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

Violations of the Leggett-Garg inequality for coherent and cat states

Hiroo Azuma1,   and   Masashi Ban2,

1Nisshin-scientia Co., Ltd.,
8F Omori Belport B, 6-26-2 MinamiOhi, Shinagawa-ku, Tokyo 140-0013, Japan
2Graduate School of Humanities and Sciences, Ochanomizu University,
2-1-1 Ohtsuka, Bunkyo-ku, Tokyo 112-8610, Japan
Email: [email protected]: [email protected]
Abstract

We show that in some cases the coherent state can have a larger violation of the Leggett-Garg inequality (LGI) than the cat state by numerical calculations. To achieve this result, we consider the LGI of the cavity mode weakly coupled to a zero-temperature environment as a practical instance of the physical system. We assume that the bosonic mode undergoes dissipation because of an interaction with the environment but is not affected by dephasing. Solving the master equation exactly, we derive an explicit form of the violation of the inequality for both systems prepared initially in the coherent state |α|\alpha\rangle and the cat state (|α+|α)(|\alpha\rangle+|-\alpha\rangle). For the evaluation of the inequality, we choose the displaced parity operators characterized by a complex number β\beta. We look for the optimum parameter β\beta that lets the upper bound of the inequality be maximum numerically. Contrary to our expectations, the coherent state occasionally exhibits quantum quality more strongly than the cat state for the upper bound of the violation of the LGI in a specific range of three equally spaced measurement times (spacing τ\tau). Moreover, as we let τ\tau approach zero, the optimized parameter β\beta diverges and the LGI reveals intense singularity.

1 Introduction

A violation of Bell’s inequality illustrates the fact that quantum mechanics is essentially incompatible with classical mechanics [1]. To compute Bell’s inequality, first we prepare states of two qubits spatially separated, second observe them independently, and third evaluate expectation values by taking averages of the set of the measurements. The violation of Bell’s inequality tells us that quantum mechanics is able to exhibit strong correlations which cannot be explained by the hidden variable theory.

The Leggett-Garg inequality (LGI) was proposed by Leggett and Garg in 1985 for testing whether or not macroscopic coherence could be realized in the laboratory [2, 3, 4]. When we consider Bell’s inequality, we pay attention to the correlation between the two spatially separated qubits. By contrast, although the LGI is an analogue of Bell’s one, it examines a correlation of measurements at two different times for a single particle, and thus we can regard it as temporal Bell’s inequality.

To derive this inequality, Leggett and Garg introduced the following two assumptions. The first one is “macroscopic realism”. This assumption requires that results of observation are determined by hidden variables that are attributes of the observed system. Moreover, we postulate that the hidden variables are independent of the observation itself. The second one is “non-invasive measurability”. This premise implies that a result of measurement at time t2t_{2} is independent of a result of previous measurement at time t1(<t2)t_{1}(<t_{2}). Thus, we can specify the state of the system uniquely without disturbing the system.

In general, classical mechanics satisfies these two assumptions, so that dynamics of all systems that evolves according to the classical theory does not violate the LGI. However, the two assumptions do not always hold in the quantum mechanics. Thus, it is possible that the dynamics of a quantum system violates the LGI. In fact, concrete examples of quantum systems which violate the inequality have been already found theoretically [5, 6, 7, 8, 9, 10, 11]. Recently, experimental demonstrations of the violation of the LGI have been reported [12, 13, 14].

In the current paper, we investigate the LGI for a cavity mode weakly coupled to a zero-temperature environment. We make the boson system undergo dissipation due to the interaction with the environment but we assume that it does not receive dephasing. Because there is no thermal fluctuation of the external reservoir, we can solve the master equation of the boson system exactly in a closed-form expression [15, 16].

In the present paper, we focus on cases where initial states of the system are given by the coherent state |α|\alpha\rangle and the cat state (|α+|α)(|\alpha\rangle+|-\alpha\rangle) and compare their violations of the LGI. The cat state is also called the even coherent state and we can consider it to be a special one of the Yurke-Stoler coherent states [17, 18, 19, 20].

In order to evaluate the LGI, selection of operators for dichotomic observation is important. In the derivation of the inequality, we performs measurements of the dichotomic variable at three equally spaced times, t1=0t_{1}=0, t2=τt_{2}=\tau, and t3=2τt_{3}=2\tau, where τ>0\tau>0. In the present paper, we choose the displaced parity operators to measure the state of the boson system [21, 22, 23, 24, 25]. These operators are characterized by an arbitrary complex number β=reiθ\beta=re^{i\theta}. Thus, adjusting the two real parameters, rr and θ\theta, we can vary the violation of the LGI.

