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

Bipartite Leggett-Garg and macroscopic Bell inequality violations using cat states: distinguishing weak and deterministic macroscopic realism

Manushan Thenabadu and M. D. Reid1 1 Centre for Quantum Science and Technology Theory, Swinburne University of Technology, Melbourne 3122, Australia
Abstract

We consider tests of Leggett-Garg’s macrorealism and of macroscopic local realism, where for spacelike separated measurements the assumption of macroscopic noninvasive measurability is justified by that of macroscopic locality. We give a mapping between the Bell and Leggett-Garg experiments for microscopic qubits based on spin 1/21/2 eigenstates and gedanken experiments for macroscopic qubits based on two macroscopically distinct coherent states (cat states). In this mapping, the unitary rotation of the Stern-Gerlach analyzer is realized by an interaction H=Ωn^4H=\Omega\hat{n}^{4} where n^\hat{n} is the number of quanta. By adjusting the time of interaction, one alters the measurement setting. We thus predict violations of Leggett-Garg and Bell inequalities in a macroscopic regime where coarse-grained measurements M^\hat{M} need only discriminate between two macroscopically distinct coherent states. To interpret the violations, we distinguish between different definitions of macroscopic realism. Deterministic macroscopic local realism (dMR) assumes a definite outcome for the measurement M^\hat{M} prior to the unitary rotation created by the analyser, and is negated by the violations. Weak macroscopic realism (wMR) assumes a definite outcome for systems prepared in a superposition ψpointer\psi_{pointer} of two macroscopically-distinct eigenstates of M^\hat{M}, after the unitary rotation. We find that wMR can be viewed as consistent with the violations. A model is presented, in which wMR holds, and for which the macroscopic violations emerge over the course of the unitary dynamics. Finally, we point out an EPR-type paradox, that a weak macro-realistic description for the system prior to the measurement M^\hat{M} is inconsistent with the completeness of quantum mechanics.

I Introduction

Bell’s theorem constrains the predictions of all local hidden variable theories to values governed by inequalities bell-brunner-rmp ; CShim-review ; Bell-2 ; det-bell . These inequalities can be violated by quantum mechanics, giving a falsification of both local realism and local causality. Bell tests involve measurements made on two separated particles, performed on timescales such that the two measurement events are spacelike separated. At each location, there is a choice between two measurement settings. To date, almost all Bell tests have examined microscopic systems. Bell violations have been predicted for large numbers of particles meso-bell-higher-spin-cat-states-bell ; cv-bell ; cat-bell ; mdr-mlr ; macro-bell-additional ; cv-bell-macro-ali , but, in almost all cases, the measurements for one of the settings require a resolution of a few particles, or at the level of Planck’s constant.

By contrast, Leggett and Garg derived inequalities that do not require such fine resolution of measurement legggarg-1 . Leggett-Garg inequalities are designed to falsify the assumption of macrorealism (M-R) for systems that at certain times are in a superposition (e.g. ψα|a+|d\psi_{\alpha}\sim|a\rangle+|d\rangle) of two macroscopically distinct states s-cat-1 . M-R combines two premises: Macroscopic realism (MR) asserts that the system must actually be in one or other of those two states at the given time. The second premise, macroscopic noninvasive measurability (NIM), asserts the existence of a measurement M^\hat{M} that can distinguish between the two macroscopic states, with negligible effect on the subsequent dynamics (at a macroscopic level). Macroscopic realism implies a predetermined value λM\lambda_{M} for the outcome of the measurement M^\hat{M} legggarg-1 . Since the two states |a|a\rangle and |d|d\rangle can be distinguished by a measurement M^\hat{M} allowing a macroscopic uncertainty, the hidden variable λM\lambda_{M} need not specify the results of measurements to a resolution of \hbar. This was recognised by Leggett and Garg who assigned the values λM=±1\lambda_{M}=\pm 1 for the two outcomes of M^\hat{M}, in order to derive inequalities legggarg-1 .

Leggett-Garg inequalities have been reported violated for a wide range of systems emary-review ; jordan_kickedqndlg2-1 ; weak-solid-state-qubits ; NSTmunro-1 ; massiveosci-1-1 ; manushan-cat-lg ; weak-hybrid ; leggett-garg-recent ; macro-bell-lg ; nst . A complication for the interpretation of the violations has been the justification of the NIM premise for practical measurements leggett-garg-recent . Different strategies are used, including (in a recent macroscopic superconducting experiment) a control experiment which quantifies the amount of disturbance induced by the measurement, if the system is indeed in one of the states |a|a\rangle or |d|d\rangle NSTmunro-1 . An alternative strategy (that does not assume the system to be in one of the fully specified quantum states) is to violate bipartite Leggett-Garg or macroscopic Bell inequalities macro-bell-lg ; weak-hybrid ; cv-bell-macro-ali . Here, the NIM premise is replaced by that of macroscopic locality. Recent work gives predictions for such violations using NOON and multi-component cat states macro-bell-lg . The question becomes how to interpret such violations. In particular, we ask whether the assumption of the macroscopic hidden variable λM\lambda_{M} for the coarse-grained measurement M^\hat{M} can be negated.

In this paper, we give full details of the results presented in the Letter letter , which analyses this question. To facilitate the analysis, we give predictions of violations of macroscopic Bell inequalities and of bipartite Leggett-Garg inequalities for the conceptually simple case where a system at certain times is found to be in a superposition of two macroscopically distinct coherent states, |α|\alpha\rangle and |α|-\alpha\rangle (cat states), as α\alpha\rightarrow\infty. Similar to the results of Ref. macro-bell-lg , the violations are obtained using only measurements M^\hat{M} which discriminate between the two coherent states, thereby allowing macroscopic uncertainties.

In order to interpret the violations, we are first careful to define macroscopic realism in the weakest (i.e. most minimal) sense, as weak macroscopic realism (wMR), the premise that a macroscopic hidden variable λM\lambda_{M} specifies the outcome of the coarse-grained pointer measurement M^\hat{M}. In contrast to many previous macrorealism tests, we find it is not necessary to include in the definition of macroscopic realism that the system be in one or other of two specific quantum states (e.g. |α|\alpha\rangle and |α|-\alpha\rangle), or even that the system be in one or other of two unspecified quantum states. The violations given in this paper therefore provide strong tests of both macrorealism (M-R) and macroscopic local realism (MLR), and also of macroscopic local causality (MLC).

In this paper, we further weaken the definition of weak macroscopic realism, by specifying M^\hat{M} to be a pointer measurement, and the distinct states |a|a\rangle and |d|d\rangle to be states with a definite outcome for the (coarse-grained) M^\hat{M}. In other words, the assumption of wMR asserts the validity of the macroscopic hidden variable λM\lambda_{M} for the state ψα\psi_{\alpha} prepared after a unitary transformation UU that determines the choice of measurement setting. We show that the violations of M-R and MLR/ MLC given in this paper can be viewed consistently with wMR.

On the other hand, we show that the violations falsify deterministic macroscopic realism (dMR), which specifies well-defined values λM\lambda_{M} prior to such a unitary transformation. The assumption of dMR also includes the assumption of macroscopic locality (ML), which states that the value of λM\lambda_{M} cannot be changed by spacelike separated measurement events at another site. The premise of dMR implies the validity of two hidden variables, λMθ\lambda_{M_{\theta}} and λMθ\lambda_{M_{\theta^{\prime}}}, ascribed to the system simultaneously (prior to UU) to predetermine the outcome of two pointer measurements, M^θ\hat{M}_{\theta} and M^θ\hat{M}_{\theta^{\prime}}, and is negated by the violation of the macroscopic Bell inequalities.

We find that the consistency of the violations with wMR is possible because of the unitary dynamics associated with the choice of measurement setting. The unitary rotation U=eiHt/U=e^{-iHt/\hbar} that determines the measurement setting transforms the system from one pointer superposition to another, over a time interval Δt\Delta t. The system is therefore not considered to be in a superposition of eigenstates of both pointer measurements, M^θ\hat{M}_{\theta} and M^θ\hat{M}_{\theta^{\prime}}, simultaneously. In our analysis, the transformation UU that determines the measurement setting is given by a nonlinear Hamiltonian H=Ωn^4H=\Omega\hat{n}^{4} where n^\hat{n} is the mode number operator. The dynamics leading to the macroscopic violations can therefore be observed over the interval Δt\Delta t. Recent work predicts violations of Bell inequalities at a microscopic level for the dynamical trajectories of two entangled particles bell-trajectories . This is consistent with our conclusion that deterministic macroscopic realism fails, since the choice of measurement setting involves dynamics, as given by UU.

There are open questions. While we show consistency with wMR, the results of this paper neither falsify nor validate this premise. The validation of this premise might be expected, in view of the known emergence of classicality with coarse-grained measurements coarse-3 ; coarse-peres ; kofler-bruck-leggett-garg-coarse . The concept might also be consistent with recent proposals to interpret quantum mechanics using theories based on multiple interacting classical worlds mhall , or with an epistemic restriction of order Planck’s constant budiyono ; bartlett-gaussian ; macro-pointer ; q-trajectories-drummond , and may also be consistent with recent interpretations of the double-slit experiment double-slit-Aharonov .

However, it is clear from the violations of the Leggett-Garg inequalities presented in this paper (and elsewhere) that if wMR holds, then the dynamics associated with the macroscopic hidden variables λM\lambda_{M} cannot be given by classical mechanics. If wMR holds, we show for the tests of this paper that the violations of the Leggett-Garg inequalities would therefore arise from the breakdown of the NIM assumption. We explain how this breakdown arises (in the bipartite model) from quantum nonlocality, the new feature being that there is a macroscopic nonlocality (associated with macroscopically distinct qubits |α|\alpha\rangle and |α|-\alpha\rangle), which we show emerges over the timescales associated with the unitary interactions UU at both sites. While the distinction between predictions for the superposition ψα\psi_{\alpha} and the classical mixture ρmix\rho_{mix} of the two macroscopically distinct states is negligible (of order e|α|2\hbar e^{-|\alpha|^{2}}), a macroscopic difference emerges over the course of the dynamics corresponding to the unitary rotations.

From an alternative point of view, we summarise in this paper that if one does postulate the validity of wMR, then inconsistencies arise. This can be shown in the form of an EPR-type paradox epr-1 , similar to that given in earlier papers macro-coherence-paradox ; eric_marg-1 ; irrealism-fringes ; macro-pointer . If one assumes wMR, then the assumption that the system is in one of two quantum states, such as |α|\alpha\rangle or |α|-\alpha\rangle, can be negated. Hence, wMR is inconsistent with the completeness of quantum mechanics, which gives predictions at the level of \hbar.

Layout of paper: The paper is organised as follows. In Section II, we review and extend previous work to give derivations of bipartite Leggett-Garg and macroscopic Bell inequalities. The macroscopic Bell inequalities are derived assuming deterministic macroscopic (local) realism or macroscopic local causality, both of which are negated by violations of the inequalities given in this paper.

In Section III, we demonstrate a mapping between the spin-1/21/2 Bell experiments performed on microscopic qubits ||\uparrow\rangle and ||\downarrow\rangle, and a gedanken experiment performed on macroscopic qubits, |α|\alpha\rangle and |α|-\alpha\rangle. The unitary rotation of the polariser or spin analyser is mapped to a transformation |αcosθ|α+isinθ|α|\alpha\rangle\rightarrow\cos\theta|\alpha\rangle+i\sin\theta|-\alpha\rangle, which is realized for θ=0,π/8,π/4\theta=0,\pi/8,\pi/4 and 3π/83\pi/8 by an evolution U=eiHt/U=e^{-iHt/\hbar} at certain times tt manushan-cat-lg . We thus show that macroscopic Bell inequalities are violated for spacelike separated systems, AA and BB, prepared in an entangled two-mode cat state |ψ|α|β|α|β|\psi\rangle\sim|\alpha\rangle|-\beta\rangle-|-\alpha\rangle|\beta\rangle. The mapping relies on the orthogonality of the coherent states, which strictly holds only for α\alpha, β\beta\rightarrow\infty. In our explicit model, we give full calculations, and compute the values of α\alpha and β\beta for which experiments could be performed.

In Section IV, we predict violations of Leggett-Garg’s macrorealism, for both single and bipartite cat-state systems, following along the lines of Ref. manushan-cat-lg . In the Bell test, the angle θ\theta is the measurement setting. In the Leggett-Garg test, the angle θ\theta gives the time tt of evolution between measurements M^\hat{M} made at different times. The dynamics associated with the unitary evolution U=eiHt/U=e^{-iHt/\hbar} can be visualised by plotting the QQ function Husimi-Q-1 . We explain weak macroscopic realism (wMR) in Sections III.E and also in Section VI. Evaluation of the QQ function dynamics is given in Sections III.D and IV.B.2.

In Section V, we examine the measurement process more carefully. For the bipartite Leggett-Garg test, whether the measurement is performed or not on system AA is determined by the duration of unitary evolution, at the second location BB. By contrast, the timing of the final irreversible (“collapse”) stage of the measurement at BB is unimportant. We show by calculation that whether the collapse occurs before or after the subsequent evolution at AA is irrelevant to the violations, making only a difference of order e|α|2\mathcal{\hbar}e^{-|\alpha|^{2}} to the final probabilities. Here we point out that although the results of this paper are consistent with weak macroscopic realism (wMR), it is possible to establish inconsistencies with quantum mechanics, similar to an EPR-type paradox. Finally, a conclusion is given in Section VI.

II Macroscopic Bell and bipartite Leggett-Garg inequalities

We first outline and extend the Bell and bipartite Leggett-Garg inequalities for a macroscopic system, as proposed in Refs. weak-hybrid ; macro-bell-lg . In order to test macroscopic realism, we combine aspects of the original Bell and Leggett-Garg proposals. As with Bell’s proposal, there are two spatially separated systems AA and BB upon which local measurements are made. Similar to the Leggett-Garg proposal, the systems at each location evolve dynamically according to a local Hamiltonian H(A)H^{(A)} (or H(B)H^{(B)}), so that measurements can be made at different times. The measurements at each site AA and BB give only two possible outcomes, corresponding to two macroscopically distinguishable states. The outcomes are designated as pseudo-spins, SA=±1S_{A}=\pm 1 and SB=±1S_{B}=\pm 1, respectively.

In fact, we present two types of test of macroscopic realism. The first (Section II.A) is a macroscopic Bell test, where for a given run of the experiment, one makes single measurements at each site, AA and BB. For each measurement, there is a choice between two measurement settings, corresponding to a choice between two different times of interaction with a measurement device. The second type of test of macroscopic realism (Section II.B) is similar to that proposed by Leggett and Garg, except there are two sites. In a given run of the experiment, one makes sequential measurements at different times t1<t2<t3t_{1}<t_{2}<t_{3}, on the systems at the separated sites, AA and BB.

Refer to caption
Figure 1: Bell test of macroscopic local realism using local dynamics: Measurements M^θ(A)\hat{M}_{\theta}^{(A)} and M^ϕ(B)\hat{M}_{\phi}^{(B)} are made at the sites AA and BB respectively. The measurements have two stages. At each site, there is a choice of two measurement settings (θ\theta and θ\theta^{\prime}, and ϕ\phi and ϕ\phi^{\prime}) corresponding to times tat_{a} and tat^{\prime}_{a} at AA, and tbt_{b} and tbt^{\prime}_{b} at BB, during which the system interacts with a local medium. The interactions at AA and BB correspond to unitary transformations UA(θ)U_{A}(\theta) and UB(ϕ)U_{B}(\phi) respectively. After the unitary interactions, final pointer measurements M^(A)\hat{M}^{(A)} and M^(B)\hat{M}^{(B)} are made each giving two macroscopically distinct outcomes (denoted ±1\pm 1).

II.1 Macroscopic Bell test

II.1.1 Deterministic macroscopic (local) realism

We consider the situation of Figure 1. One prepares the state of the system at a time t=0t=0. At each site there are two possible measurement settings (θ\theta and θ\theta^{\prime} at AA, and ϕ\phi and ϕ\phi^{\prime} at BB) and for each of these measurements there are only two outcomes, corresponding to states of the system that are macroscopically distinct. We refer to the outcomes SA=±1S_{A}=\pm 1 and SB=±1S_{B}=\pm 1 as “spin” outcomes, and label the measurements at sites AA and BB as M^θ(A)\hat{M}_{\theta}^{(A)} and M^ϕ(B)\hat{M}_{\phi}^{(B)} respectively. Distinguishing only between macroscopically distinct outcomes, the measurements M^θ(A)\hat{M}_{\theta}^{(A)} and M^ϕ(B)\hat{M}_{\phi}^{(B)} can allow a significant error, beyond the level of ,\hbar, as we will show by example in Sections III and IV. As such, we refer to the measurements as coarse-grained, or macroscopic, measurements.

The assumption of macroscopic realism (MR), as specified by Leggett and Garg legggarg-1 , asserts that each system prior to the measurement M^θ(A)\hat{M}_{\theta}^{(A)} and M^ϕ(B)\hat{M}_{\phi}^{(B)} is in one or other of the two macroscopically distinguishable states, which have a definite outcome for the spin. Here, we do not wish to restrict that the two macroscopically distinct states are specific states, ψ1\psi_{1} and ψ2\psi_{2}, or even that they are quantum states. The definition of MR is generalised to postulate that prior to the measurement M^θ(A)\hat{M}_{\theta}^{(A)} and M^ϕ(B)\hat{M}_{\phi}^{(B)}, the system is in a state for which the outcome of the macroscopic, coarse-grained measurement M^θ(A)\hat{M}_{\theta}^{(A)} or M^ϕ(B)\hat{M}_{\phi}^{(B)} is definite, being either +1+1 or 1-1. Assuming MR, macroscopic hidden variables λθ(A)\lambda_{\theta}^{(A)} and λϕ(B)\lambda_{\phi}^{(B)} with values +1+1 or 1-1 can be ascribed to the systems AA and BB at the time immediately prior to the measurement, the values giving the predetermined outcomes for the measurements M^θ(A)\hat{M}_{\theta}^{(A)} and M^ϕ(B)\hat{M}_{\phi}^{(B)} respectively. These variables, introduced by Leggett and Garg, are hidden variables, since they are not explicitly given by quantum mechanics.

