i\pi/4}\langle x|\alpha\rangle+\frac{(c_{1}+ic_{2})}{\sqrt{2}}e^{-i\pi/4}\langle x|-\alpha\rangle\}|x\rangle.

Hence,

P(+)\displaystyle P(+) =\displaystyle= 12x0|(c1ic2)eiπ/4x|α\displaystyle\frac{1}{2}\sum_{x\geq 0}|(c_{1}-ic_{2})e^{i\pi/4}\langle x|\alpha\rangle (71)
+(c1+ic2)eiπ/4x|α|2.\displaystyle+(c_{1}+ic_{2})e^{-i\pi/4}\langle x|-\alpha\rangle|^{2}.

The leading term is given by (69), as required. Any errors in the approximate measurement procedure lead to errors which compared to the uncertainty bound of 11 are negligible for α>2\alpha>2.

XI.4 Considerations for a realistic test of the EPR-Bohm paradox based on weak local realism

The EPR-Bohm paradox based on wLR (or wMR) as described in the set-ups of Figures 2 and 4 can be signified when

(Δinfσ^yA)2+(Δdσ^zA)2<1.(\Delta_{inf}\hat{\sigma}_{y}^{A})^{2}+(\Delta_{d}\hat{\sigma}_{z}^{A})^{2}<1. (72)

for measurements on the system prepared at the time tft_{f}. Here (Δinfσ^yA)2(\Delta_{inf}\hat{\sigma}_{y}^{A})^{2} is the square of the error in the inferred value for σy^A\hat{\sigma_{y}}^{A} given the result for the measurement σ^yB\hat{\sigma}_{y}^{B} on system BB. The (Δdσ^zA)2(\Delta_{d}\hat{\sigma}_{z}^{A})^{2} is the square of the error in distinguishing the spin states for the state of system AA as prepared at the time tft_{f}. The inference variance can be measured using standard techniques, as in Appendix C. It is also necessary to confirm that system AA is given quantum mechanically as a spin 1/21/2 system, which includes defining the two spin eigenstates and demonstrating both spin measurements σ^yA\hat{\sigma}_{y}^{A} and σ^zA\hat{\sigma}_{z}^{A} (and their non-commutativity) for the entangled system at time tft_{f}, as well as confirming the lower bound of the inequality (7). The experiment of Ref. sch-epr-exp-atom reports simultaneous measurement along these lines. There is no violation of the uncertainty principle for the state of system AA at time tft_{f}, because the system AA is in one or other spin state with equal probability, implying (Δσ^zA)2=1(\Delta\hat{\sigma}_{z}^{A})^{2}=1. In the macroscopic proposals, the pseudo-spin states are the coherent states |α|\alpha\rangle and |α|-\alpha\rangle, or else |z|zN|\uparrow\rangle_{z}\equiv|\uparrow\rangle_{z}^{\otimes N} and |z|zN|\downarrow\rangle_{z}\equiv|\downarrow\rangle_{z}^{\otimes N}. Noise can diminish the effectiveness of the measurement σ^zA\hat{\sigma}_{z}^{A} , increasing (Δdσ^zA)2(\Delta_{d}\hat{\sigma}_{z}^{A})^{2}. The analysis in Appendix C indicates that the error due to the overlap of the coherent states becomes negligible for α>2\alpha>2, so that (Δdσ^zA)20(\Delta_{d}\hat{\sigma}_{z}^{A})^{2}\rightarrow 0.

References

  • (1) A. Einstein, B. Podolsky, and N. Rosen, Can Quantum-Mechanical Description of Physical Reality Be Considered Complete?, Phys. Rev. 47, 777 (1935).
  • (2) J. S. Bell, On the Einstein-Podolsky-Rosen paradox, Physics 1,195 (1964).
  • (3) J. F. Clauser, M. A. Horne, A. Shimony and R. A. Holt, Proposed experiment to test local hidden-variable theories, Phys. Rev. Lett. 23, 880 (1969).
  • (4) J. S. Bell, Introduction to the hidden-variable question, in: Foundations of Quantum Mechanics ed B d’Espagnat (New York: Academic) pp171-181 (1971).
  • (5) J. F. Clauser and A. Shimony, Bell’s theorem: experimental tests and implications, Rep. Prog. Phys. 41, 1881 (1978).
  • (6) N. Brunner, D. Cavalcanti, S. Pironio, V. Scarani, and S. Wehner, Bell nonlocality, Rev. Mod. Phys. 86, 419 (2014).
  • (7) D. M. Greenberger, M. Home, A. Zeilinger, Going Beyond Bell’s Theorem, in Quantum Theory, and Conceptions of the Universe, M. Kafatos, ed., Kluwer, Dordrecht, The Netherlands (1989), p. 69.
  • (8) D. M. Greenberger, M. A. Horne, A. Zeilinger, Bell’s theorem without inequalities, American Journal of Physics 58, 1131(1990).
  • (9) N. David Mermin, What’s wrong with these elements of reality? Phys. Today 43 (6), 9 (1990).
  • (10) R. K. Clifton, M. L. G. Redhead, J. N. Butterfield, Generalization of the Greenberger-Horne-Zeilinger algebraic proof of nonlocality, Foundations of Physics 21, 149 (1991).
  • (11) D. Bouwmeester, J.-W. Pan, M. Daniell, H. Weinfurter, A. Zeilinger, Observation of three-photon Greenberger-Horne-Zeilinger entanglement, Phys. Rev. Lett. 82, 1345 (1999).
  • (12) J.-W. Pan, D. Bouwmeester, M. Daniell, H. Weinfurter, A. Zeilinger, Experimental test of quantum nonlocality in three-photon Greenberger–Horne–Zeilinger entanglement, Nature 403, 515 (2000).
  • (13) M. D. Reid, P. D. Drummond, W. P. Bowen, E. G. Cavalcanti, P. K. Lam, H. A. Bachor, U. L. Andersen and G. Leuchs, The Einstein-Podolsky-Rosen paradox: From concepts to applications. Rev. Mod. Phys. 81, 1727 (2009).
  • (14) E. Schrödinger, Die gegenwärtige Situation in der Quantenmechanik, Die Naturwissenschaften, 23, 807-812; 823-828; 844-849 (1935).
  • (15) A. Leggett and A. Garg, Quantum mechanics versus macroscopic realism: is the flux there when nobody looks?, Phys. Rev. Lett. 54, 857 (1985).
  • (16) D. Bohm, Quantum Theory (Prentice-Hall, New York, 1951).
  • (17) D. Bohm and Y. Aharonov, Discussion of Experimental Proof for the Paradox of Einstein, Rosen and Podolsky, Phys. Rev. 108, 1070 (1957).
  • (18) A. Aspect, P. Grangier, G. Roger, Experimental Realization of Einstein-Podolsky-Rosen Bohm Gedankenexperiment: A new violation of Bell’s inequalities. Phys. Rev. Lett. 49, 91 (1982).
  • (19) M. Thenabadu and M.D. Reid, Bipartite Leggett-Garg and macroscopic Bell inequality violations using cat states: distinguishing weak and deterministic macroscopic realism, Phys. Rev. A 105, 052207 (2022). M. D. Reid and M Thenabadu, Weak versus deterministic macroscopic realism, arXiv:2101.09476 [quant-ph].
  • (20) M. Thenabadu and M. D. Reid, Macroscopic delayed-choice and retrocausality: quantum eraser, Leggett-Garg and dimension witness tests with cat states, Phys. Rev. A 105, 062209 (2022).
  • (21) J. S. Bell, On the Problem of Hidden Variables in Quantum Mechanics, Rev. Mod. Phys. 38, 447-52 (1966).
  • (22) H. Jeong, M. Paternostro and T. C. Ralph, Failure of Local Realism Revealed by Extremely-Coarse-Grained Measurements, Phys. Rev. Lett. 102, 060403 (2009).