If there is a choice of measurement settings θ\theta and θ\theta^{\prime} at a given site, then one assumes (based on the assumption of MR) that two hidden variables λθ(A)\lambda_{\theta}^{(A)} and λθ(A)\lambda_{\theta^{\prime}}^{(A)} describe the state of the system as it exists prior to the time the choice is made, to predetermine the results of both measurements, M^θ(A)\hat{M}_{\theta}^{(A)} and M^θ(A)\hat{M}_{\theta^{\prime}}^{(A)}. Similar to the original Bell derivation Bell-2 ; det-bell , we call the definition of MR as given here deterministic macroscopic realism (dMR).

Assuming locality, that the value of the macroscopic hidden variable λ\lambda for the system at one site cannot be changed by the local measurement at the other site bell-brunner-rmp ; CShim-review ; Bell-2 ; det-bell , it is straightforward to derive the Clauser-Horne-Shimony-Holt (CHSH) Bell inequality, |B|2|B|\leq 2, where CShim-review

B\displaystyle B =\displaystyle= E(θ,ϕ)E(θ,ϕ)+E(θ,ϕ)+E(θ,ϕ)\displaystyle E(\theta,\phi)-E(\theta,\phi^{\prime})+E(\theta^{\prime},\phi)+E(\theta^{\prime},\phi^{\prime}) (1)

Here θ\theta and θ\theta^{\prime} (ϕ\phi and ϕ\phi^{\prime}) are two measurement settings at sites A (and B) respectively, and E(θ,ϕ)E(\theta,\phi) is the expectation value for the product SASBS_{A}S_{B}, for the choice of settings, M^θ(A)\hat{M}_{\theta}^{(A)} and M^ϕ(B)\hat{M}_{\phi}^{(B)}. The combined assumptions of locality and deterministic macroscopic realism are referred to as deterministic macroscopic local realism (MLR), or simply deterministic macroscopic realism, as the condition of locality is naturally implied by dMR for two sites det-bell . The derivation of (1) follows from the assumption of MLR, which implies E(θ,ϕ)=SASB=λθ(A)λϕ(B)E(\theta,\phi)=\langle S_{A}S_{B}\rangle=\langle\lambda_{\theta}^{(A)}\lambda_{\phi}^{(B)}\rangle.

The local measurements M^θ(A)\hat{M}_{\theta}^{(A)} and M^ϕ(B)\hat{M}_{\phi}^{(B)} have two stages: a unitary stage UU for which there is an interaction with an analyzer device, and a final stage M^\hat{M} which includes an irreversible coupling to a reservoir i.e. to a detector. The unitary stage determines the measurement setting, and is analogous to the transmission of a photon or spin-1/21/2 particle through a polarizing beam splitter or Stern-Gerlach analyzer. Thus, in the traditional Bell tests, θ\theta and ϕ\phi are polarizer or analyzer angles, determining the direction of polarisation or spin that is to be measured at each site. The final pointer stage M^\hat{M} corresponds to detection in one of the arms of the polarizing beam splitter.

For the examples given in this paper, following Refs. macro-bell-lg ; manushan-cat-lg , the unitary stage of the measurement at each site comprises an interaction with a nonlinear medium, and the measurement setting is determined by the time duration of the interaction. This need not be a choice of time itself, but can instead be a choice of the degree of nonlinearity of the medium, the choice of measurement setting being made by a switch. We identify at each site two choices of times for the measurements: tat_{a} and tat_{a}^{\prime} (tbt_{b} and tbt_{b}^{\prime}) at AA (and BB) respectively. The Bell inequality (1) becomes |B|2|B|\leq 2 where

B\displaystyle B =\displaystyle= E(ta,tb)E(ta,tb)+E(ta,tb)+E(ta,tb)\displaystyle E(t_{a},t_{b})-E(t^{\prime}_{a},t^{\prime}_{b})+E(t^{\prime}_{a},t_{b})+E(t^{\prime}_{a},t^{\prime}_{b}) (2)

Modelling the measurement in this way as an interaction with a device emphasises that the measurement will occur over a finite time interval. To justify the locality assumption, it is assumed that the measurement events at each site are spacelike separated.

II.1.2 Macroscopic local causality

When deriving (1) and (2), it is assumed that two macroscopic hidden variables describe each system at the time immediately prior to the measurement (t=0t=0) bell-brunner-rmp ; CShim-review ; Bell-2 ; det-bell . These are λθ(A)\lambda_{\theta}^{(A)} and λθ(A)\lambda_{\theta^{\prime}}^{(A)} for system AA, and λϕ(B)\lambda_{\phi}^{(B)} and λϕ(B)\lambda_{\phi^{\prime}}^{(B)} for system BB. Each of these variables assumes values +1+1 or 1-1. Thus, the system at AA is in a state with simultaneous predetermined values for the measurements M^θ\hat{M}_{\theta} and M^θ\hat{M}_{\theta^{\prime}}, and similarly for BB.

However, it is well known that the Bell inequality (1) can be derived with a weaker assumption allowing for a stochastic interaction with a local measurement device det-bell ; CShim-review . Here, one assumes that the system at time t=0t=0 is in a hidden variable state given by a set of hidden variables {λ}\{\lambda\} with probability density ρ(λ)\rho(\lambda). One defines the probability PA(±1|λ,θ)P_{A}(\pm 1|\lambda,\theta) for obtaining a spin outcome +1+1 or 1-1 at site AA, given the system prior to measurement is in the state {λ}\{\lambda\} and given the choice of measurement setting θ\theta for the local device at AA. The probability PB(±1|λ,ϕ)P_{B}(\pm 1|\lambda,\phi) is defined similarly, for the spin outcome ±1\pm 1 at BB, given the measurement setting ϕ\phi for the local measurement device at BB. The locality assumption is that probability PA(±1|λ,θ)P_{A}(\pm 1|\lambda,\theta) does not depend on ϕ\phi, and PB(±1|λ,ϕ)P_{B}(\pm 1|\lambda,\phi) does not depend on θ\theta. These assumptions can be viewed as a single assumption, local causality det-bell .

Thus, the Bell inequalities (1) and (2) follow from a weaker assumption, that we refer to as macroscopic local causality (MLC). The assumption is similar to local causality, except here one postulates that prior to the measurement, the system is in a hidden variable state which gives a definite probability for just two macroscopically distinct outcomes. In this paper, the Bell inequalities can be derived from either MLR or MLC, and we will use MLR to imply either definition.

II.2 Bipartite Leggett-Garg tests

The Leggett-Garg inequalities were derived for the situation where one considers succcessive measurements on a single system, at times tit_{i}. Here, following macro-bell-lg ; weak-hybrid and different to the original treatment legggarg-1 , we consider two systems AA and BB (Figure 2). Measurements Mi^(A)\hat{M_{i}}^{(A)} and M^j(B)\hat{M}_{j}^{(B)} are made on systems AA and BB at times tit_{i} and tjt_{j} respectively. The measurements give two outcomes, +1+1 and 1-1, corresponding to two macroscopically distinguishable states. However, here there is no choice of measurement setting at a given time: Rather, the unitary rotation stage UAU_{A} (or UBU_{B}) of the measurement is considered to be part of the dynamics. In the examples we consider in this paper, the measurements Mi^(A)\hat{M_{i}}^{(A)} and M^j(B)\hat{M}_{j}^{(B)} that occur at the times tit_{i} and tjt_{j} are analogous to the pointer measurements, M^(A)\hat{M}^{(A)} and M^(B)\hat{M}^{(B)}, of Figure 1.

The Leggett-Garg inequalities are derived based on the assumption of macroscopic realism legggarg-1 . Hence one may identify specific macroscopic hidden variables λi(A)\lambda_{i}^{(A)} and λi(B)\lambda_{i}^{(B)} to describe the states immediately prior to the pointer measurements Mi(A)^\hat{M_{i}^{(A)}} and M^j(B)\hat{M}_{j}^{(B)}. Different to the Bell derivation of Section II.A.1 however, one does not ascribe to the system simultaneously at a time tit_{i} predetermined results λi(A)\lambda_{i}^{(A)} and λj(A)\lambda_{j}^{(A)} for both future measurements, M^i(A)\hat{M}_{i}^{(A)} and M^j(A)\hat{M}_{j}^{(A)}. What was the local unitary interaction (UAU_{A} or UBU_{B}) with the measurement device in the Bell test is now the local trajectory. One therefore assumes macroscopic realism for a measurement with just a single setting, at the time tit_{i}. This means that the Leggett-Garg-Bell inequalities for the systems of interest may be derived based on the assumption of weak macroscopic realism (wMR): that for a system AA prepared in a superposition of macroscopically distinct pointer states, the system can be described by a single hidden variable λi\lambda_{i} at that time tit_{i} (and this is not affected by a space-like separated measurement event at the different site BB).

The traditional Leggett-Garg test invokes a second assumption, that it is in principle possible to determine which of the two macrosopically distinct states the system is in at any given time, by performing a noninvasive measurement that has a negligible effect on the subsequent dynamics legggarg-1 . This assumption is necessary because the traditional Leggett-Garg tests involve measurements at different times performed on the same system, and there is the possibility of direct disturbance of the system due to measurement. In this paper, we follow macro-bell-lg , and consider two systems (Figure 2). One assumes validity of (weak) macroscopic realism, but the additional assumption of noninvasive measureability is replaced by the assumption of locality. Specifically, for measurements at AA and BB that are spacelike separated events, one assumes the hidden variable value λi(A)\lambda_{i}^{(A)} is not changed by whether the measurement at BB takes place or not, and similarly for λj(B)\lambda_{j}^{(B)}.

Refer to caption
Figure 2: A Leggett-Garg test of macrorealism using bipartite systems. Two systems AA and BB prepared in an entangled state at time t1t_{1} undergo dynamical evolution according to local unitary operators, UA(t)U_{A}(t) and UB(t)U_{B}(t). At the later times t2t_{2} and t3t_{3}, each system is found to be in one or other of two macroscopically distinct states 1-1 or +1+1, and the outcomes are anti-correlated. One can infer the outcome for AA by measuring at BB. A possible evolution is shown by the dashed lines.

We consider measurements made at times t1t_{1} and t3t_{3} at site AA, and at times t2t_{2} and t4t_{4} at site BB, where t1<t2<t3<t4t_{1}<t_{2}<t_{3}<t_{4}. Let us denote the spin outcome SAS_{A} at AA for the measurement made at time tit_{i} by Si(A)S_{i}^{(A)}. The spin Sj(B)S_{j}^{(B)} is defined similarly, for site BB. With the assumptions of macroscopic realism and noninvasive measurability (locality), the Leggett-Garg-Bell inequality

2\displaystyle-2 \displaystyle\leq E(t1,t2)E(t1,t4)+E(t2,t3)+E(t3,t4)2\displaystyle E(t_{1},t_{2})-E(t_{1},t_{4})+E(t_{2},t_{3})+E(t_{3},t_{4})\leq 2
(3)

where E(ti,tj)=Si(A)Sj(B)E(t_{i},t_{j})=\langle S_{i}^{(A)}S_{j}^{(B)}\rangle can be shown to hold legggarg-1 . This inequality was derived as a Leggett-Garg inequality in Ref. legggarg-1 , and as a Leggett-Garg-Bell inequality, in Refs. macro-bell-lg ; weak-hybrid . Since the expectation values E(ti,tj)E(t_{i},t_{j}) of the inequality involve measurements made at diffferent sites, the inequality follows based on the assumption of locality, which justifies that E(ti,tj)=Si(A)Sj(B)=λi(A)λj(B)E(t_{i},t_{j})=\langle S_{i}^{(A)}S_{j}^{(B)}\rangle=\langle\lambda_{i}^{(A)}\lambda_{j}^{(B)}\rangle. We note that while we assume the time order t1<t2<t3<t4t_{1}<t_{2}<t_{3}<t_{4}, in fact because we have two localised sites, we need only take t1<t3t_{1}<t_{3} and t2<t4t_{2}<t_{4}.

A three-time Leggett-Garg-Bell inequality

3\displaystyle-3 \displaystyle\leq E(t1,t2)E(t1,t3)+E(t2,t3)1,\displaystyle E(t_{1},t_{2})-E(t_{1},t_{3})+E(t_{2},t_{3})\leq 1, (4)

where E(ti,tj)=Si(B)Sj(A)E(t_{i},t_{j})=\langle S_{i}^{(B)}S_{j}^{(A)}\rangle is also useful. This is based on the original three-time Leggett-Garg inequality, derived for the single-system set-up, in Refs. weak-solid-state-qubits ; jordan_kickedqndlg2-1 . Here, one assumes however that a measurement at time t1t_{1} at site AA does not affect the outcome at the later time t3t_{3} at the same site, which raises the issue of a disturbance due to the measurement at t1t_{1}. However, for correlated initial states, one may use that the result for S1(B)S_{1}^{(B)} can be inferred from a measurement of S1(A)S_{1}^{(A)}. Thus, assuming locality, one may deduce macro-bell-lg :

E(t1,t3)=S1(A)S3(A)=S1(B)S3(A)E(t_{1},t_{3})=\langle S_{1}^{(A)}S_{3}^{(A)}\rangle=-\langle S_{1}^{(B)}S_{3}^{(A)}\rangle (5)

We note that the no-signalling-in-time (NSIT) equality has been shown to give a stronger test of Leggett-Garg’s macrorealism NSTmunro-1 ; nst . However, it is more challenging to use this equality for a bipartite test.

II.3 Summary of assumptions

Since we are to show violation of the macroscopic Bell and Leggett-Garg-Bell inequalities, it is important to summarise what is the weakest (i.e. minimal) assumption needed to derive the inequalities. The Bell inequalities (2) can be derived from the minimal assumption of macroscopic local causality (MLC), as in Section 1.A.2. In this paper, however, we use the terms macroscopic local realism (MLR) and macroscopic local causality (MLC) interchangeably. The Leggett-Garg-Bell inequalities (3) and (4) on the other hand are derived assuming weak macroscopic realism, combined with macroscopic noninvasiveness (NIM) of the measurement, which for two sites is justified by macroscopic locality (ML).

III Violating the macroscopic Bell inequalities

We consider the system of Figure 1 prepared in the initial “cat” state cat-bell-wang-1

|ψBell=𝒩(|αa|βb|αa|βb)|\psi_{Bell}\rangle=\mathcal{N}(|\alpha\rangle_{a}|-\beta\rangle_{b}-|-\alpha\rangle_{a}|\beta\rangle_{b}) (6)

Here |α|\alpha\rangle and |β|\beta\rangle are coherent states for two modes, labelled aa and bb, and for convenience we assume α\alpha, β\beta to be real and positive. We will ultimately take α\alpha, β\beta\rightarrow\infty. The normalisation constant is 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}. In the limit where α\alpha, β\beta are large, we note that |αa|\alpha\rangle_{a} and |αa|-\alpha\rangle_{a} become orthogonal, and similarly |βb|\beta\rangle_{b} and |βb|-\beta\rangle_{b}, so that 𝒩12\mathcal{N}\rightarrow\frac{1}{\sqrt{2}}. In this limit, we may write the cat state as a macroscopic two-qubit state

|ψBell=12(|+a|b|a|+b)|\psi_{Bell}\rangle=\frac{1}{\sqrt{2}}(|+\rangle_{a}|-\rangle_{b}-|-\rangle_{a}|+\rangle_{b}) (7)

Here |+a|+\rangle_{a} and |a|-\rangle_{a} are orthogonal states giving an outcome of +1+1 and 1-1 respectively, for a measurement M^(A)\hat{M}^{(A)} of the sign SAS_{A} of the coherent amplitude, α\alpha. Similarly, |+b|+\rangle_{b} and |b|-\rangle_{b} are orthogonal states, with outcome +1+1 and 1-1 for the measurement M^(B)\hat{M}^{(B)} of the sign SBS_{B} of the amplitude β\beta. The two modes aa and bb are spatially separated, being located at sites AA and BB, respectively. The sign SS of the coherent amplitude is taken to be the value of the “spin” for each mode (site).

III.1 Bell violations using unitary local transformations

At locations AA and BB, we assume local unitary transformations UA(ta)U_{A}(t_{a}) and UB(tb)U_{B}(t_{b}) take place, as in Figure 1. States |±a|\pm\rangle_{a} and |±b|\pm\rangle_{b} are transformed after a time tat_{a} and tbt_{b} according to (at site AA)

UA(t)a|+a\displaystyle U_{A}(t{}_{a})|+\rangle_{a} =\displaystyle= a|+a+i1a2|a\displaystyle a|+\rangle_{a}+i\sqrt{1-a^{2}}|-\rangle_{a}
UA(t)a|a\displaystyle U_{A}(t{}_{a})|-\rangle_{a} =\displaystyle= a|a+i1a2|+a\displaystyle a|-\rangle_{a}+i\sqrt{1-a^{2}}|+\rangle_{a} (8)

and (at site BB)

UB(t)b|+b\displaystyle U_{B}(t{}_{b})|+\rangle_{b} =\displaystyle= b|+b+i1b2|b\displaystyle b|+\rangle_{b}+i\sqrt{1-b^{2}}|-\rangle_{b}
UB(t)b|b\displaystyle U_{B}(t{}_{b})|-\rangle_{b} =\displaystyle= b|b+i1b2|+b\displaystyle b|-\rangle_{b}+i\sqrt{1-b^{2}}|+\rangle_{b} (9)

For a system prepared in the superposition |ψBell|\psi_{Bell}\rangle of eqn. (7), the system will evolve into the final state

|ψf\displaystyle|\psi_{f}\rangle =12(UA|+aUB|bUA|aUB|+b)\displaystyle=\frac{1}{\sqrt{2}}(U_{A}|+\rangle_{a}U_{B}|-\rangle_{b}-U_{A}|-\rangle_{a}U_{B}|+\rangle_{b})
=12{(ab+a¯b¯)(|+a|b|a|+b)\displaystyle=\frac{1}{\sqrt{2}}\{(ab+\bar{a}\bar{b})(|+\rangle_{a}|-\rangle_{b}-|-\rangle_{a}|+\rangle_{b})
+i(a¯bab¯)(|+a|+b|a|b)}\displaystyle\thinspace\thinspace{\color[rgb]{0,0,1}{\normalcolor+}}i(\bar{a}b-a\bar{b})(|+\rangle_{a}|+\rangle_{b}-|-\rangle_{a}|-\rangle_{b})\} (10)

where we have abbreviated a¯=1a2\bar{a}=\sqrt{1-a^{2}} and b¯=1b2.\bar{b}=\sqrt{1-b^{2}}. Using that the states |±|\pm\rangle are orthogonal, and defining the measurement S^\hat{S} which measures the sign SS (which we label as its “spin”), according to S^(A)|±a=±|±a\hat{S}^{(A)}|\pm\rangle_{a}=\pm|\pm\rangle_{a} and S^(B)|±b=±|±b\hat{S}^{(B)}|\pm\rangle_{b}=\pm|\pm\rangle_{b}, we find that the expectation value E(ta,tb)E(t_{a},t_{b}) of the spin product S^(A)S^(B)\hat{S}^{(A)}\hat{S}^{(B)} is

E(ta,tb)=|a¯bab¯|2|ab+a¯b¯|2E(t_{a},t_{b})=|\bar{a}b-a\bar{b}|^{2}-|ab+\bar{a}\bar{b}|^{2} (11)

For the traditional Bell experiment where the transformation is achieved using a polarising beam splitter, we have a=cosθa=\cos\theta and b=cosϕb=\cos\phi, and

E(θ,ϕ)=cos2(θϕ)E(\theta,\phi)=-\cos 2(\theta-\phi) (12)

It is known that the choice of angles θ=0\theta=0, ϕ=π/8\phi=\pi/8, θ=π/4\theta^{\prime}=\pi/4 and ϕ=3π/8\phi^{\prime}=3\pi/8 in the Bell inequality (1) will lead to B=22B=-2\sqrt{2} which violates the inequality.