  • (23) M. Thenabadu, G-L. Cheng, T. L. H. Pham, L. V. Drummond, L. Rosales-Zárate and M. D. Reid, Testing macroscopic local realism using local nonlinear dynamics and time settings, Phys. Rev. A 102, 022202 (2020).
  • (24) R. Rushin Joseph, M. Thenabadu, C. Hatharasinghe, J. Fulton, R-Y Teh, P. D. Drummond and M. D. Reid, Wigner’s Friend paradoxes: consistency with weak-contextual and weak-macroscopic realism models, arXiv 2211.02877 [quant-ph].
  • (25) P. Colciaghi, Y. Li, P. Treutlein, and T. Zibold, Einstein-Podolsky-Rosen experiment with two Bose-Einstein condensates, Physical Review X13, 021031 (2023).
  • (26) H. Hofmann and S. Takeuchi, Violation of local uncertainty relations as a signature of entanglement, Phys. Rev. A 68, 032103 (2003).
  • (27) N. Bohr, Can quantum-mechanical description of physical reality be considered complete?, Phys. Rev. 48, 696 (1935).
  • (28) C. Wang et al., A Schrödinger cat living in two boxes, Science 352, 1087 (2016).
  • (29) Zaki Leghtas, Gerhard Kirchmair, Brian Vlastakis, Michel H. Devoret, Robert J. Schoelkopf, and Mazyar Mirrahimi, Deterministic protocol for mapping a qubit to coherent state superpositions in a cavity. Phys. Rev. A 87, 042315 (2013).
  • (30) B. Yurke and D. Stoler, Generating quantum mechanical superpositions of macroscopically distinguishable states via amplitude dispersion, Phys. Rev. Lett. 57, 13 (1986).
  • (31) E. Wright, D. Walls and J. Garrison, Collapses and Revivals of Bose-Einstein Condensates Formed in Small Atomic Samples, Phys. Rev. Lett. 77, 2158 (1996).
  • (32) M. Greiner, O. Mandel, T. Hånsch and I. Bloch, Collapse and revival of the matter wave field of a Bose-Einstein condensate, Nature 419, 51 (2002).
  • (33) G. Kirchmair et al., Observation of the quantum state collapse and revival due to a single-photon Kerr effect, Nature 495, 205 (2013).
  • (34) M. Thenabadu and M. D. Reid, Leggett-Garg tests of macrorealism for dynamical cat states evolving in a nonlinear medium, Phys. Rev. A 99, 032125 (2019).
  • (35) B. Vlastakis, G. Kirchmair, Z. Leghtas, S. E. Nigg, L. Frunzio, S. M. Girvin, M. Mirrahimi, M. H. Devoret, R. J. Schoelkopf, Deterministically encoding quantum information using 100-photon schrödinger cat states, Science 342, 607 (2013).
  • (36) M. Mirrahimi, Z. Leghtas, V. V. Albert, S. Touzard, R. J. Schoelkopf, L. Jiang, M. H. Devoret, Dynamically protected cat-qubits: a new paradigm for universal quantum computation, New Journal of Physics 16, 045014 (2014).
  • (37) F. Frowis, P. Sekatski, W. Dur, N. Gisin, and N. Sangouard, Macroscopic quantum states: measures, fragility, and implementations, Rev. Mod. Phys. 90, 025004 (2018).
  • (38) A. Ourjoumtsev, H. Jeong, R. Tualle-Brouri, P. Grangier, Generation of optical ‘Schrödinger cats’ from photon number states, Nature 448,784 (2007).
  • (39) M. Wolinsky, H. J. Carmichael, Quantum noise in the parametric oscillator: From squeezed states to coherent-state superpositions, Phys. Rev. Lett. 60 1836 (1988).
  • (40) L. Krippner, W. J. Munro, M. D. Reid, Transient macroscopic quantum superposition states in degenerate parametric oscillation: Calculations in the large-quantum-noise limit using the positive p representation, Phys. Rev. A 50, 4330 (1994).
  • (41) E. E. Hach III, C. C. Gerry, Generation of mixtures of schrödinger-cat states from a competitive two-photon process, Phys. Rev. A 49, 490 (1994).
  • (42) L. Gilles, B. M. Garraway, P. L. Knight, Generation of nonclassical light by dissipative two-photon processes, Phys. Rev. A 49, 2785 (1994).