III.2 Unitary transformations via nonlinear media

Now we consider how to realise the transformations (8) and (9). We seek a unitary transformation Uθ(A)UA(θ)U_{\theta}^{(A)}\equiv U_{A}(\theta) such that at site AA the transformation

UA(θ)|α\displaystyle U_{A}(\theta)|\alpha\rangle =\displaystyle= cosθ|α+isinθ|α\displaystyle\cos\theta|\alpha\rangle+i\sin\theta|-\alpha\rangle
UA(θ)|α\displaystyle U_{A}(\theta)|-\alpha\rangle =\displaystyle= cosθ|α+isinθ|α\displaystyle\cos\theta|-\alpha\rangle+i\sin\theta|\alpha\rangle (13)

holds, for the particular angle choices θ=0\theta=0 and θ=π/4\theta=\pi/4. At site BB, we seek a transformation such that

UB(ϕ)|β\displaystyle U_{B}(\phi)|\beta\rangle =\displaystyle= cosϕ|β+isinϕ|β\displaystyle\cos\phi|\beta\rangle+i\sin\phi|-\beta\rangle
UB(ϕ)|β\displaystyle U_{B}(\phi)|-\beta\rangle =\displaystyle= cosϕ|β+isinϕ|β\displaystyle\cos\phi|-\beta\rangle+i\sin\phi|\beta\rangle (14)

for the angle choices ϕ=π/8\phi=\pi/8, and ϕ=3π/8\phi=3\pi/8. We drop subscripts aa and bb, where the meaning is clear.

These transformations can be achieved by interacting each single mode system with a nonlinear medium at the given site, where the interaction Hamiltonian is HNL=Ωn^kH_{NL}=\Omega\hat{n}^{k} (k>2k>2 and kk even) manushan-cat-lg ; yurke-stoler-1 ; wrigth-walls-gar-1 ; macro-bell-lg . Here n^\hat{n} is the mode number operator and Ω\Omega is the nonlinear coefficient. We choose units such that =1\hbar=1. The interaction takes place independently for each mode aa and bb, and we denote the respective Hamiltonians as HNL(A)H_{NL}^{(A)} and HNL(B)H_{NL}^{(B)}. Thus we define local interactions at each site

HNL(A)=Ωn^ak,HNl(B)=Ωn^bkH_{NL}^{(A)}=\Omega\hat{n}_{a}^{k},\ \ H_{Nl}^{(B)}=\Omega\hat{n}_{b}^{k} (15)

The boson destruction mode operators for modes aa and bb are denoted by a^\hat{a} and b^\hat{b}, respectively, and the number operators are 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}.

Refer to caption
Figure 3: An initial coherent state |α|\alpha\rangle evolving according to the Hamltonian HNL(A)H_{NL}^{(A)}, k=4k=4, α=3\alpha=3. The Figure shows contour plots of the Q function Q(xA,pA)Q(x_{A},p_{A}) for the states created at a time tt. For the times shown, a superposition of the states |α|-\alpha\rangle and |α|\alpha\rangle is formed.
Refer to caption
Figure 4: The detailed dynamics of the creation of the superposition state shown in Figure 4. The system is not a superposition of states |α|-\alpha\rangle and |α|\alpha\rangle at all times tt. Here, α=4\alpha=4.

One may solve for the dynamics given by HNLH_{NL}. If the system AA is prepared in a coherent state |αa|\alpha\rangle_{a}, the state after time tt is yurke-stoler-1

|α,ta\displaystyle\left|\alpha,t_{a}\right\rangle =\displaystyle= UA(ta)|αa\displaystyle U_{A}(t_{a})|\alpha\rangle_{a} (16)
=\displaystyle= exp[|α|22]n=0αnexp(iϕn)n!|na\displaystyle\exp[-\frac{\left|\alpha\right|^{2}}{2}]\underset{n=0}{\sum^{\infty}}\frac{\alpha^{n}\exp(-i\phi_{n})}{\sqrt{n!}}\left|n\right\rangle_{a}

where UA(ta)=eiHNL(A)ta/U_{A}(t_{a})=e^{-iH_{NL}^{(A)}t_{a}/\hbar} and ϕn=Ωtank\phi_{n}=\varOmega t_{a}n^{k}. For an initial coherent state |βb|\beta\rangle_{b} at BB, the state after a time tt is given similarly, as |β,tb=UB(tb)|βb\left|\beta,t_{b}\right\rangle=U_{B}(t_{b})|\beta\rangle_{b} where UB(tb)=eiHNL(B)tb/U_{B}(t_{b})=e^{-iH_{NL}^{(B)}t_{b}/\hbar}. These solutions are well known manushan-cat-lg ; yurke-stoler-1 and are depicted graphically in Figures 3 and 4 in terms of the QQ function, defined as Q(xA,pA)=1πα0|ρ|α0Q(x_{A},p_{A})=\frac{1}{\pi}\langle\alpha_{0}|\rho|\alpha_{0}\rangle Husimi-Q-1 . Here α0=xA+ipA\alpha_{0}=x_{A}+ip_{A} and |α0|\alpha_{0}\rangle is a coherent state. At certain intervals of time, the system evolves into a superposition of the macroscopically distinct coherent states |α|-\alpha\rangle and |α|\alpha\rangle (Figure 4).

Suppose at time ta=tb=0t_{a}=t_{b}=0, the systems are in the initial states |α|\alpha\rangle and |β|\beta\rangle. After a time tb=π/4Ωt_{b}=\pi/4\Omega, the system at BB is in the state UB(π/4Ω)|βU_{B}(\pi/4\Omega)|\beta\rangle (second picture of Figure 3). This gives the transformation of eq. (14) with ϕ=π/8\phi=\pi/8 manushan-cat-lg . We write the state at this time as

Uπ/8(B)|β\displaystyle U_{\pi/8}^{(B)}|\beta\rangle =\displaystyle= eiπ/8{(cosπ/8|β+isinπ/8|β}\displaystyle e^{-i\pi/8}\{(\cos\pi/8|\beta\rangle+i\sin\pi/8|-\beta\rangle\}

After a time ta=π/2Ωt_{a}=\pi/2\Omega, the state is UA(π/2Ω)|αU_{A}(\pi/2\Omega)|\alpha\rangle which corresponds to the transformation eq. (13) with θ=π/4\theta=\pi/4. We write the state at ta=π/2Ωt_{a}=\pi/2\Omega as manushan-cat-lg

Uπ/4(A)|α\displaystyle U_{\pi/4}^{(A)}|\alpha\rangle =\displaystyle= eiπ/42{|α+eiπ/2|α}\displaystyle\frac{e^{-i\pi/4}}{\sqrt{2}}\{|\alpha\rangle+e^{i\pi/2}|-\alpha\rangle\} (18)

To obtain the transformation (14) with ϕ=3π/8\phi=3\pi/8, we consider the state at time tb=3π/4Ωt_{b}=3\pi/4\Omega manushan-cat-lg , which is given by UB(3π/4Ω)|βU_{B}(3\pi/4\Omega)|\beta\rangle. The state at the time tb=3π/4Ωt_{b}=3\pi/4\Omega is

U3π/8(B)|β\displaystyle U_{3\pi/8}^{(B)}|\beta\rangle =\displaystyle= eiπ/8{(cos3π/8|β+isin3π/8|β}\displaystyle e^{i\pi/8}\{(\cos 3\pi/8|\beta\rangle+i\sin 3\pi/8|-\beta\rangle\}

These transformations are shown graphically in Figure 3. The transformations at each site AA and BB are summarised as

UA(ta)=eiHNL(A)ta/,UB(tb)=eiHNL(B)tb/U_{A}(t_{a})=e^{-iH_{NL}^{(A)}t_{a}/\hbar},\ \ U_{B}(t_{b})=e^{-iH_{NL}^{(B)}t_{b}/\hbar} (20)

where ta=0t_{a}=0, tb=π/4Ωt_{b}=\pi/4\Omega, ta=π/2Ωt^{\prime}_{a}=\pi/2\Omega and tb=3π/4Ωt_{b}^{\prime}=3\pi/4\Omega. Noting from above, UB(π/4Ω)=Uπ/8(B)U_{B}(\pi/4\Omega)=U_{\pi/8}^{(B)}, UA(π/2Ω)=Uπ/4(A)U_{A}(\pi/2\Omega)=U_{\pi/4}^{(A)} and UB(3π/4Ω)=U3π/8(B)U_{B}(3\pi/4\Omega)=U_{3\pi/8}^{(B)}.

The Bell test of Section III.A requires a measurement of the spin S^\hat{S} for the states after the unitary rotations. As we have seen above, the unitary parts of the measurements M^θ(A)\hat{M}_{\theta}^{(A)} and M^ϕ(B)\hat{M}_{\phi}^{(B)} are to be made by interacting with the nonlinear medium at each site. This creates the superpositions (LABEL:eq:state3-20) of the coherent states |α|-\alpha\rangle and |α|\alpha\rangle (|β|-\beta\rangle and |β|\beta\rangle) which are rotated relative to the initial state |α|\alpha\rangle (|β|\beta\rangle). The initial states determine the orientiation of real axis (i.e. the phase direction). The coherent states |α|-\alpha\rangle and |α|\alpha\rangle (|β|-\beta\rangle and |β|\beta\rangle) are hence the basis states for a particular pointer measurement, M^(A)\hat{M}^{(A)} (M^(B)\hat{M}^{(B)}) that determines the sign of the coherent amplitude i.e. the value of the macroscopic qubit. The pointer measurements are hence measurements of the quadrature phase amplitudes yurke-stoler-1

X^A=12(a^+a^),X^B=12(b^+b^)\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}),\ \ \hat{X}_{B}=\frac{1}{\sqrt{2}}(\hat{b}+\hat{b}^{\dagger}) (21)

on the evolved states at each site. The orientation of the axis that defines quadratures is aligned with the orientation of the initial coherent states |α|\alpha\rangle (|β|\beta\rangle) in phase space. This is set in the experiment by a local oscillator. Finally, such a measurement involves a coupling to detectors. Let us denote the results of the pointer measurements X^A\hat{X}_{A} and X^B\hat{X}_{B} by XAX_{A} and XBX_{B} respectively. The final spin outcome of the measurement M^θ(A)\hat{M}_{\theta}^{(A)} is taken to be SA=+1S_{A}=+1 if XA>0X_{A}>0, and 1-1 otherwise. Similarly, the outcome of the measurement M^ϕ(B)\hat{M}_{\phi}^{(B)} is SB=+1S_{B}=+1 if XB>0X_{B}>0, and 1-1 otherwise.

III.3 Macroscopic Bell violations

At each site AA and BB, the system of Figure 1 prepared in the cat state (6) undergoes unitary transformations UAU_{A} and UBU_{B}, given by (20). There is the choice of two rotation angles, corresponding to two different times, at each site. These give transformations of the form (13) and (14).

A violation of the Bell inequality (1) is obtained if we select angles θ=0\theta=0 and θ=π/4\theta=\pi/4 at site AA, and ϕ=π/8\phi=\pi/8 and ϕ=3π/8\phi=3\pi/8 at site BB (Figure 1). Therefore, in order to violate the Bell inequality (2), we select ta=0t_{a}=0 (which gives rotation θ=0\theta=0) and ta=π/2Ωt^{\prime}_{a}=\pi/2\Omega (which gives rotation θ=π/4\theta=\pi/4). Similarly at site B, we select tb=π/4Ωt_{b}=\pi/4\Omega (rotation ϕ=π/8\phi=\pi/8) and tb=3π/(4Ω)t^{\prime}_{b}=3\pi/(4\Omega) (rotation ϕ=3π/8\phi=3\pi/8). Using the prediction E(ta,tb)=E(θ,ϕ)E(t_{a},t_{b})=E(\theta,\phi) given by (12) for orthogonal states undergoing the unitary transformations (8) and (9), we see that this choice of time-settings will give a violation of the Bell inequality (2).

Refer to caption
Figure 5: Distributions giving macroscopic Bell violations: Contour plots of P(XA,XB)P(X_{A},X_{B}) for the system depicted in Figure 1 with an initial entangled state |ψBell|\psi_{Bell}\rangle (6) and unitary transformations UAU_{A} and UBU_{B} (20), for measurement settings tat_{a} and tbt_{b}. Here, α=β=3\alpha=\beta=3 and Ω=1\Omega=1. For each setting θ\theta and ϕ\phi given by (tat_{a}, tbt_{b}), the outcomes SAS_{A} and SBS_{B} for M^(A)\hat{M}^{(A)} and M^(B)\hat{M}^{(B)} are +1+1 or 1-1 if the measurement of X^A\hat{X}_{A} or X^B\hat{X}_{B} yields a positive or negative value.

To account for the nonorthogonality due to finite amplitudes α\alpha and β\beta, we evaluate the distributions P(XA,XB)P(X_{A},X_{B}) for the outcomes of the joint measurement of the amplitudes X^A\hat{X}_{A} and X^B\hat{X}_{B} (eqn (21), for each of the time-settings. This constitutes the prediction for the final pointer measurements. The final state after the unitary transformations is

|ψf\displaystyle|\psi_{f}\rangle =\displaystyle= 𝒩UA(ta)UB(tb)(|α|β|α|β)\displaystyle\mathcal{N}\thinspace U_{A}(t_{a})U_{B}(t_{b})\Bigr{(}|\alpha\rangle|-\beta\rangle-|-\alpha\rangle|\beta\rangle\Bigl{)}

where the unitary operators are given by (20) and 𝒩\mathcal{N} is the normalisation constant. The probability is

P(XA,XB)=|XA|XB|ψf|2P(X_{A},X_{B})=|\langle X_{A}|\langle X_{B}|\psi_{f}\rangle|^{2} (23)

where |XA|X_{A}\rangle, |XB|X_{B}\rangle are eigenstates of XAX_{A} and XBX_{B} respectively. Details of calculation are given in the Appendix. Integrating gives the results needed for the evaluation of the Bell inequality term, BB, of eqn (2). We use E(ta,tb)=P+++PP+P+E(t_{a},t_{b})=P_{++}+P_{--}-P_{+-}-P_{-+} where

P++\displaystyle P_{++} =\displaystyle= 00P(XA,XB)𝑑XA𝑑XB\displaystyle\int_{0}^{\infty}\int_{0}^{\infty}P(X_{A},X_{B})dX_{A}dX_{B} (24)

and PP_{--}, P+P_{-+} and P+P_{+-} are defined similarly for the remaining three quadrants.

The results for the four choices of time settings are shown in Figures 5 and 6. For α,β1.5\alpha,\beta\geq 1.5, the overlap of the coherent states |α|\alpha\rangle and |α|-\alpha\rangle (and |β|\beta\rangle and |β|-\beta\rangle) is small and there is agreement with the calculations based on the assumption of orthogonality of the states. For each time setting, the results of the measurements S(A)S^{(A)} and S(B)S^{(B)} have two possible values, +1+1 and 1-1, and there are four joint outcomes. In Figure 5, the plots of P(XA,XB)P(X_{A},X_{B}) show four distinct Gaussian peaks corresponding to the probabilities of these joint outcomes, the peaks becomes macroscopically separated for large α\alpha and β\beta. We thus verify that maximal violations of the Bell inequality (2) (|B|=22|B|=2\sqrt{2}) are indeed predicted for the macroscopic qubits, where measurements distinguish macroscopically distinct states, as α\alpha, β\beta\rightarrow\infty.

Refer to caption
Refer to caption
Figure 6: The violation of the macroscopic Bell inequality (2) for the system depicted in Figure 1, as described in Figure 5. We show (top) BB versus β\beta where α=β.\alpha=\beta. In the lower plot, β\beta is varied for fixed α\alpha. A violation is obtained when |B|>2|B|>2.

III.4 Dynamics of the unitary part of the measurement

In the set-up of Figure 1, two spacelike separated systems display failure of macroscopic local realism. As the measurements take place, the systems evolve dynamically over an interval of time. This is a realistic model because the analyser measurements involve interactions over a finite time duration. In order to gain insight into how the macroscopic Bell violations arise over the course of the dynamics, we present plots in terms of the Q function.

The Husimi Q function is defined uniquely as a positive function for a two-mode quantum state ρ\rho by Q(α0,β0)=1π2α0|β0|ρ|β0|α0Q(\alpha_{0},\beta_{0})=\frac{1}{\pi^{2}}\langle\alpha_{0}|\langle\beta_{0}|\rho|\beta_{0}\rangle|\alpha_{0}\rangle where |α0|\alpha_{0}\rangle and |β0|\beta_{0}\rangle are coherent states for the modes Husimi-Q-1 . It is a function of four real variables xAx_{A}, pAp_{A}, xBx_{B} and pBp_{B}, where α0=xA+ipA\alpha_{0}=x_{A}+ip_{A} and β0=xB+ipB\beta_{0}=x_{B}+ip_{B}. The Q function is a quasi-probability distribution, corresponding to an anti-normal ordering of moments, and therefore (different to a positive Wigner function) does not provide a probability distribution for the actual outcomes XAX_{A}, PA,P_{A}, XBX_{B}, PBP_{B} of the measurements X^A=12(a^+a^),X^B=12(b^+b^)\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}),\hat{X}_{B}={\color[rgb]{1,0,0}{\color[rgb]{0,0,1}{\color[rgb]{0,0,0}\frac{1}{\sqrt{2}}}}}(\hat{b}+\hat{b}^{\dagger}), P^A=1i2(a^a^),P^B=1i2(b^b^)\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}),\hat{P}_{B}={\color[rgb]{1,0,0}{\color[rgb]{0,0,1}{\color[rgb]{0,0,0}\frac{1}{i\sqrt{2}}}}}(\hat{b}-\hat{b}^{\dagger}). However, the QQ function provides at any given time the correct probability distribution for the outcomes of the macroscopically separated spins Sj(B)S_{j}^{(B)} and Si(A)S_{i}^{(A)}, as we confirm below.

Refer to caption
Refer to caption
Figure 7: Top: Contour plots of the QQ function marginal Q(xA,xB)Q(x_{A},x_{B}) for the states created at the times relevant for the violation of the macroscopic Bell inequality (2), as in Figure 5. The system is prepared at ta=tb=0t_{a}=t_{b}=0 in the entangled Bell cat state |ψBell|\psi_{Bell}\rangle (6). Lower: The corresponding plots if the system is prepared in the non-entangled mixture ρmix\rho_{mix} (LABEL:eq:mixqstate), which does not violate the Bell inequality. Here α=β=3\alpha=\beta=3 and Ω=1\Omega=1.
Refer to caption
Figure 8: The two-state nature of the systems created at the specific times given in Figures 5 and 7 is evident on examining the contour plots of the marginal Q(xA,pA)Q(x_{A},p_{A}). The plots for the Bell cat state and the mixed state ρmix\rho_{mix} are indistinguishable.

Assuming the system is prepared in the cat state (6), the state of the system after interaction times tat_{a} at site AA and tbt_{b} at site BB is given by eq. (LABEL:eq:234-1-1-1). The Q function at the time tat_{a} and tbt_{b} is Q(xA,xB,pA,pB)=1π2|β0|α0|ψf|2Q(x_{A},x_{B},p_{A},p_{B})=\frac{1}{\pi^{2}}|\langle\beta_{0}|\langle\alpha_{0}|\psi_{f}\rangle|^{2}. The marginal for the two XX quadratures, given by

Q(xA,xB)\displaystyle Q(x_{A},x_{B}) =\displaystyle= 𝑑pA𝑑pBQ(xA,xB,pA,pB)\displaystyle\int dp_{A}dp_{B}Q(x_{A},x_{B},p_{A},p_{B}) (25)

One can also evaluate the marginals for each system e.g. Q(xA,pAQ(x_{A},p_{A}) by integrating over xBx_{B} and pBp_{B}.

In Figures 7 and 8, we plot the marginals Q(xA,xB)Q(x_{A},x_{B}) and Q(xA,pA)Q(x_{A},p_{A}) corresponding to the states created after the interaction times tat_{a} and tbt_{b}, as in Figure 5, where there is a violation of the Bell inequality (2). As expected, there is similarity in the macroscopic limit of the Q function Q(xA,xB)Q(x_{A},x_{B}) with the actual distributions P(XA,XB)P(X_{A},X_{B}), of Figure 5. The additional noise of order \hbar (scaled to order 1\sim 1 in the plots) which is characteristic of the Q function does not change the probabilities for the macroscopically distinct outcomes, SAS_{A} and SBS_{B}.

Refer to caption
Figure 9: The entangled cat state |ψBell|\psi_{Bell}\rangle (6) evolves while being measured at the different sites. We show the contour plots of Q(xA,xB)Q(x_{A},x_{B}) corresponding to the 44 joint measurements leading to violation of the Bell inequality (2), as in Figures 5 and 7. The top sequence shows the dynamics for the measurements with settings (ta=0,tb=π/4)(t_{a}=0,t_{b}=\pi/4) and (ta=0,tb=3π/4)(t_{a}=0,t_{b}=3\pi/4). The centre and lower sequences correspond to settings (ta=π/2,tb=3π/4)(t_{a}=\pi/2,t_{b}=3\pi/4) and (ta=π/2,tb=π/4)(t_{a}=\pi/2,t_{b}=\pi/4), respectively. Here α=β=3\alpha=\beta=3 and Ω=1\Omega=1.
Refer to caption
Figure 10: The plots are as for Figure 9, but here the system evolves from a nonentangled state ρmix\rho_{mix}. There is no violation of the Bell inequality. The plots with ta=0t_{a}=0 (top) involving a unitary rotation at only one site are indistinguishable from those of Figure 9. The centre and lower plots involving transformations at two sites show a macroscopic difference at time ta=π/(2Ω)t_{a}=\pi/(2\Omega).

To investigate the dynamics leading to the violation of macroscopic local realism (MLR), we give a comparison with the predictions if the system is initially prepared in the non-entangled mixed state

ρmix\displaystyle\rho_{mix} =\displaystyle= 12(|α|βα|β|+|α|βα|β|)\displaystyle\frac{1}{2}(|\alpha\rangle|-\beta\rangle\langle\alpha|\langle-\beta|+|-\alpha\rangle|\beta\rangle\langle-\alpha|\langle\beta|)

The mixture evolves to

ρmix(ta,tb)\displaystyle\rho_{mix}(t_{a},t_{b}) =\displaystyle= UB(tb)UA(ta)ρmixUA(ta)UB(tb)\displaystyle U_{B}(t_{b})U_{A}(t_{a})\rho_{mix}U_{A}^{\dagger}(t_{a})U_{B}^{\dagger}(t_{b})

We plot the Q function of the mixed state ρmix\rho_{mix}, eqn. (LABEL:eq:mixunit), in Figure 7 (lower) for the choice of time-settings that for the state |ψBell|\psi_{Bell}\rangle lead to a violation of the Bell inequality (top). The mixed state has the interpretation that the system is in one or other of the quantum states with a definite outcome for SAS_{A} and SBS_{B} at a time t=0t=0, as α\alpha, β\beta\rightarrow\infty. Consistent with that interpretation, the dynamics does not lead to a violation of the Bell inequality. We see that the Q functions for the initial states (|ψBell|\psi_{Bell}\rangle and ρmix\rho_{mix}) at t=0t=0 are indistinguishable. In fact, calculation shows that there is a microscopic difference 𝒞\mathcal{C}, of order e|α|2|β|2e^{-|\alpha|^{2}-|\beta|^{2}}, between these two Q functions. We also see that the Q functions arising from initial states (|ψBell|\psi_{Bell}\rangle and ρmix\rho_{mix}) are indistinguishable if either tat_{a} or tbt_{b} is zero. However, for the time settings where there is evolution (rotation) at both the sites, the final outcomes are macroscopically different.

The timescales of the dynamics leading to the violation of macroscopic local realism (MLR) can be visualised if we calculate the full dynamics associated with the unitary measurements (rotations) (Figures 9 and 10). The predictions for the joint distributions P(XA,XB)P(X_{A},X_{B}) depend only on the absolute value of the times tat_{a} and tbt_{b} of interaction with the measurement apparatus at each site. The same violation of the Bell inequalities can thus be achieved in different ways relative to a shared clock. For example, one may measure system AA first, and subsequently measure system BB; or vice versa. Alternatively, one may measure by evolving the systems simultaneously. In Figures 9 and 10, we choose to evolve sequentially, with BB first. On comparing the sequences of Q functions for the entangled state and for the mixture ρmix\rho_{mix} in Figures 9 and 10, we observe the transition from indistinguishable Q functions at ta=tb=0t_{a}=t_{b}=0 though to a small but noticable difference at times ta=π/8Ωt_{a}=\pi/8\Omega and tb=π/4Ωt_{b}=\pi/4\Omega. Finally, at times ta=π/2Ωt_{a}=\pi/2\Omega and tb=π/4Ωt_{b}=\pi/4\Omega, the difference has become macroscopic.

To understand why there is little distinction between the results for |ψBell|\psi_{Bell}\rangle and ρmix\rho_{mix} if the unitary evolution acts on the state at one site only (ta=0t_{a}=0), but not where for ta0t_{a}\neq 0 and tb0t_{b}\neq 0, we calculate the probabilities P(XA,XB)P(X_{A},X_{B}) for each case, in the Appendix. The difference arises from quantum interference terms damped by exp(xA22|α|2)\exp(-x_{A}^{2}-2|\alpha|^{2}) in the first case but which are not damped for the entangled state for certain values of xAx_{A}, in the second case. The effect for large α\alpha and β\beta is in direct analogy with the calculations given for the two-qubit Bell experiment in Section III.A, as expected since these calculations rely on the orthogonality of the states forming the qubit.

The plots in Figure 8 show the marginals Q(xA,pA)Q(x_{A},p_{A}) and Q(xB,pB)Q(x_{B},p_{B}) at each site. These plots highlight the two-state nature of the systems immediately after the unitary rotations UA(ta)U_{A}(t_{a}) and UB(tb)U_{B}(t_{b}) for the choice of measurement settings required for the Bell violation. Identical two-state plots are obtained for the coarsely selected times given in Figure 9 that show the dynamics during those measurements. However, as seen from Figure 4, over very much shorter time scales, the unitary dynamics involves a continuous transition, and the systems cannot be regarded as two-state systems at all times.

III.5 Conclusions

It is clear from the summary given in Section II.C that the violations predicted for the macroscopic Bell inequality (1) and (2) imply failure of deterministic macroscopic (local) realism (dMR) and macroscopic local causality (MLC), which we collectively refer to as macroscopic local realism (MLR). We explain that there is no inconsistency however with the assumption of weak macroscopic realism (wMR), as introduced in the Sections I and II.

The peaks of the Q function in the Figures give the values for the probabilities of the macroscopically distinct outcomes (+1+1 and 1-1) for the spins, SAS_{A} and SBS_{B}. The description is of an entire ensemble, not of an individual system. Therefore the negligible difference of order 𝒞e|α|2\mathcal{C}\sim e^{-|\alpha|^{2}} between the Q functions of the classical mixture ρmix\rho_{mix} (eqn (LABEL:eq:mixqstate)) and the superposition |ψBell|\psi_{Bell}\rangle (eqn (6)) refers to the average over many identically prepared systems, not to any individual system at any given time.

In what we refer to here as the weak macroscopic realism (wMR) model, an interpretation is given that goes beyond this. Consider the cat-state |ψBell|\psi_{Bell}\rangle, which for large α,\alpha,β\beta is a superposition of two states with definite and macroscopically distinct outcomes for the pointer measurement M^(A)\hat{M}^{(A)} (M^(B)\hat{M}^{(B)}) of the spins SAS_{A} (SBS_{B}). Similar superpositions (10) are created after the unitary rotations UU at each site, as evident in Figures 7 and 9. In analogy with the polarising beam splitter or analyser in a Bell experiment, UU brings about a change of basis, in preparation for a pointer measurement. The Q function for such superposition states |ψpointer|\psi_{pointer}\rangle gives macroscopically distinct peaks corresponding to outcomes SAS_{A} (SBS_{B}) for M^(A)\hat{M}^{(A)} (M^(B)\hat{M}^{(B)}).

The wMR interpretation is that immediately prior to the measurement M^\hat{M} at each site, the local system was in one or other of two states (referred to as φ1\varphi_{1} and φ2\varphi_{2}) predetermined to give one or other of the macroscopically distinguishable outcomes +1+1 or 1-1 respectively. This assumption is defined in parallel with that of macroscopic locality of the pointer (MLP): that the measurement at BB does not affect the (macroscopic) spin outcome at AA (and vice versa). In this model, the violation of macrorealism and of macroscopic local realism arises over the course of the unitary dynamics. The distinction between predictions for the initial Bell superposition and a classical mixture of the two pointer eigenstates is quantified by terms 𝒞\mathcal{C} of order e|α|2\hbar e^{-|\alpha|^{2}}. The violations of macro-realism and MLR emerge with the amplification of the effect of 𝒞\mathcal{C}, over the course of the unitary dynamics corresponding to the rotation of basis.

In this interpretation, it is not assumed that the states φ1\varphi_{1} and φ2\varphi_{2} correspond to |α|\alpha\rangle and |α|-\alpha\rangle, nor that φ1\varphi_{1} and φ2\varphi_{2} correspond to quantum states (indeed, they cannot, as shown in Section V). Nonetheless, the wMR model asserts the predetermination of the macroscopic prediction.

IV Leggett-Garg tests of macro-realism and macroscopic local realism

In this section, we propose tests of macrorealism and of macroscopic local realism using cat states and the Leggett-Garg inequalities. For measurements of spin Sj(A)S_{j}^{(A)} made on a single system AA at consecutive times t1<t2<t3t_{1}<t_{2}<t_{3}, macrorealism implies the Leggett-Garg inequality weak-solid-state-qubits ; jordan_kickedqndlg2-1

S1(A)S2(A)+S2(A)S3(A)S1(A)S3(A)1{\color[rgb]{0,0,0}{\color[rgb]{0,0,0}\langle S_{1}^{(A)}S_{2}^{(A)}\rangle+\langle S_{2}^{(A)}S_{3}^{(A)}\rangle-\langle S_{1}^{(A)}S_{3}^{(A)}\rangle}\leq 1} (28)

IV.1 Leggett-Garg test with a single system

We first outline the approach using a single system, but as applied to the cat states. Here, we follow an analysis similar to that given by Ref. manushan-cat-lg . The system AA is prepared at time t1=0t_{1}=0 in a coherent state |α|\alpha\rangle. The spin S1S_{1} if measured at this time gives the result +1+1. The result however is known deterministically, from the preparation. Following preparation, there is evolution according to the Hamiltonian HNL(A)H_{NL}^{(A)} of eqn (15), for a time t2=π/4Ωt_{2}=\pi/4\Omega. This gives the state

Uπ/8|α\displaystyle U_{\pi/8}|\alpha\rangle =\displaystyle= eiπ/8{(cosπ/8|α+isinπ/8|α}\displaystyle e^{-i\pi/8}\{(\cos\pi/8|\alpha\rangle+i\sin\pi/8|-\alpha\rangle\} (29)

At the time t2t_{2}, after the rotation, we may choose to measure the value of the spin S2S_{2}. This gives S1S2=cos(π/4)\langle S_{1}S_{2}\rangle=\cos(\pi/4). If not measured, the system continues to evolve according to HNLH_{NL} until the time t3=π/2Ωt_{3}=\pi/2\Omega. The state of the system at t3t_{3} is

Uπ/4|α=eiπ/42{|α+eiπ/2|α}U_{\pi/4}|\alpha\rangle=\frac{e^{-i\pi/4}}{\sqrt{2}}\{|\alpha\rangle+e^{i\pi/2}|-\alpha\rangle\} (30)

which implies S1S3=0\langle S_{1}S_{3}\rangle=0.

To test the Leggett-Garg inequality, one requires to evaluate S2S3\langle S_{2}S_{3}\rangle. Here, one assumes macrorealism. We follow the usual approach, and suppose that an ideal measurement M^\hat{M} of S2S_{2} takes place at time t2t_{2}, which would give a result of either +1+1 or 1-1. If the result is ±1\pm 1, then one assumes the system was in the state |±α|\pm\alpha\rangle at the time t2t_{2}, based on the assumption that the measurement did not disturb the system, and noting that tomography of the (average) state conditional on the result ±1\pm 1 would be consistent with the quantum state |±α.|\pm\alpha\rangle. For either state |±α|\pm\alpha\rangle, one deduces S2S3=cos(π/4)\langle S_{2}S_{3}\rangle=\cos(\pi/4), since the system evolves for a further time π/4Ω\pi/4\Omega. Thus, the prediction is S2S3=cos(π/4)\langle S_{2}S_{3}\rangle=\cos(\pi/4). Overall, the value for the left-side of the Leggett-Garg inequality (28) is 1.41.4, which violates the Leggett-Garg inequality.

A criticism of the above approach is that there is a change due to the measurement M^\hat{M} at time t2t_{2}, and the state prior to the measurement was different to |α|\alpha\rangle. However, macrorealism postulates the existence of an ideal measurement M^\hat{M} where any such effect will have negligible impact on the future dynamics, as the system becomes macroscopic (α\alpha\rightarrow\infty). This would not change the prediction of a violation, if macrorealism is correct.

The dynamics as depicted in terms of the Q function is given for the sequence of times t1<t2<t3t_{1}<t_{2}<t_{3} in Figure 11. There are two cases to compare: whether a measurement M^\hat{M} is performed at time t2t_{2}, or not. Details are given in the Appendix. The lower sequence plots where a measurement is performed at time t2t_{2}, assuming that for the ideal measurement this is done in such a way to instigate a collapse into one or other of the coherent states, as above. The Q functions (LABEL:eq:q2-2) and (48) describe the states for an ensemble of systems AA immediately after the time t2t_{2}, if a measurement has or has not taken place at time t2,t_{2}, respectively. The two Q functions differ by the term

𝒞=epA2xA22πeα2sin(2pAα0)\mathcal{C}=-\frac{e^{-p_{A}^{2}-x_{A}^{2}}}{\sqrt{2}\pi}e^{-\alpha^{2}}\sin(2p_{A}\alpha_{0}) (31)

This term is negligible for even moderate α1.5\alpha\sim 1.5 and there is no distinguishable difference between the functions at the time t2t_{2} (refer Figure 11). Despite the negligible difference between the (average) states at time t2t_{2}, the states at time t3t_{3} are macroscopically different. The Q function model is consistent with the Leggett-Garg premise that a measurement exists which has increasingly negligible effect on the immediate state of the system as the system becomes larger legggarg-1 . This is because 𝒞\mathcal{C} is proportional to eα2e^{-\alpha^{2}}. In this model, the Leggett-Garg inequalities are violated because the small effect nonetheless impacts the future dynamics at a macroscopic level.