  • (43) R. Y. Teh, F.-X. Sun, R. Polkinghorne, Q. Y. He, Q. Gong, P. D. Drummond, M. D. Reid, Dynamics of transient cat states in degenerate parametric oscillation with and without nonlinear kerr interactions, Phys. Rev. A 101, 043807 (2020).
  • (44) H.-Y. Ku, N. Lambert, F.-J. Chan, C. Emary, Y.-N. Chen, F. Nori, Experimental test of non-macrorealistic cat states in the cloud, npj Quantum Information 6, 98 (2020).
  • (45) A. Omran, H. Levine, A. Keesling, G. Semeghini, T. T. Wang, S. Ebadi, H. Bernien, A. S. Zibrov, H. Pichler, S. Choi, et al., Generation and manipulation of Schrödinger cat states in Rydberg atom arrays, Science 365, 570 (2019).
  • (46) C. Emary, N. Lambert, and F. Nori, Leggett-Garg inequalities, Rep. Prog. Phys 77, 016001 (2014).
  • (47) A. Asadian, C. Brukner, and P. Rabl, Probing Macroscopic Realism via Ramsey Correlation Measurements, Phys. Rev. Lett. 112, 190402 (2014).
  • (48) G. C. Knee, K. Kakuyanagi, M.-C. Yeh, Y. Matsuzaki, H. Toida, H. Yamaguchi, S. Saito, A. J. Leggett and W. J. Munro, A strict experimental test of macroscopic realism in a superconducting flux qubit, Nat. Commun. 7, 13253 (2016).
  • (49) R. Uola, G. Vitagliano and C. Budroni, Leggett-Garg macrorealism and the quantum nondisturbance conditions, Phys. Rev. A 100, 042117 (2019).
  • (50) M. D. Reid and W. J. Munro, Macroscopic boson states exhibiting the Greenberger-Horne-Zeilinger contradiction with local realism, Phys. Rev. Lett. 69, 997 (1992).
  • (51) Adan Cabello, Multiparty multilevel Greenberger-Horne-Zeilinger states, Phys. Rev. A 63, 022104 (2001).
  • (52) W. Son, Jinhyoung Lee, and M. S. Kim, Generic Bell Inequalities for Multipartite Arbitrary Dimensional Systems, Phys. Rev. Lett. 96, 060406 (2006).
  • (53) N. D. Mermin, Extreme Quantum Entanglement in a Superposition of Macroscopically Distinct States, Phys. Rev. Lett. 65, 1838 (1990).
  • (54) D. Bohm, A suggested interpretation of quantum theory in terms of “hidden” variables”, Phys. Rev. 85, pp 166-179 (1952).
  • (55) R. Spekkens, Evidence for the epistemic view of quantum states: A toy theory, Phys Rev A 75, 032110 (2007).
  • (56) M. J. Hall, D-A. Deckert, and H. J. Wiseman, Quantum Phenonema modeled by interactions between Many Classical Worlds, Phys. Rev. X4, 041013 (2014).
  • (57) A. Auffèves and P. Grangier, Contexts, Systems and Modalities: A new ontology for quantum mechanics, Found. Phys. 46, 121 (2016).
  • (58) P. Grangier, Contextual Inferences, Nonlocality, and the Incompleteness of Quantum Mechanics, Entropy 23, 1660 (2021).
  • (59) O. J. E. Maroney, Measurements, disturbance and the three-box paradox, Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics, 58, 41 (2017).
  • (60) O. J. E. Maroney and C. G. Timpson, Quantum- vs. Macro- Realism: What does the Leggett-Garg Inequality actually test?, arXiv:1412.6139 (2017).
  • (61) P. D. Drummond and M. D. Reid, Retrocausal model of reality for quantum fields, Phys. Rev. Research 2, 033266 (2020).
  • (62) P. D. Drummond and M. D. Reid, Objective Quantum Fields, Retrocausality and Ontology, Entropy 23, 749 (2021).
  • (63) S. Frederich, Introducing the Q-based interpretation of quantum theory, The British Journal for the Philosophy of Science, 0, ja (2021), pp null. arXiv 2106.13502 (2021).
  • (64) M. D. Reid and P. D. Drummond, A quantum phase-space equivalence leads to hidden causal loops in a model for measurement consistent with macroscopic realism, arXiv 2205.06070 and arXiv:2303.02373.