Refer to caption
Refer to caption
?figurename? 11: The Q function Q(xA,pA)Q(x_{A},p_{A}) for times t1t_{1}, t2t_{2} and t3t_{3} (α=3\alpha=3). The top sequence assumes no measurement is made at t2t_{2}. The lower sequence assumes a measurement is made at t2t_{2}, of the type that collapses the system into one or other of two coherent states. While the difference in the Q function at time t2t_{2} due to measurement is negligible, this leads to distinct dynamics, and a macroscopic difference at time t3t_{3}.

A challenge for testing macrorealism is to overcome the criticism that the ideal measurement cannot be realised. While arguments can be made that the system at t2t_{2} was in a state sufficiently close to |α|\alpha\rangle or |α|-\alpha\rangle, if macrorealism is correct NSTmunro-1 , it becomes difficult to exclude that the system, in a theory alternative to quantum mechanics, is described by some other state at time t2t_{2}, prior to measurement. This motivates the next section.

IV.2 Bipartite Leggett-Garg tests

IV.2.1 Violation of inequalities: idealised measurement at BB

We next analyse a different model for the ideal measurement that occurs at time t2t_{2}, where strong justification can be given that the measurement at t2t_{2} is nondisturbing. We will see that the predictions of a violation of the Leggett-Garg inequality (28) are unchanged. Consider the set-up of Figure 2 where the system is prepared in a Bell cat state

|ψBell=𝒩(|α|β|α|β)|\psi_{Bell}\rangle=\mathcal{N}(|\alpha\rangle|-\beta\rangle-|-\alpha\rangle|\beta\rangle) (32)

as in eqn. (6). We propose a test of the Leggett-Garg inequality (28) for system AA, where the second system BB is used to perform measurements on system AA. After preparation, both systems AA and BB interact with the nonlinear medium as given by (20) at the respective locations and evolve according to the same shared clock. At the later time t2=π/4Ωt_{2}=\pi/4\Omega, the overall state is

Uπ/8(A)Uπ/8(B)|ψBell\displaystyle U_{\pi/8}^{(A)}U_{\pi/8}^{(B)}|\psi_{Bell}\rangle =\displaystyle= 𝒩eiπ/4(|α|β|α|β)\displaystyle\mathcal{N}e^{-i\pi/4}(|\alpha\rangle|-\beta\rangle-|-\alpha\rangle|\beta\rangle)

The system remains in a Bell cat state and there is a perfect anticorrelation between the spin results S2(A)S_{2}^{(A)} and S2(B)S_{2}^{(B)} at AA and BB. Hence, a measurement of Si(A)S_{i}^{(A)} can be made noninvasively, by performing a measurement of the spin Si(B)S_{i}^{(B)}, where i=1,2i=1,2. The assumption of noninvasiveness is justified by that of locality: It is assumed that the outcome of the measurement on the spacelike separated system AA is independent of the choice of duration of the unitary evolution at BB.

One requires to measure S1(A)S2(A)\langle S_{1}^{(A)}S_{2}^{(A)}\rangle. One infers the result for S1(A)S_{1}^{(A)} by measuring S1(B)S_{1}^{(B)}. A measurement is then made of S2(A)S_{2}^{(A)} at the time t2t_{2}, assuming such a measurement will accurately reflect the macroscopic value of the spin S2(A)S_{2}^{(A)}of system AA immediately prior to the time t2t_{2}. Thus S1(A)S2(A)=S1(B)S2(A)\langle S_{1}^{(A)}S_{2}^{(A)}\rangle=-\langle S_{1}^{(B)}S_{2}^{(A)}\rangle. To obtain the prediction for S1(A)S2(A)\langle S_{1}^{(A)}S_{2}^{(A)}\rangle, one argues as follows. The result of the first measurement at time t1t_{1} at AA (as measured at BB) is either +1+1 or 1-1. Assuming the measurement at BB is such to cause a “collapse” so that the system at AA is then either in the coherent state |α|\alpha\rangle or |α|-\alpha\rangle, the system evolves as given by eq. (29) and S1(A)S2(A)=cos(π/4)\langle S_{1}^{(A)}S_{2}^{(A)}\rangle=\cos(\pi/4). Similarly, S1(A)S3(A)=S1(B)S3(A)=cos(π/2)=0\langle S_{1}^{(A)}S_{3}^{(A)}\rangle=-\langle S_{1}^{(B)}S_{3}^{(A)}\rangle=\cos(\pi/2)=0.

One also requires to measure S2(A)S3(A)\langle S_{2}^{(A)}S_{3}^{(A)}\rangle. The result for S2(A)S_{2}^{(A)} at time t2t_{2} is inferred by measuring S2(B)S_{2}^{(B)} at time t2t_{2}. However, to measure S2(A)S_{2}^{(A)} this way, we use state (LABEL:eq:bellunitary), which means that no measurement is made on system BB (or AA) at the earlier time t1t_{1}. The measurement at time t2t_{2} implies either +1+1 or 1-1, for the spin of AA at time t2t_{2}, and hence we obtain the prediction for S2(A)S3(A)=cos(π/4)\langle S_{2}^{(A)}S_{3}^{(A)}\rangle=\cos(\pi/4). This gives the left side of the Leggett-Garg inequality (28) as 1.41.4 and a violation is predicted, as in Section IV.A.

The analysis also implies violation of the Leggett-Garg-Bell inequality (4) presented in Section II. We have taken the anti-correlated states (6) (and (32)), so that the Leggett-Garg inequality (28) is written

S1(B)S2(A)S2(B)S3(A)+S1(B)S3(A)1.{\color[rgb]{0,0,0}{\color[rgb]{0,0,0}-\langle S_{1}^{(B)}S_{2}^{(A)}\rangle-\langle S_{2}^{(B)}S_{3}^{(A)}\rangle+\langle S_{1}^{(B)}S_{3}^{(A)}\rangle}\leq 1}. (34)

It is convenient to consider the correlated Bell state (6) (and (32)) given by ββ\beta\rightarrow-\beta so that the values of the spins Si(A)S_{i}^{(A)} and Si(B)S_{i}^{(B)} are equal. The Leggett-Garg inequality (28) becomes

S1(B)S2(A)+S2(B)S3(A)S1(B)S3(A)1{\color[rgb]{0,0,0}{\color[rgb]{0,0,0}\langle S_{1}^{(B)}S_{2}^{(A)}\rangle+\langle S_{2}^{(B)}S_{3}^{(A)}\rangle-\langle S_{1}^{(B)}S_{3}^{(A)}\rangle}\leq 1} (35)

which may also be measured as

S1(A)S2(B)+S2(B)S3(A)S1(B)S3(A)1{\color[rgb]{0,0,0}{\color[rgb]{0,0,0}\langle S_{1}^{(A)}S_{2}^{(B)}\rangle+\langle S_{2}^{(B)}S_{3}^{(A)}\rangle-\langle S_{1}^{(B)}S_{3}^{(A)}\rangle}\leq 1} (36)

This reduces to the Leggett-Garg-Bell inequality (4) of Section I. Thus, E(t1,t2)=S1(A)S2(B)E(t_{1},t_{2})=\langle S_{1}^{(A)}S_{2}^{(B)}\rangle, E(t2,t3)=S2(B)S3(A)E(t_{2},t_{3})=\langle S_{2}^{(B)}S_{3}^{(A)}\rangle and E(t1,t3)=S1(A)S3(A)E(t_{1},t_{3})=\langle S_{1}^{(A)}S_{3}^{(A)}\rangle.

IV.2.2 Predictions for the realistic measurement

The above calculation assumes that the measurement at BB “collapses” the system at AA into one or other of the coherent states, |α|\alpha\rangle or |α|-\alpha\rangle, at the time of the measurement, t2t_{2}. The calculations based on that assumption can now be rigorously validated, since we give a specific proposal, that the measurement at site BB be performed as a quadrature phase amplitude measurement X^B\hat{X}_{B}. The spin Sj(B)S_{j}^{(B)}is the sign of the outcome XBX_{B}. In this section, we carry out a complete analysis by evaluating P(XA,XB)P(X_{A},X_{B}) as in Section III.C. The results shown in Figures 12 and 13 confirm the prediction of the violation of the Leggett-Garg-Bell inequality (28) and (4) for large α,\alpha, β\beta.

Refer to caption
?figurename? 12: Distributions giving the violation of the bipartite Leggett-Garg-Bell inequality (34). Two systems AA and BB are prepared in the entangled cat state |ψBell|\psi_{Bell}\rangle (6) and then evolve according to the unitary transformations UA(ta)U_{A}(t_{a}) and UB(tb)U_{B}(t_{b}) (20) as in Figure 2. We show the contour plots of the joint probability of P(XA,XB)P(X_{A},X_{B}) for the measurements X^A\hat{X}_{A} and X^B\hat{X}_{B} made at the times tat_{a} and tbt_{b}, with α=β=3\alpha=\beta=3. The last three distributions give the moments S1(B)S2(A)\langle S_{1}^{(B)}S_{2}^{(A)}\rangle, S1(B)S3(A)\langle S_{1}^{(B)}S_{3}^{(A)}\rangle and S2(B)S3(A)\langle S_{2}^{(B)}S_{3}^{(A)}\rangle. Here, t1=0t_{1}=0, t2=π/4Ωt_{2}=\pi/4\Omega, and t3=π/2Ωt_{3}=\pi/2\Omega. Similar distributions and violations are obtained for the bipartite Leggett-Garg-Bell inequality (4) for the correlated Bell state, where XBXBX_{B}\rightarrow-X_{B}.
Refer to caption
Refer to caption
?figurename? 13: The violation of the Leggett-Garg-Bell inequality (34) and (4) for the system depicted in Figures 2 and 12. We plot BlgB_{lg} given by (37) versus β\beta for the state |ψBell|\psi_{Bell}\rangle (6), where α=β.\alpha=\beta. Violation is obtained when Blg>1B_{lg}>1. The verification of Si(A)Sj(A)=Si(B)Sj(A)\langle S_{i}^{(A)}S_{j}^{(A)}\rangle=-\langle S_{i}^{(B)}S_{j}^{(A)}\rangle for i=1,2i=1,2 is given by examination of the conditional distribution P(Si(A)=1|Si(B)=1)P(S_{i}^{(A)}=1|S_{i}^{(B)}=-1) as shown. The same results are obtained for the Bell state (6) with α=β\alpha=-\beta, on defining Blg=S1(A)S2(B)S1(B)S3(A)+S2(B)S3(A)B_{lg}=\langle S_{1}^{(A)}S_{2}^{(B)}\rangle-\langle S_{1}^{(B)}S_{3}^{(A)}\rangle+\langle S_{2}^{(B)}S_{3}^{(A)}\rangle and considering P(Si(A)=1|Si(B)=1)P(S_{i}^{(A)}=1|S_{i}^{(B)}=1).

The dynamics of the measurement and its disturbance is visualised, by plotting sequences of the QQ function (Figure 14). In order to measure S1(A)S2(A)\langle S_{1}^{(A)}S_{2}^{(A)}\rangle or S1(A)S3(A)\langle S_{1}^{(A)}S_{3}^{(A)}\rangle, via the measurable moments S1(B)S2(A)\langle S_{1}^{(B)}S_{2}^{(A)}\rangle and S1(B)S3(A)\langle S_{1}^{(B)}S_{3}^{(A)}\rangle, one measures the spin BB at t1=0t_{1}=0 (top and centre sequences). In order to measure S2(A)S3(A)\langle S_{2}^{(A)}S_{3}^{(A)}\rangle via the measurable moment S2(B)S3(A)\langle S_{2}^{(B)}S_{3}^{(A)}\rangle, one measures the spin BB at time t2t_{2} (with no measurement at t1t_{1}). The measurements are made by stopping the evolution at BB at the time t1t_{1}, or t2t_{2}, respectively.

Refer to caption
?figurename? 14: Q function dynamics as the entangled cat state |ψBell|\psi_{Bell}\rangle evolves through the three measurement sequences of the bipartite Leggett-Garg test. Plots show the marginal distribution Q(xA,xB)Q(x_{A},x_{B}) after evolving for times tat_{a} and tbt_{b} at the sites AA and BB. The subsystems evolve acording to the same shared clock, with ta=tbt_{a}=t_{b}, up to the time of measurement at the site BB. Here, t1=0t_{1}=0, t2=π/4Ωt_{2}=\pi/4\Omega, and t3=π/2Ωt_{3}=\pi/2\Omega. The top, centre and lower rows give a sequence of snapshots corrresponding to the measurement of S1(B)S2(A)\langle S_{1}^{(B)}S_{2}^{(A)}\rangle, S1(B)S3(A)\langle S_{1}^{(B)}S_{3}^{(A)}\rangle and S2(B)S3(A)\langle S_{2}^{(B)}S_{3}^{(A)}\rangle respectively. The results agree with those of Figure 11.

The top two sequences of Figure 14 show measurement of S1(A)S2(A)\langle S_{1}^{(A)}S_{2}^{(A)}\rangle or S1(A)S3(A)\langle S_{1}^{(A)}S_{3}^{(A)}\rangle, where the system at BB is frozen at t1=tb=0t_{1}=t_{b}=0, so that information about the state of the system AA at time t1t_{1} is stored at that site BB, and AA continues to evolve. We see from the plot of Q(xA,xB)Q(x_{A},x_{B}) for tb=0t_{b}=0 and ta=t3t_{a}=t_{3} (last plot of the centre sequence) that S1(B)S3(A)=0\langle S_{1}^{(B)}S_{3}^{(A)}\rangle=0. Similarly, the plot for tb=0t_{b}=0 and ta=t2t_{a}=t_{2} (last plot of the top sequence) shows S1(B)S2(A)=0.707\langle S_{1}^{(B)}S_{2}^{(A)}\rangle=-0.707. The lower sequence of Figure 14 corresponds to a measurement at site BB made at time t2t_{2}, in order to evaluate S2(B)S3(A)=S2(A)S3(A)\langle S_{2}^{(B)}S_{3}^{(A)}\rangle=-\langle S_{2}^{(A)}S_{3}^{(A)}\rangle. The evolution at BB ceases at t2t_{2} so that the value for S2(A)S_{2}^{(A)} can be measured. The system AA continues to evolve, until t3t_{3}. The correlations giving S2(B)S3(A)=S2(A)S3(A)\langle S_{2}^{(B)}S_{3}^{(A)}\rangle=-\langle S_{2}^{(A)}S_{3}^{(A)}\rangle are indicated by the final plot of the lower sequence. We find S2(B)S3(A)=0.707\langle S_{2}^{(B)}S_{3}^{(A)}\rangle=-0.707 and hence S2(A)S3(A)=0.707\langle S_{2}^{(A)}S_{3}^{(A)}\rangle=0.707. The exact values for the correlations are evaluated numerically by integration of the joint distributions P(XA,XB)P(X_{A},X_{B}) at the specified times. The calculation given in the Appendix leads to

Blg\displaystyle{\color[rgb]{1,0,0}{\color[rgb]{0,0,0}B{}_{lg}}} =\displaystyle= {S1(B)S2(A)S1(B)S3(A)+S2(B)S3(A)}\displaystyle-\{\langle S_{1}^{(B)}S_{2}^{(A)}\rangle-\langle S_{1}^{(B)}S_{3}^{(A)}\rangle+\langle S_{2}^{(B)}S_{3}^{(A)}\rangle\} (37)
=\displaystyle= 2erf(2α)erf(2β){1e2|α|22|β|2}\displaystyle\frac{\sqrt{2}erf(\sqrt{2}\alpha)erf(\sqrt{2}\beta)}{{\color[rgb]{0,0,1}{\color[rgb]{0,0,0}\{1-e^{-2\left|\alpha\right|^{2}-2\left|\beta\right|^{2}}\}}}}

where erferf is the error function. Results are shown in Figure 13, in agreement with the simple approach of the last section.

The pictures of Figure 11 are also justified for this measurement procedure. The impact of whether the measurement takes place at time t2t_{2}, or not, is seen to be minimal if we examine the marginal (reduced) state for the system A, immediately after time t2t_{2}. This is consistent with the idealised model of the measurement given in Figure 11. If there was no measurement at t2t_{2}, and the measurement was made at time t1t_{1}, then the system AA at time t2t_{2} is in a superposition of the two coherent states. On the other hand, if a measurement is made at time t2t_{2}, then the marginal (reduced) state for AA is the mixture, since no preparation took place at t1t_{1} and the statistics for the system is given by (LABEL:eq:bellunitary). The marginal Q functions in each case are (48) and (LABEL:eq:q2-2), given as the second plots of the two sequences in Figure 11, which give indistinguishable predictions for large α\alpha, justifying the arguments that this is an (apparently) non-disturbing measurement.

Despite the negligible difference in the marginal Q functions for system AA at the time t2t_{2}, a macroscopic difference in the joint correlations measured between AA and BB emerges at the later time t3t_{3} (and indeed, has already emerged at the time t2t_{2}). The correlations shown in Figures 12 and 14 imply results for the two-time moments corresponding exactly to those shown in the final plots of the two sequences of Figure 11. The macroscopic difference between the final plots of the two sequences in Figure 11 arises over the timescale of the unitary evolution that determines the measurement - this is the evolution seen in the first three plots of the centre and lower sequence, of Figure 14. The subsequent dynamics at AA after t2t_{2} leads to the macroscopic difference in correlations seen in the final plots of the sequences of Figure 11. On very short timescales (evident at ta=π/16Ωt_{a}=\pi/16\Omega of Figure 14) as the unitary rotation takes place, we note the system cannot be considered a two-state system.

V Delayed collapse

V.1 Delaying the collapse stage of the measurement

We now clarify a possible point of confusion about the timing of the final readout (the “collapse” stage) of the measurement at BB, for the Leggett-Garg-Bell proposal of Section IV.B. The measurement of S1(A)S_{1}^{(A)} is done by measuring S1(B)S_{1}^{(B)} at time t1=0t_{1}=0 and inferring from the correlation between the two systems, AA and BB. The measurement of S2(B)S_{2}^{(B)} comes in two stages: First, it is necessary to stop the unitary interaction of system BB at the time t1=0t_{1}=0, so that information from the correlation can be stored. The second stage of the measurement at BB constitutes the irreversible “collapse”, where there is a coupling to a detector. While the timing of the first stage is crucial, the timing of the “collapse” stage at BB is immaterial.

To clarify, we compare where the collapse at BB has, or has not, occurred prior to the time of the (collapse) measurement of XAX_{A} at AA. First, let us assume the collapse at BB has not occurred. Here, the two systems AA and BB are prepared in the entangled Bell state at time t1=0t_{1}=0. System BB does not evolve further (tb=0t_{b}=0), while system AA evolves for a time ta>0t_{a}>0. The state formed after the evolution at AA is the superposition

UA(ta)|ψBell=𝒩UA(ta){|αa|βb|αa|βb}U_{A}(t_{a})|\psi_{Bell}\rangle=\mathcal{N}U_{A}(t_{a})\{|\alpha\rangle_{a}|-\beta\rangle_{b}-|-\alpha\rangle_{a}|\beta\rangle_{b}\} (38)

The final joint distribution P(XA,XB)supP(X_{A},X_{B})_{sup} that describes the statistics at the later time tat_{a}, assuming there has been no prior collapse at BB, is calculated straightforwardly for this superposition state, as shown in the Appendix.

Refer to caption
Refer to caption
?figurename? 15: The timing of the final collapse stage of the measurement S1(B)S_{1}^{(B)} at time t1t_{1} on BB is immaterial, for α,β>1\alpha,\beta>1: The plots show a sequence of snapshots of Q(xA,xB)Q(x_{A},x_{B}) as the system described in Figure 14 evolves for the evaluation of S1(A)S2(A)\langle S_{1}^{(A)}S_{2}^{(A)}\rangle and S1(A)S3(A)\langle S_{1}^{(A)}S_{3}^{(A)}\rangle. (a) The top two sequences: Here, α=β=0.5\alpha=\beta=0.5. The top sequence shows the evolution if no collapse of the measurement at BB has taken place before the time t3=π/2Ωt_{3}=\pi/2\Omega. The lower sequence shows the evolution assuming the collapse has taken place at time tb=t1=0t_{b}=t_{1}=0. (b) The lower two sequences: As for (a), but here α=β=1.5\alpha=\beta=1.5. The states for the top and lower sequences although different are visually indistinguishable.

Now let us assume the final stage of the measurement at BB was made at the time t=0t=0, meaning that the pointer measurement consisting of a readout of the value of spin S1(B)S_{1}^{(B)} occurred at this time. In this case, the system BB has been coupled to a third system, so that a “collapse” occurs, for the system BB (and also for AA). The system immediately after t=0t=0 is given as the mixture ρmix\rho_{mix} (eqn (LABEL:eq:mixqstate)). The subsequent evolution is different to the evolution of the superposition (38). If a measurement is made at the later time tat_{a}, the joint probability P(XA,XB)mixP(X_{A},X_{B})_{mix} is readily calculated, as shown in the Appendix. The two probabilities P(XA,XB)supP(X_{A},X_{B})_{sup} and P(XA,XB)mixP(X_{A},X_{B})_{mix} are indeed different, due to the interference terms present for the superposition state, where the collapse does not occur. However, for α\alpha and β\beta large, the two probabilities are indistinguishable.

The result is illustrated in Figure 15 where we consider measurements of S1(A)S2(A)\langle S_{1}^{(A)}S_{2}^{(A)}\rangle and S1(A)S3(A)\langle S_{1}^{(A)}S_{3}^{(A)}\rangle, for both small and large α\alpha and β\beta. The Q functions depicted in the final plots of the top and lower sequences directly reflect the probabilities P(XA,XB)supP(X_{A},X_{B})_{sup} and P(XA,XB)mixP(X_{A},X_{B})_{mix} for the two cases, where collapse has not, and has, taken place, respectively, by the time of the detection at AA. Although numerically different, these plots are, in effect, indistinguishable, for α,β>1\alpha,\beta>1.

A similar result occurs if we consider the measurement of S2(A)S3(A)\langle S_{2}^{(A)}S_{3}^{(A)}\rangle. It makes no effective difference to the joint probabilities P(XA,XB)P(X_{A},X_{B}) whether or not the final collapse stage of the measurement S2(B)S_{2}^{(B)} at BB is delayed until a later time. The details are given in the Appendix.

V.2 Interpreting the delayed-collapse cat-state experiment

We ask what can be concluded from the Leggett-Garg-Bell experiments described in Sections IV and V.A, where the collapse stage of the measurement can be delayed? The latter will imply that one can perform the macroscopic Bell and Leggett-Garg tests as delayed choice and quantum eraser experiments delayed-choice-qubit . This follows naturally from the mapping from the microscopic to macroscopic qubits, which is ideal in the limit of large α\alpha, for the angle choices needed for the delayed choice experiments. The new feature, compared to former delayed-choice experiments, is that the results are at the macroscopic level, where the qubit spin values reflect the macroscopically distinct amplitudes.

An argument can be given for interpretating the delayed collapse gedanken experiment in a way that is consistent with the premise of weak macroscopic realism (wMR). Here, the definition of wMR includes the assumption of macroscopic locality for the pointer (MLP) at A, as defined in Section III.E. The argument is based on the macroscopic nature of the experiment. After the time t2t_{2}, the final collapse stage of the measurement S2(B)S_{2}^{(B)} at BB gives information about the macroscopic value of the spin S2(A)S_{2}^{(A)} of system AA (should it be measured). The delayed-collapse analysis of Sections V.A confirms that this information can be revealed at system BB an infinite time after t2t_{2}, and after further events at AA. The system BB can be separated from AA by an infinite distance. A strong argument can then be given that carrying out the collapse stage of the measurement at BB does not macroscopically change the state-description of AA at the former time t2t_{2} i.e. does not change the outcome for the macroscopic spin S2(A)S_{2}^{(A)} at AA.

We note that the wMR interpretation does not mean that prior to the measurement of spin S2(B)S_{2}^{(B)} at BB the system is in the quantum state |α|\alpha\rangle or |α|-\alpha\rangle, since these states are microscopically specified, giving predictions for all measurements that might be performed on AA. That the system in the cat state cannot be viewed as a classical mixture can be negated in multiple ways, e.g. by observing fringes in the distribution for the orthogonal quadrature P^A\hat{P}_{A} (for a single system AA) or by deducing an Einstein-Podolsky-Rosen (EPR) paradox for the entangled state |ψBell|\psi_{Bell}\rangle, along the lines given in Refs. eric_marg-1 ; irrealism-fringes ; macro-coherence-paradox ; macro-pointer ; Bohm-1 ; epr-1 ; bohm-eric ; epr-r2 . We give an example of this in the next subsection.

The consistency with wMR is explained as follows. Suppose the systems AA and BB are prepared at a time tjt_{j} in a macroscopic superposition |ψpointer|\psi_{pointer}\rangle of states with definite outcomes for pointer measurements M^(A)\hat{M}^{(A)} and M^(B)\hat{M}^{(B)}, an example being (as α,\alpha,β\beta\rightarrow\infty)

|ψBell=𝒩(|α|β|α|β)|\psi_{Bell}\rangle=\mathcal{N}(|\alpha\rangle|-\beta\rangle-|-\alpha\rangle|\beta\rangle) (39)

Here, the states with definite outcomes for M^(A)\hat{M}^{(A)} and M^(B)\hat{M}^{(B)} are the states with definite outcomes for the spin Sj(A)S_{j}^{(A)}. The assumption of wMR asserts that the system AA at the time tjt_{j} is in one or other of two macroscopically specified states φ+\varphi_{+} and φ\varphi_{-}, for which the result of a measurement of spin Sj(A)S_{j}^{(A)} is deterministically predetermined i.e. the system AA at time tjt_{j} may be described by a macroscopic hidden variable λj(A)\lambda_{j}^{(A)}. For the bipartite system, wMR is to be consistent with MLP. MLP asserts that the value of the macroscopic hidden variable λj(A)\lambda_{j}^{(A)} for the system AA cannot be changed by any spacelike separated event or measurement at the system BB that takes place at time ttjt\geq t_{j} e.g. it cannot be changed by a future event at BB. The interpretation is that the system AA at each time t1t_{1}, t2t_{2} and t3t_{3} was in one or other of states φj,+\varphi_{j,+} or φj,\varphi_{j,-} with a definite value of spin Sj(A)S_{j}^{(A)}, and that the failure of macrorealism arises from the initial entanglement as the systems evolve dynamically at the intermediate times.

The assumption of MLP is to be distinguished from the stronger assumption, macroscopic locality (ML), introduced earlier, as part of the premise of macroscopic local realism (MLR). The premise of ML assumes locality to apply for spacelike separations where the measurement on system AA can vary, so that AA is not necessarily prepared in the pointer basis. The above wMR interpretation is thus not contradicted by the violation of macroscopic Bell inequalities. In fact, we have seen in Section III that, consistent with the validity of wMR, there is no violation of these Bell inequalities when there is no unitary transformation UAU_{A} at AA. This is because in that case the system is in (immediately prior to the measurements) a state |ψpointer|\psi_{pointer}\rangle where wMR applies.

V.3 Einstein-Podolsky-Rosen-type paradox based on macroscopic realism

The EPR argument argues the incompleteness of quantum mechanics based on the assumption of local realism, or local causality epr-1 . One may also argue an EPR-type paradox for the cat states based on the validity of weak macroscopic realism. Here, we apply the argument given in Ref. macro-coherence-paradox .

One considers the superposition state c1|α+ic2|αc_{1}|\alpha\rangle+ic_{2}|-\alpha\rangle (for α\alpha large), similar to that prepared at the time t2t_{2} in the Leggett-Garg-Bell tests. Here, |c1|2+|c2|2=1|c_{1}|^{2}+|c_{2}|^{2}=1. Macroscopic realism postulates that the system AA in such a state is actually in one or other state φ+\varphi_{+} and φ\varphi_{-} for which the value of the macroscopic spin S2(A)S_{2}^{(A)} is predetermined. The spin S2(A)S_{2}^{(A)} is measured by the sign of XAX_{A}: the distribution P(XA)P(X_{A}) gives two separated Gaussian peaks, each with variance (ΔXA)2=1/2(\Delta X_{A})^{2}=1/2.

One may specify the degree of predetermination of XAX_{A} for the macroscopic states, φ+\varphi_{+} and φ\varphi_{-}, by postulating for each state φ+\varphi_{+} and φ\varphi_{-} that the variance be some specified value, which we take as (ΔXA)+2(\Delta X_{A})_{+}^{2} and (ΔXA)2(\Delta X_{A})_{-}^{2} respectively. If we assume each of φ+\varphi_{+} and φ\varphi_{-} to be quantum states, then for each the Heisenberg uncertainty relation implies (ΔXA)(ΔPA)1/2(\Delta X_{A})(\Delta P_{A})\geq 1/2. We then argue, as was done in ref. macro-coherence-paradox , that for the ensemble of systems in a classical mixture of such states, φ+\varphi_{+} and φ\varphi_{-}, we require (ΔXA)ave(ΔPA)1/2(\Delta X_{A})_{ave}(\Delta P_{A})\geq 1/2, where (ΔXA)ave2=P+(ΔXA)+2+P(ΔXA)2(\Delta X_{A})_{ave}^{2}=P_{+}(\Delta X_{A})_{+}^{2}+P_{-}(\Delta X_{A})_{-}^{2} , P+=|c+|2P_{+}=|c_{+}|^{2} and P=|c|2P_{-}=|c_{-}|^{2}. The observation of

εM(ΔXA)ave(ΔPA)<1/2\varepsilon_{M}\equiv(\Delta X_{A})_{ave}(\Delta P_{A})<1/2 (40)

implies incompatibility of macroscopic realism with the completeness of quantum mechanics, since the localised states φ+\varphi_{+} and φ\varphi_{-} cannot be given as quantum states.

If we take (ΔXA)+2=(ΔXA)2=1/2(\Delta X_{A})_{+}^{2}=(\Delta X_{A})_{-}^{2}=1/2, to match that the states have the variances associated with the two Gaussian peaks (the coherent states), we find the condition (40) is satisfied for (ΔPA)2<1/2(\Delta P_{A})^{2}<1/2. This is indeed the case for the superposition above. The distribution P(PA)P(P_{A}) is

P(PA)=ePA2π{1+12sin(22PA|α|)}P(P_{A})=\frac{e^{-P_{A}^{2}}}{\sqrt{\pi}}\{1+\frac{1}{\sqrt{2}}\sin(2\sqrt{2}P_{A}|\alpha|)\} (41)

which shows fringes, as given in ref. yurke-stoler-1 . The variance is (ΔPA)2=12α2e4α2(\Delta P_{A})^{2}=\frac{1}{2}-\alpha^{2}e^{-4\alpha^{2}} macro-pointer , indicating a paradox.

VI Conclusion

In this paper, we have provided ways to test quantum mechanics against macroscopic realism using cat states. We consider two definitions of macroscopic realism. The first, macrorealism (M-R), supplements the assumption of macroscopic realism with the premise of macroscopic noninvasive measurability. The second, macroscopic local realism (MLR), supplements with the premise of macroscopic locality. In this paper, “macroscopically distinct” refers to separations well beyond the quantum noise level \hbar.

In Sections II-IV, we show that quantum mechanics predicts failure of MLR (and M-R) for two spacelike-separated systems prepared in entangled cat states, in a way that is directly analogous to the original Bell and Leggett-Garg proposals. To test MLR, we consider CHSH-type Bell inequalities and three-time Leggett-Garg-Bell inequalities, where measurements are made at two spacelike separated locations, at successive times. One may construct similar tests, based on the Leggett-Garg-Bell inequality described in Section II, involving four times.

The tests give a convincing strategy to demonstrate the non-invasiveness of a measurement that is assumed in a Leggett-Garg test of macro-realism. For a macroscopic system, the usual argument is that the disturbance (at least for some ideal measurement) becomes vanishingly small, so as to have a negligible affect on the later dynamics. In order to refute the criticism that there is a disturbance, methods have been developed to quantify the effect of such a disturbance, if macroscopic realism is satisfied. One approach performs a control experiment that prepares two macroscopically distinguishable states (ψ1\psi_{1} and ψ2\psi_{2}) and then quantifies the effect of disturbance on the inequality, if the system is indeed in one of those states NSTmunro-1 . A counterargument however is that the system at time t2t_{2} is in a state microscopically different to ψ1\psi_{1} (or ψ2\psi_{2}), a state which was not prepared in the laboratory. In the present paper, such criticisms are avoided, because macroscopic realism is defined more broadly. One does not assume that the macroscopically distinct states are specific states ψ1\psi_{1} and ψ2\psi_{2}, nor that the states are quantum states. A direct disturbance due to measurement is ruled out, because the result for the measurement of AA is inferred from the measurement on the separated system BB. The non-classicality is then explained by quantum nonlocality, but at a macroscopic level.

In Section V, we have extended the bipartite Leggett-Garg analysis, to demonstrate that a delay of the final irreversible stage of the measurement makes no difference to the predictions, provided the amplitude of the coherent states is large. This implies that one could perform delayed choice experiments, similar to those examined for qubits in Refs. delayed-choice-qubit ; delayed-choice-interpretation . The possibility of delayed choice experiments is expected because of the mapping onto the qubit system, for large α\alpha, β\beta (as the coherent states become orthogonal), for the rotation angles required for the qubit experiments. The new feature is the observation at a macroscopic level, since the qubits correspond to macroscopically distinct coherent states. As with the qubit experiments, this does not however imply acausal effects delayed-choice-interpretation .

It is clear from the violations of the macroscopic inequalities presented in this paper that MLR, macroscopic local causality (MLC), and M-R fail. We have explained in Section II that the MLR is defined as deterministic macroscopic (local) realism (dMR), that the system is predetermined to be in a state giving a definite outcome for the pointer measurement (M^θ\hat{M}_{\theta} or M^θ\hat{M}_{\theta^{\prime}}) prior to the unitary rotation UU that determines the measurement setting, θ\theta or θ\theta^{\prime}. The violation of the macroscopic Bell inequalities given in this paper show that dMR fails.

The results are however consistent with an interpretation in which a weak macroscopic realism (wMR) holds. The initial Bell state gives predictions for joint probabilities P(XA,PA,XB,PB)P(X_{A},P_{A},X_{B},P_{B}) at time t1t_{1} that are only microscopically different (of order e|α|2\hbar e^{-|\alpha|^{2}}) to those of a non-entangled mixture. Yet, after the choice of measurement setting, the final joint probabilities are macroscopically different. In the wMR interpretation, the macroscopic differences arise during the unitary dynamics, corresponding in a Bell experiment to the choice of measurement setting. Such dynamics shifts the system into a superposition |ψpointer|\psi_{pointer}\rangle of states with definite outcomes for the measurement of the macroscopic spin Si(A)S_{i}^{(A)}, at a given time tit_{i}. The assumption of wMR postulates that a system in a superposition |ψpointer|\psi_{pointer}\rangle is describable by a macroscopic “element of reality”, λM\lambda_{M}, which predetermines the result of the pointer measurement for the macroscopic qubit value Si(A)S_{i}^{(A)}. The interpretation is that the system is in one or other states φ+\varphi_{+} and φ\varphi_{-} giving a definite value for the macroscopic pointer measurement.

An EPR-type paradox arises from the assumption of weak macroscopic realism, if correct. The element of reality “state”, φ+\varphi_{+} or φ\varphi_{-}, of the system prior to the measurement cannot be a quantum state. This is evident from calculations given in macro-coherence-paradox and explained in section V.C, where the states φ+\varphi_{+} or φ\varphi_{-} that would apply to the cat state |α+|α|\alpha\rangle+|-\alpha\rangle can be shown to be inconsistent with the uncertainty principle. Similar paradoxes have been illustrated for the two-slit experiment using the concept of irrealism irrealism-fringes , and for entangled cat states eric_marg-1 , based on the logic of the original EPR paradoxes that reveal the inconsistency of the completeness of quantum mechanics with local realism epr-1 ; Bohm-1 ; bohm-eric ; epr-r2 .

Finally, we comment on the possibility of an experiment. Entangled cat states have been generated cat-bell-wang-1 . The challenge is to realise the unitary rotation, given by the nonlinear Hamiltonian which has a quartic dependence on the field intensity. This is necessary for the full Bell experiment. However, a macroscopic delayed choice/ quantum eraser experiment may be carried out more straightforwardly, since this depends only on the creation of a simple cat state superposition, which can be realised at times t=π/2Ωt=\pi/2\Omega from the nonlinear Kerr interaction HNL=Ωn^2H_{NL}=\Omega\hat{n}^{2} yurke-stoler-1 ; manushan-cat-lg . This interaction has been experimentally achieved collapse-revival-bec-2 ; collapse-revival-super-circuit-1 .