  • (65) G. Castagnoli, Unobservable causal loops explain both the quantum computational speedup and quantum nonlocality, Phys. Rev. A 104, 032203 (2021).
  • (66) R. B. Griffiths, Consistent Quantum Theory (Cambridge University Press, Cambridge, 2002).
  • (67) Robert B. Griffiths, Nonlocality claims are inconsistent with Hilbert-space quantum mechanics, Phys. Rev. A 101, 022117 (2020).
  • (68) Sandro Donadi and Sabine Hossenfelder, Toy model for local and deterministic wave-function collapse, Phys. Rev. A 106, 022212 (2022).
  • (69) R. Schnabel, The solution to the Einstein-Podolsky-Rosen paradox, arXiv:2208.13831.
  • (70) C. Brukner, A No-Go Theorem for Observer-Independent Facts, Entropy 20, 350 (2018).
  • (71) D. Frauchiger and R. Renner, Quantum theory cannot consistently describe the use of itself, Nat. Commun. 9, 1 (2018).
  • (72) M. Losada, R. Laura, and O. Lombardi, Frauchiger-Renner argument and quantum histories, Phys. Rev. A 100, 052114 (2019).
  • (73) A. Matzkin and D. Sokolovski, Wigner Friend scenarios with non-invasive weak measurements, Phys. Rev. A 102, 062204 (2020).
  • (74) D. Lazarovici and M. Hubert, How Quantum Mechanics can consistently describe the use of itself, Sci. Rep. 9, 470 (2019).
  • (75) S. Kochen and E. Specker, The problem of hidden variables in quantum mechanics, Journal of mathematics and mechanics 17, 59 (1967).
  • (76) J. Bell, Against measurement, Phys. World 3, 8 pp 33 (1990).
  • (77) J. Bell, Speakable and unspeakable in quantum mechanics: Collected papers on quantum philosophy, (Cambridge University Press, 2004).
  • (78) N. Harrigan and R. Spekkens, Einstein, incompleteness, and the epistemic view of quantum states, Found. Phys. 40, 125 (2010).
  • (79) P. Grangier, Contextual Inferences, Nonlocality, and the Incompleteness of Quantum Mechanics, Entropy 23, 1660 (2021) and references therein.
  • (80) E. G. Cavalcanti, P. D. Drummond, H. A. Bachor, and M. D. Reid, Spin entanglement, decoherence and Bohm’s EPR paradox, Optics Express 17 (21), 18693 (2009).
  • (81) M. D. Reid, Demonstration of the Einstein-Podolsky-Rosen Paradox using Nondegenerate Parametric Amplification, Phys. Rev. A 40, 913 (1989).
  • (82) M. Ardehali, Bell inequalities with a magnitude of violation that grows exponentially with the number of particles, Phys. Rev. A 46, 5375 (1992).
  • (83) A. V. Belinskiı, D. N. Klyshko, Interference of light and bell’s theorem, Phys.-Usp. 36, 653 (1993).
  • (84) H. J. Lipkin, N. Meshkov, and A. J. Glick, Validity of many-body approximation methods for a solvable model: exact solutions and perturbation theory, Nucl. Phys. 62 188 (1965).
  • (85) M. Steel and M. J. Collett, Quantum state of two trapped Bose-Einstein condensates with a Josephson coupling, Phys. Rev. A 57, 2920 (1998).
  • (86) M. W. Mitchell, J. S. Lundeen, and A. M. Steinberg, Super-resolving phase measurements with a multiphoton entangled state, Nature 429, 161 (2004).
  • (87) I. Afek, O. Ambar, and Y. Silberberg, High-NOON States by Mixing Quantum and Classical Light, Science 328, 879 (2010).
  • (88) J. P. Dowling, Quantum optical metrology –the lowdown on high-n00n states, Contemporary Physics 49, 125 (2008).
  • (89) S. Slussarenko, M. M. Weston, H. M. Chrzanowski, L. K. Shalm, V. B. Verma, S. W. Nam, G. J. Pryde, Unconditional violation of the shot-noise limit in photonic quantum metrology, Nature Photonics 11, 700 (2017).
  • (90) E. G. Cavalcanti, S. J. Jones, H. M. Wiseman, M. D. Reid, Experimental criteria for steering and the Einstein-Podolsky-Rosen paradox, Phys. Rev. A 80, 032112 (2009)