Acknowledgements

This research has been supported by the Australian Research Council Discovery Project Grants schemes under Grant DP180102470.

Appendix

VI.1 Calculation of Bell violations

The joint probability is P(XA,XB)=|XA|XB|ψf|2P(X_{A},X_{B})=|\langle X_{A}|\langle X_{B}|\psi_{f}\rangle|^{2} where |XA|X_{A}\rangle, |XB|X_{B}\rangle are eigenstates of XAX_{A} and XBX_{B} respectively. Using the overlap x|eiθα=1π1/4exp(x22+2x|α|eiθ2{|α|eiθ2}2|α|22)\langle x|e^{i\theta}\alpha\rangle=\frac{1}{\pi^{1/4}}\exp(-\frac{x^{2}}{2}+\frac{2x|\alpha|e^{i\theta}}{\sqrt{2}}-\{\frac{|\alpha|e^{i\theta}}{\sqrt{2}}\}^{2}-\frac{|\alpha|^{2}}{2}), we find

XA|XB|ψf\displaystyle\langle X_{A}|\langle X_{B}|\psi_{f}\rangle =exp(|α|2+|β|2)2π{expXA22+2XA|α|[bexpXB222XB|β|ib¯expXB22+2XB|β|]\displaystyle=\frac{\exp^{-(|\alpha|^{2}+|\beta|^{2})}}{\sqrt{2\pi}}\{\exp^{-\frac{X_{A}^{2}}{2}+\sqrt{2}X_{A}|\alpha|}[b\exp^{-\frac{X_{B}^{2}}{2}-\sqrt{2}X_{B}|\beta|}-i\bar{b}\exp^{-\frac{X_{B}^{2}}{2}+\sqrt{2}X_{B}|\beta|}]
expXA222XA|α|[bexpXB22+2XB|β|+ib¯expXB222XB|β|]}\displaystyle-\exp^{-\frac{X_{A}^{2}}{2}-\sqrt{2}X_{A}|\alpha|}[b\exp^{-\frac{X_{B}^{2}}{2}+\sqrt{2}X_{B}|\beta|}+i\bar{b}\exp^{-\frac{X_{B}^{2}}{2}-\sqrt{2}X_{B}|\beta|}]\} (42)

VI.2 Interference effects

Where ta=0t_{a}=0, the joint probability P(XA,XB)P(X_{A},X_{B}) for outcomes XAX_{A} and XBX_{B} for the system prepared in the entangled superposition (32) is

P(XA,XB)\displaystyle P(X_{A},X_{B}) =\displaystyle= 𝒩2{|XA|αxB|U(B)(tb)|β|2+|XA|αXB|U(B)(tb)|β|2\displaystyle\mathcal{N}^{2}\{|\langle X_{A}|\alpha\rangle\langle x_{B}|U^{(B)}(t_{b})|-\beta\rangle|^{2}+|\langle X_{A}|-\alpha\rangle\langle X_{B}|U^{(B)}(t_{b})|\beta\rangle|^{2} (43)
+XA|αXB|U(B)(tb)|βXA|αXB|U(B)(tb)|β\displaystyle+\langle X_{A}|\alpha\rangle\langle X_{B}|U^{(B)}(t_{b})|-\beta\rangle\langle X_{A}|-\alpha\rangle^{*}\langle X_{B}|U^{(B)}(t_{b})|\beta\rangle^{*}
+XA|αXB|U(B)(tb)|βXA|αXB|U(B)(tb)|β}\displaystyle+\langle X_{A}|\alpha\rangle^{*}\langle X_{B}|U^{(B)}(t_{b})|-\beta\rangle^{*}\langle X_{A}|-\alpha\rangle\langle X_{B}|U^{(B)}(t_{b})|\beta\rangle\}

This contains an interference term proportional to exp(XA22|α|2)\exp(-X_{A}^{2}-2|\alpha|^{2}) that vanishes with large α\alpha. The expression reduces to the result for the mixture ρmix\rho_{mix}, given by

P(XA,XB)\displaystyle P(X_{A},X_{B}) =\displaystyle= 12{|XA|αXB|U2(tb)|β|2+|XA|αXB|U2(tb)|β|2}\displaystyle\frac{1}{2}\{|\langle X_{A}|\alpha\rangle\langle X_{B}|U_{2}(t_{b})|-\beta\rangle|^{2}+|\langle X_{A}|-\alpha\rangle\langle X_{B}|U_{2}(t_{b})|\beta\rangle|^{2}\} (44)

because the interference term contains the expression XA|αXA|α=1π1/2exp(XA22|α|2)\langle X_{A}|\alpha\rangle\langle X_{A}|-\alpha\rangle^{*}=\frac{1}{\pi^{1/2}}\exp(-X_{A}^{2}-2|\alpha|^{2}) which is small for orthogonal state, with large α0\alpha_{0}.

By contrast, for ta0t_{a}\neq 0 and tb0t_{b}\neq 0, we compare the expressions P(XA,XB)P(X_{A},X_{B}) for the evolved cat state given by |ψf|\psi_{f}\rangle of eq. (LABEL:eq:234-1-1-1), with that of the evolved mixture

ρmix(ta,tb)\displaystyle\rho_{mix}(t_{a},t_{b}) =\displaystyle= 12(U(B)(tb)|βU(A)|αα|U(A)β|U(B)(tb)\displaystyle\frac{1}{2}\Bigl{(}U^{(B)}(t_{b})|-\beta\rangle U^{(A)}|\alpha\rangle\langle\alpha|U^{(A)\dagger}\langle-\beta|U^{(B)\dagger}(t_{b}) (45)
+U(B)|βU(A)|αα|U(A)β|U(B))\displaystyle+U^{(B)}|\beta\rangle U^{(A)}|-\alpha\rangle\langle-\alpha|U^{(A)\dagger}\langle\beta|U^{(B)\dagger}\Bigl{)}

The expressions are the same apart from interference terms, such as

XA|U(A)(ta)|αXB|U(B)(tb)|β×XA|U(A)(ta)|αXB|U(B)(tb)|β\langle X_{A}|U^{(A)}(t_{a})|\alpha\rangle\langle X_{B}|U^{(B)}(t_{b})|-\beta\rangle\times\langle X_{A}|U^{(A)}(t_{a})|-\alpha\rangle^{*}\langle X_{B}|U^{(B)}(t_{b})|\beta\rangle^{*} (46)

We see from the solutions given by eqs. (LABEL:eq:state3-20) that for the suitable choices of tat_{a} , U(A)(ta)|αU^{(A)}(t_{a})|-\alpha\rangle (for example) may involve terms such as |α|\alpha\rangle, and U(A)(ta)|αU^{(A)}(t_{a})|\alpha\rangle involves terms in |α|\alpha\rangle. This then leads to a contribution of type XA|αXA|α=1π1/2exp(XA2+22XAα2|α|2)\langle X_{A}|\alpha\rangle\langle X_{A}|\alpha\rangle^{*}=\frac{1}{\pi^{1/2}}\exp(-X_{A}^{2}+2\sqrt{2}X_{A}\alpha-2|\alpha|^{2}) and therefore significant interference terms (when XA2αX_{A}\sim\sqrt{2}\alpha). These terms are the origin of the macroscopic difference between the probability distributions in Figures 6 for nonzero tat_{a} and tbt_{b}, which leads to the macroscopic Bell violation.

VI.3 Leggett-Garg test for a single system: Q function dynamics

The Q function at time t1t_{1} is that of a coherent state |α0|\alpha_{0}\rangle where we take α0\alpha_{0} to be real.

Q(xA,pA,t1)=epA2πe(xAα0)2Q(x_{A},p_{A},t_{1})=\frac{e^{-p_{A}^{2}}}{\pi}e^{-(x_{A}-\alpha_{0})^{2}} (47)

At time t2t_{2}, the system is in the superposition state (29) which has the Q function

Q(xA,pA,t2)=\displaystyle Q(x_{A},p_{A},t_{2})= epA222π{c+e(xAα0)2+ce(xA+α0)22exA2eα02sin(2pAα0)}\displaystyle\frac{e^{-p_{A}^{2}}}{2\sqrt{2}\pi}\{c_{+}e^{-(x_{A}-\alpha_{0})^{2}}+c_{-}e^{-(x_{A}+\alpha_{0})^{2}}-2e^{-x_{A}^{2}}e^{-\alpha_{0}^{2}}\sin(2p_{A}\alpha_{0})\} (48)

where c±=2±1c_{\pm}=\sqrt{2}\pm 1. Assuming no measurement takes place at time t2t_{2}, the system at time t3t_{3} is in the superposition state (30) with

Q(xA,pA,t3)=\displaystyle Q(x_{A},p_{A},t_{3})= epA22π{e(xAα0)2+e(xA+α0)22exA2eα02sin(2pAα0)}\displaystyle\frac{e^{-p_{A}^{2}}}{2\pi}\{e^{-(x_{A}-\alpha_{0})^{2}}+e^{-(x_{A}+\alpha_{0})^{2}}-2e^{-x_{A}^{2}}e^{-\alpha_{0}^{2}}\sin(2p_{A}\alpha_{0})\} (49)

If the measurement is performed at t2t_{2}, the system “collapses” to either |α0|\alpha_{0}\rangle (with probability cos2π/8\cos^{2}\pi/8) or |α0|-\alpha_{0}\rangle (with probability sin2π/8\sin^{2}\pi/8) and then evolves at time t3t_{3} respectively to either Uπ/8|αU_{\pi/8}|\alpha\rangle or Uπ/8|αU_{\pi/8}|-\alpha\rangle as given by eqn. (29).

There are two cases to compare: whether or not a measurement is performed at time t2t_{2}. Figure 12 shows the sequence for the Q functions at the times t1t_{1} , t2t_{2} and t3t_{3} for these two cases. The top sequence, modelling where a measurement is not performed, shows sequentially the three Q functions, (47), (48) and (49). The lower sequence plots where a measurement is performed at time t2t_{2}, assuming this is done in such a way to instigate a collapse into one or other of the coherent states, as described above. The second plot of the lower sequence is therefore the Q function

Q(xA,pA,t2)\displaystyle Q(x_{A},p_{A},t_{2}) =epA222π{c+e(xAα0)2+ce(xA+α0)2}\displaystyle=\frac{e^{-p_{A}^{2}}}{2\sqrt{2}\pi}\{c_{+}e^{-(x_{A}-\alpha_{0})^{2}}+c_{-}e^{-(x_{A}+\alpha_{0})^{2}}\}

representing the average state ρmix(t2)\rho_{mix}(t_{2}) of the system at time t2t_{2}, immediately after the measurement at time t2t_{2}: ρmix(t2)=cos2π/8|α0α0|+sin2π/8|α0α0|\rho_{mix}(t_{2})=\cos^{2}\pi/8|\alpha_{0}\rangle\langle\alpha_{0}|+\sin^{2}\pi/8|-\alpha_{0}\rangle\langle-\alpha_{0}|. The Q function for the final state at time t3t_{3}, if the measurement has taken place at time t2t_{2}, is given by the evolution of ρmix(t2)\rho_{mix}(t_{2}) for the time π/4Ω\pi/4\Omega. The final average state is ρmix(t3)=cos2π/8|ψ+ψ+|+sin2π/8|ψψ|\rho_{mix}(t_{3})=\cos^{2}\pi/8|\psi_{+}\rangle\langle\psi_{+}|+\sin^{2}\pi/8|\psi_{-}\rangle\langle\psi_{-}| where |ψ±=Uπ/8|±α|\psi_{\pm}\rangle=U_{\pi/8}|\pm\alpha\rangle. The Q function is

Q(xA,pA,t3)=\displaystyle Q(x_{A},p_{A},t_{3})= epA28π{6e(xAα0)2+e(xA+α0)24exA2eα02sin(2pAα0)}\displaystyle\frac{e^{-p_{A}^{2}}}{8\pi}\{6e^{-(x_{A}-\alpha_{0})^{2}}+e^{-(x_{A}+\alpha_{0})^{2}}-4e^{-x_{A}^{2}}e^{-\alpha_{0}^{2}}\sin(2p_{A}\alpha_{0})\} (51)

which is plotted as the third function of the lower sequence.

VI.4 Bipartite Leggett-Garg tests: correlation between the outcomes at the different sites

Here, we evaluate Blg={S1(B)S2(A)S1(B)S3(A)+S2(B)S3(A)}B_{lg}=-\{\langle S_{1}^{(B)}S_{2}^{(A)}\rangle-\langle S_{1}^{(B)}S_{3}^{(A)}\rangle+\langle S_{2}^{(B)}S_{3}^{(A)}\rangle\} and the value of the conditional probabilities P(Si(A)=1|Si(B)=1)P(S_{i}^{(A)}=1|S_{i}^{(B)}=-1) where i=1,2i=1,2 for the case α=β\alpha=\beta. These quantities are worked out exactly as one would measure them. We first evaluate P(XA,XB)P(X_{A},X_{B}) at time t2t_{2}.

P(XA,XB)\displaystyle P(X_{A},X_{B}) =|XB|XA|Uπ/8(A)Uπ/8(B)|ψBell|2\displaystyle=\left|\langle X_{B}|\langle X_{A}|U_{\pi/8}^{(A)}U_{\pi/8}^{(B)}|\psi_{Bell}\rangle\right|^{2}
=2eXA2XB22|α|22|β|2π(1e2|α|22|β|2)sinh2(2XA|α|2XB|β|)\displaystyle=2\frac{e^{-X_{A}^{2}-X_{B}^{2}-2\left|\alpha\right|^{2}-2\left|\beta\right|^{2}}}{\pi{\color[rgb]{0,0,1}{\normalcolor\left(1-e^{-2\left|\alpha\right|^{2}-2\left|\beta\right|^{2}}\right)}}}\sinh^{2}(\sqrt{2}X_{A}\left|\alpha\right|-\sqrt{2}X_{B}\left|\beta\right|) (52)

where we use

XB|XA|Uπ/8(A)Uπ/8(B)|ψBell\displaystyle\langle X_{B}|\langle X_{A}|U_{\pi/8}^{(A)}U_{\pi/8}^{(B)}|\psi_{Bell}\rangle =𝒩eiπ/4(XA|αXB|βXA|αXB|β)\displaystyle=\mathcal{N}e^{-i\pi/4}\left(\langle X_{A}|\alpha\rangle\langle X_{B}|-\beta\rangle-\langle X_{A}|-\alpha\rangle\langle X_{B}|\beta\rangle\right)
=2𝒩eXA22XB22π1/2e|α|2|β|2iπ/4sinh(2XA|α|2XB|β|)\displaystyle=2\mathcal{N}\frac{e^{-\frac{X_{A}^{2}}{2}-\frac{X_{B}^{2}}{2}}}{\pi^{1/2}}e^{-\left|\alpha\right|^{2}-\left|\beta\right|^{2}-i\pi/4}\sinh(\sqrt{2}X_{A}\left|\alpha\right|-\sqrt{2}X_{B}\left|\beta\right|) (53)

In fact, because the system at time t2t_{2} remains in a Bell state |ψBell|\psi_{Bell}\rangle (apart from a phase factor), this also represents P(XA,XB)P(X_{A},X_{B}) for the time t1t_{1}. We note that

P(XB)\displaystyle P(X_{B}) =P(XA,XB)𝑑XA=eXB22|α|22|β|2π{1e2|α|22|β|2}(e2|α|2cosh(22|β|XB)1)\displaystyle=\int P(X_{A},X_{B})dX_{A}=\frac{e^{-X_{B}^{2}-2\left|\alpha\right|^{2}-2\left|\beta\right|^{2}}}{\sqrt{\pi}{\color[rgb]{0,0,1}{\normalcolor\{1-e^{-2\left|\alpha\right|^{2}-2\left|\beta\right|^{2}}\}}}}\left(e^{2|\alpha|^{2}}\cosh(2\sqrt{2}|\beta|X_{B})-1\right) (54)

which on integration gives P(XB>0)=1/2P(X_{B}>0)=1/2. For the value of the conditional probability P(Si(A)=1|Si(B)=1)P(S_{i}^{(A)}=1|S_{i}^{(B)}=-1), i=1,2i=1,2, one measures

P(XA>0|XB0)\displaystyle{\color[rgb]{0,0,0}{\color[rgb]{0,0,1}}P(X_{A}>0|X_{B}\leq 0)} =00P(XA,XB)𝑑XA𝑑XB0P(XB)𝑑XB=12+erf(2α)×erf(2β)2(1e2α22β2)\displaystyle{\color[rgb]{0,0,0}=\frac{\int_{-\infty}^{0}\int_{0}^{\infty}P(X_{A},X_{B})dX_{A}dX_{B}}{\int_{-\infty}^{0}P(X_{B})dX_{B}}{\color[rgb]{0,0,0}={\color[rgb]{0,0,0}\frac{1}{2}+\frac{erf(\sqrt{2}\alpha)\times erf(\sqrt{2}\beta)}{{\color[rgb]{0,0,0}2\left(1-e^{-2\alpha^{2}-2\beta^{2}}\right)}}}}} (55)

Similarly, we evaluate

S1(B)S2(A)\displaystyle{\color[rgb]{0,0,0}}{\normalcolor}{\normalcolor}{\normalcolor}\langle S_{1}^{(B)}S_{2}^{(A)}\rangle =2erf(2α)erf(2β)2{1e2|α|22|β|2}\displaystyle=-\frac{\sqrt{2}erf(\sqrt{2}\alpha)erf(\sqrt{2}\beta)}{2{\color[rgb]{0,0,0}{\color[rgb]{0,0,1}{\color[rgb]{0,0,0}\{1-e^{-2\left|\alpha\right|^{2}-2\left|\beta\right|^{2}}\}}}}}
S1(B)S3(A)\displaystyle\langle S_{1}^{(B)}S_{3}^{(A)}\rangle =0\displaystyle=0 (56)

where S2(B)S3(A)=S1(B)S2(A)\langle S_{2}^{(B)}S_{3}^{(A)}\rangle=\langle S_{1}^{(B)}S_{2}^{(A)}\rangle. This gives the result (37). The plots of Figure 13 and 15 reveal that for α=β>1\alpha=\beta>1, the positive and negative values for the outcomes of XAX_{A} correspond to distinct macroscopically separated Gaussian peaks in the different quadrants. The conditional probability goes to 11 as α\alpha is larger, which justifies that the measurement of Si(B)S_{i}^{(B)}at BB indicates the value Si(A)S_{i}^{(A)}at AA.

VI.5 Delayed Collapse for measurement S1(A)S2(A)\langle S_{1}^{(A)}S_{2}^{(A)}\rangle

The state formed after the evolution at AA is the superposition

UA(ta)|ψBell=𝒩{UA(ta)|αa|βbUA(ta)|αa|βb}U_{A}(t_{a})|\psi_{Bell}\rangle=\mathcal{N}\{U_{A}(t_{a})|\alpha\rangle_{a}|-\beta\rangle_{b}-U_{A}(t_{a})|-\alpha\rangle_{a}|\beta\rangle_{b}\} (57)

The final joint distribution at time tat_{a} is (assuming there has been no prior collapse at BB)

P(XA,XB)sup\displaystyle P(X_{A},X_{B})_{sup} =\displaystyle= 𝒩2(XA|UA(ta)|αaXB|βbXA|UA(ta)|αaXB|βb)\displaystyle\mathcal{N}^{2}(\langle X_{A}|U_{A}(t_{a})|\alpha\rangle_{a}\langle X_{B}|-\beta\rangle_{b}-\langle X_{A}|U_{A}(t_{a})|-\alpha\rangle_{a}\langle X_{B}|\beta\rangle_{b}) (58)
×(α|aUA(ta)|XAβ|bXbα|aUA(ta)|XAβ|bXb)\displaystyle\times(\langle\alpha|_{a}U_{A}^{\dagger}(t_{a})|X_{A}\rangle\langle-\beta|_{b}X_{b}\rangle-\langle-\alpha|_{a}U_{A}^{\dagger}(t_{a})|X_{A}\rangle\langle\beta|_{b}X_{b}\rangle)

Now let us assume the final stage of the measurement at BB was made at the time t=0t=0, meaning that the pointer measurement consisting of a readout of the value of spin S1(B)S_{1}^{(B)} occurred at this time. In this case, the system BB has been coupled to a third system, so that a “collapse” occurs, for the system BB (and also for AA). The system (or rather, an ensemble of them) immediately after t=0t=0 is given as the mixture ρmix\rho_{mix} (eqn (LABEL:eq:mixqstate)). The system then evolves from a state that is either |α|β|\alpha\rangle|-\beta\rangle or |α|β|-\alpha\rangle|\beta\rangle, which is different to the evolution of the superposition (57). If a measurement is made at the later time tat_{a}, the joint probability is

P(XA,XB)mix\displaystyle P(X_{A},X_{B})_{mix} =\displaystyle= 12(XA|UA(ta)|αaXB|βbα|aUA(ta)|XAβ|bXb\displaystyle\frac{1}{2}(\langle X_{A}|U_{A}(t_{a})|\alpha\rangle_{a}\langle X_{B}|-\beta\rangle_{b}\langle\alpha|_{a}U_{A}^{\dagger}(t_{a})|X_{A}\rangle\langle-\beta|_{b}X_{b}\rangle (59)
+XA|UA(ta)|αaXB|βb)α|aUA(ta)|XAβ|bXb)\displaystyle+\langle X_{A}|U_{A}(t_{a})|-\alpha\rangle_{a}\langle X_{B}|\beta\rangle_{b})\langle-\alpha|_{a}U_{A}^{\dagger}(t_{a})|X_{A}\rangle\langle\beta|_{b}X_{b}\rangle)

The two probabilities P(XA,XB)P(X_{A},X_{B}) are indeed different. However, for α\alpha and β\beta large, examination shows that the two probabilities (58) and (59) become indistinguishable.

VI.6 Delayed collapse for measurement S2(A)S3(A)\langle S_{2}^{(A)}S_{3}^{(A)}\rangle

We may also examine where a measurement is to be made at time t2t_{2}, as in the evaluation of S2(A)S3(A)\langle S_{2}^{(A)}S_{3}^{(A)}\rangle. Again, it makes no difference as to the evaluation, if the “collapse” stage is delayed until a later time. Wishing to measure the spin of subsystem BB at time t2t_{2} means we then “stop” the evolution of system BB at tb=t2t_{b}=t_{2}. If the system AA continues to evolve, then the state of the system is (ta=tAt2t^{\prime}_{a}=t_{A}-t_{2})

U(A)(tA)|ψBell\displaystyle U^{(A)}(t^{\prime}_{A})|\psi_{Bell}\rangle =\displaystyle= N±eiπ/4(U(A)(tA)|α|βU(A)(tA)|α|β)\displaystyle N_{\pm}e^{-i\pi/4}(U^{(A)}(t^{\prime}_{A})|\alpha\rangle|-\beta\rangle-U^{(A)}(t^{\prime}_{A})|-\alpha\rangle|\beta\rangle) (60)

The final distributions P(XA,XB)supP(X_{A},X_{B})_{sup} and P(XA,XB)mixP(X_{A},X_{B})_{mix} are similar to those above, replacing tAt_{A} by tAt^{\prime}_{A}. The lower sequence of Figure 15 shows the Q function dynamics for the evolution where there has been no collapse at time t2t_{2}. When plotted, the sequence where a collapse has occured shows, even for moderate β\beta, no distinguishable difference, in either the final measured probabilities or the Q function.

?refname?

  • (1) J. S. Bell, Physics 1, 195 (1964).
  • (2) J. F. Clauser, M. A. Horne, A. Shimony and R. A. Holt, Phys. Rev. Lett. 23, 880 (1969). J. S. Bell, Foundations of Quantum Mechanics ed B d’Espagnat (New York: Academic) pp171-81 (1971). J. F. Clauser and A. Shimony, Rep. Prog. Phys. 41, 1881 (1978).
  • (3) N. Brunner, D. Cavalcanti, S. Pironio, V. Scarani, and S. Wehner, Rev. Mod. Phys. 86, 419 (2014).
  • (4) H. M. Wiseman, Journ. Phys A 47, 424001 (2014).
  • (5) N. D. Mermin, Phys. Rev. D 22, 356 (1980). P. D. Drummond, Phys. Rev. Lett. 50, 1407 (1983). J. C. Howell, A. Lamas-Linares, and D. Bouwmeester, Phys. Rev. Lett. 88, 030401 (2002). M. D. Reid and W. J. Munro, Phys. Rev. Lett. 69, 997 (1992). M. D. Reid, Phys. Rev. Lett. 84, 2765 (2000). H. Jeong, M. Paternostro and T. C. Ralph, Phys. Rev. Lett. 102, 060403 (2009). M. Navascués, D. Pérez-García and I. Villanueva, J. Phys. A: Math. Theor. 46 085304 (2013). J. Tura, R. Augusiak, A. B. Sainz, B. Lücke, C. Klempt, M. Lewenstein, A. Acínac, Annals of Physics 362, 370 (2015). B. J. Dalton, arXiv (2018), arXiv:1812.09651 [quant-ph]. A. B. Watts, N. Y. Halpern and A. Harrow, arXiv:1911.09122v1 [quant-ph].
  • (6) N. D. Mermin, Phys. Rev. Lett. 65, 1838 (1990). A. Gilchrist, P. Deuar and M. D. Reid, Phys. Rev. A60, 4259 (1999). K. Wodkiewicz, New Journal of Physics 2, 21 (2000). G. Svetlichny, Phys. Rev. D 35, 3066 (1987). D. Collins et al., Phys. Rev. Lett. 88, 170405 (2002). J. Lavioe et al., New J. Phys 11, 073051 (2009). C. Wildfeuer, A. Lund and J. Dowling, Phys. Rev. A 76, 052101 (2007). F. Toppel, M. V. Chekhova, and G. Leuchs, arXiv:1607.01296 [quant-ph] (2016).
  • (7) U. Leonhardt and J. Vaccaro, J. Mod. Opt. 42, 939 (1995). A. Gilchrist, P. Deuar and M. Reid, Phys. Rev. Lett. 80 3169 (1998). K. Banaszek and K. Wodkiewicz, Phys. Rev. Lett. 82 2009 (1999). Oliver Thearle et al., Phys. Rev. Lett. 120, 040406 (2018). A. Ketterer, A. Keller, T. Coudreau, and P. Milman, Phys. Rev. A 91, 012106 (2015).
  • (8) A.S. Arora and A. Asadian, Phys. Rev. A 92, 062107 (2015). Ming-Da Huang, Ya-Fei Yu, and Zhi-Ming Zhang, Phys. Rev. A 102, 022229 (2020).
  • (9) M. D. Reid, Phys. Rev. A 62, 022110 (2000). M. D. Reid, Phys. Rev. A 97, 042113 (2018).
  • (10) M. Navascués and H. Wunderlich, Proc. Roy. Soc. Lond. A 466, 881 (2009).
  • (11) A. Leggett and A. Garg, Phys. Rev. Lett. 54, 857 (1985).
  • (12) E. Schroedinger, 23, 807 (1935). F. Frowis, P. Sekatski, W. Dur, N. Gisin, and N. Sangouard, Rev. Mod. Phys. 90, 025004 (2018).
  • (13) A. N. Jordan, A. N. Korotkov, and M. Buttiker, Phys. Rev. Lett. 97, 026805 (2006).
  • (14) N. S. Williams and A. N. Jordan, Phys. Rev. Lett. 100, 026804 (2008).
  • (15) C. Emary, N. Lambert, and F. Nori, Rep. Prog. Phys 77, 016001 (2014). A. Palacios-Laloy, F. Mallet, F. Nguyen, P. Bertet, Denis Vion, Daniel Esteve and Alexander N. Korotkov, Nature Phys. 6, 442 (2010). M. E. Goggin, M. P. Almeida, M. Barbieri, B. P. Lanyon, J. L. O’Brien, A. G. White, and G. J. Pryde, Proc. Natl. Acad. Sci. 108, 1256 (2011). G. Waldherr, P. Neumann, S. F. Huelga, F. Jelezko, and J. Wrachtrup, Phys.Rev. Lett. 107, 090401 (2011). G. C. Knee et al., Nature Commun. 3, 606 (2012). Z. Q. Zhou, S. Huelga, C-F Li, and G-C Guo, Phys. Rev. Lett. 115, 113002 (2015). C. Robens, W. Alt, D. Meschede, C. Emary, and A. Alberti, Phys. Rev. X 5, 011003 (2015). J. A. Formaggio et al., Phys. Rev. Lett. 117, 050402 (2016).
  • (16) J. Kofler and C. Brukner, Phys. Rev. A 87, 052115 (2013); L. Clemente and J. Kofler, Phys. Rev. Lett. 116, 150401 (2016).
  • (17) G. C. Knee, K. Kakuyanagi, M.-C. Yeh, Y. Matsuzaki, H. Toida, H. Yamaguchi, S. Saito, A. J. Leggett and W. J. Munro, Nat. Commun. 7, 13253 (2016).
  • (18) J. Dressel and A. N. Korotkov Phys. Rev. A 89, 012125 (2014). J. Dressel, C. J. Broadbent, J. C. Howell and A. N. Jordan, Phys. Rev. Lett. 106, 040402 (2011). T. C. White et al., NPJ Quantum Inf. 2, 15022 (2016). B. L. Higgins, M. S. Palsson, G. Y. Xiang, H. M. Wiseman, and G. J. Pryde, Phys. Rev. A 91, 012113 (2015).
  • (19) A. Asadian, C. Brukner, and P. Rabl, Phys. Rev. Lett. 112, 190402 (2014). C. Budroni, G. Vitagliano, G. Colangelo, R. J. Sewell, O. Gühne, G. Tóth, and M. W. Mitchell, Phys. Rev. Lett. 115, 200403 (2015). J. J. Halliwell, Phys. Rev. A 93, 022123 (2016). L. Rosales-Zárate, B. Opanchuk, Q. Y. He, and M. D. Reid, Phys. Rev. A 97, 042114 (2018). L. Rosales-Zárate, B. Opanchuk, and M. D. Reid, Phys. Rev. A 97, 032123 (2018). A. K. Pan, Phys. Rev. A 102, 032206 (2020). J. Halliwell, A. Bhatnagar, E. Ireland, H. Nadeem and V. Wimalaweera, arXiv:2009.03856v3 [quant-ph]
  • (20) M. Thenabadu and M. D. Reid, Phys. Rev. A 99, 032125 (2019).
  • (21) R. Uola, G. Vitagliano and C. Budroni, Phys. Rev. A 100, 042117 (2019).
  • (22) M. Thenabadu, G-L. Cheng, T. L. H. Pham, L. V. Drummond, L. Rosales-Zárate and M. D. Reid, Phys. Rev. A 102, 022202 (2020).
  • (23) See M. D Reid and M. Thenabadu, accompanying letter.
  • (24) D. Gočanin, A. Dimić, F. D. Santo and B. Dakić, Phys. Rev. A, 102, 020201(R) (2020).
  • (25) A Peres, “Quantum Theory Concepts and Methods”, (Kluwer Academic Publishers) (1995).
  • (26) J. Kofler and C. Brukner, Phys. Rev. Lett. 101, 090403 (2008).
  • (27) D. Poulin, Physical Review A 71, 022102 (2005). J. Kofler and C. Brukner, Phys. Rev. Lett. 99, 180403 (2007). S. Raeisi, P. Sekatski, and C. Simon, Phys. Rev. Lett. 107, 250401 (2011). T. Wang, R. Ghobadi, S. Raeisi, and C. Simon, Phys. Rev. A 88, 062114 (2013). H. Jeong, Y. Lim, and M. S. Kim, Phys. Rev. Lett. 112, 010402 (2014). J. Park, S.-W. Ji, J. Lee, and H. Nha, Phys. Rev. A 89, 042102 (2014). C. Duarte, G. D. Carvalho, N. K. Bernardes, and F. de Melo, Phys. Rev. A 96, 032113 (2017). P. Silva Correia and F. de Melo, Phys. Rev. A 100, 022334 (2019). O. Kabernik, Phys. Rev. A 97, 052130 (2018). C. Duarte, arXiv preprint arXiv:1908.04432 (2019). I. Veeren and F. De Melo, Phys. Rev. A 102, 022205 (2020).
  • (28) S. D. Bartlett, T. Rudolph and R. W. Spekkens,Phys. Rev. A 86, 012103 (2012). R. W. Spekkens, Phys. Rev. A 75, 032110 (2007). N. Harrigan and R. W. Spekkens, Found. Phys. 40, 125 (2010).
  • (29) A. Budiyono and D. Rohrlich, Nature Communications 8, 1306 (2017). A. Budiyono, Phys. Rev. A 100, 032125 (2019). ibid, Phys. Rev. A 100, 062102 (2019).
  • (30) M. D. Reid, J. Phys. A: Math. Theor. 50, 41LT01 (2017).
  • (31) P. D. Drummond and M. D. Reid, Phys. Rev. Research 2, 033266 (2020).
  • (32) M. J. Hall, D. Deckert and H. Wiseman, Phys. Rev. X 4, 041013 (2014).
  • (33) Y. Aharonov, E. Cohen, F. Colombo, T. Landsberger, I. Sabadini, D. Struppa and J. Tollaksen, Proc. Nat. Acad. Sci. 114, 6480 (2017).
  • (34) A. Einstein, B. Podolsky, and N. Rosen, Phys. Rev. 47, 777 (1935).
  • (35) M. D. Reid, Phys. Rev. A 100, 052118 (2019).
  • (36) F. R. Lustosa, P. R. Dieguez, and I. G. da Paz, Phys. Rev. A 102, 052205 (2020).
  • (37) E. G. Cavalcanti and M. D. Reid, Phys. Rev. A. 77, 062108 (2008). L. Rosales-Zarate, R. Y. Teh, S. Kiesewetter, A. Brolis, K. Ng, and M. D. Reid, J. Opt. Soc. Am. B 32 A82 (2015). R. Y. Teh, L. Rosales-Zárate, B. Opanchuk, and M. D. Reid, Phys. Rev. A 94, 042119 (2016).
  • (38) Kôdi Husimi (1940). Proc. Phys. Math. Soc. Jpn. 22: 264.
  • (39) C. Wang et al., Science 352, 1087 (2016).
  • (40) B. Yurke and D. Stoler, Phys. Rev. Lett. 57, 13 (1986).
  • (41) E. Wright, D. Walls and J. Garrison, Phys. Rev. Lett. 77, 2158 (1996).
  • (42) M. O. Scully and K. Drühl, Phys. Rev. A 25, 2208 (1982). S. P. Walborn, M. O. Terra Cunha, S. Pádua, and C. H. Monken, Phys. Rev. A 65, 033818 (2002). Y.-H. Kim, R. Yu, S. P. Kulik, Y. H. Shih and M. O. Scully, Phys. Rev. Lett. 84 (2000). A. Peruzzo, P. Shadbolt, N. Brunner, S. Popescu and J. L. O’Brien, Science 338 6107, (2012).
  • (43) Xiao-Song Ma et al., Proceedings of the National Academy of Sciences. 110 (4) (2013); B. G. Englert, M. O. Scully, and H. Walther, Am. J. Phys. 67, 325 (1999). Sandro Faetti, arXiv:1912.04101 [quant-ph]. Xi. Ma, J. Kofler, and A. Zeilinger, Rev. Mod. Phys. 88, 015005 (2016).
  • (44) D. Bohm, “Quantum Theory” (Constable, London 1951).
  • (45) M. D. Reid, Phys. Rev. A 40, 913 (1989). M. D. Reid, P. D. Drummond, W. P. Bowen, E. G. Cavalcanti, P. K. Lam, H. A. Bachor, U. L. Andersen and G. Leuchs, Rev. Mod. Phys. 81, 1727 (2009).
  • (46) E. G. Cavalcanti, P. D. Drummond, H. A. Bachor and M. D. Reid, Optics Express 17 (21), 18693 (2009).
  • (47) M. Greiner, O. Mandel, T. Hånsch and I. Bloch, Nature 419, 51 (2002).
  • (48) G. Kirchmair et al., Nature 495, 205 (2013).