In the current paper, we concentrate on a maximization problem, where we look for the optimum values of the parameters, θ\theta and rr, for maximizing the violation. Because the LGI reflects the macroscopic coherence of the system of interest, we can regard its violation as a quantity of the quantum feature of the system. One of the motivations of the present paper is to clarify which state exhibits a characteristic of quantum nature more strongly, the coherent state or the cat state.

We obtain the following result: the violation of the inequality for the coherent state is larger than that for the cat state if we let the time difference τ\tau of the LGI be a value belonging to a specific range. This result is unexpected because the cat state is a superposition of two different coherent states, |α|\alpha\rangle and |α|-\alpha\rangle, and we can consider intuitively that the cat state (|α+|α)(|\alpha\rangle+|-\alpha\rangle) exhibits the characteristic of quantum nature more intensely than the ordinary coherent state |α|\alpha\rangle. This counter intuitive fact is novel and interesting.

Moreover, if we let the time difference τ\tau approach zero, the parameter β\beta diverges to infinity as |β|=r|\beta|=r\to\infty for the optimized displaced parity operators and the maximized violation of the inequality converges to 3/23/2 but not unity, so that the LGI shows strong singularity. Because such behaviour of the violation does not occur in the inequality for a two-level system (i.e. using projection operators as observables, the upper bound of the inequality for the two-level system gets closer to unity as the time difference τ\tau approaches zero), we can guess that the origin of the singularity is the infiniteness of the dimension of the Hilbert space for the boson system.

Here, we mention previous works. In the current paper, we study the LGI for a cavity mode interacting with a zero-temperature environment. In Ref. [7], Chen et al. investigated the LGI of a qubit coupled to a zero-temperature environment. In Refs. [8, 9], the LGI of a qubit interacting with a thermal environment was studied. In Ref. [11], Thenabadu and Reid discussed the violation of the LGI of the cat state of the boson system but they did not consider the interaction with an environment.

In Ref. [26], two inequalities inspired by the LGI were derived to distinguish quantum from classical transport through nanostructures. In Refs. [27, 28], the extended LGI proposed in Ref. [26] was utilized on a double quantum dot and a multiple-quantum-well structure. In Ref. [29], new practical quantum witnesses to verify quantum coherence for complex systems were proposed and they were regarded as refined ones compared with the LGI. In Ref. [30], the violation of the LGI in electronic Mach-Zehnder interferometers was investigated.

The current paper is organized as follows. In Sect. 2, we review the master equation of the boson system, the LGI, and the displaced parity operators. In Sect. 3, we derive the explicit form of the LGI for the system prepared initially in the coherent state. In Sect. 4, we derive the rigorous form of the inequality with assuming that the system is initially in the cat state. In Sect. 5, we estimate the violation of the inequality numerically in the case where the system is initialized in the coherent state |α|\alpha\rangle. In Sect. 6, we evaluate the violation numerically in the case where the system is initially put in the cat state (|α+|α)(|\alpha\rangle+|-\alpha\rangle). In Sects. 5 and 6, we perform optimization of the displaced parity operators. In Sect. 7, we compare the maximized violations of the inequality for the two initial states, the coherent and cat states. In Sect. 8, we give brief discussion. In Appendices A and B, we give explicit mathematical forms that appear in Sects. 3 and 4, respectively.

2 Reviews of the master equation for the boson system, the LGI, and the displaced parity operators

In the current section, first of all, we review the master equation for the boson system and its exact solution [15, 16]. Next, we introduce the LGI [2, 3, 4] and the displaced parity operators [21, 22, 23, 24, 25].

The master equation of the boson system weakly coupled to a zero-temperature environment is given by

ρ˙(t)=i[H,ρ(t)]+Γ[2aρ(t)aaaρ(t)ρ(t)aa],\dot{\rho}(t)=-i[H,\rho(t)]+\Gamma[2a\rho(t)a^{\dagger}-a^{\dagger}a\rho(t)-\rho(t)a^{\dagger}a], (1)
H=ωaa,H=\omega a^{\dagger}a, (2)

where we put =1\hbar=1, and Γ\Gamma and ω\omega represent the spontaneous emission rate and the angular velocity, respectively. We assume that Γ\Gamma and ω\omega are real. Here, we consider how to solve Eq. (1) rigorously.

Taking the interaction picture, we simplify the master equation (1) as

ρ˙(t)=Γ[2aρ(t)aaaρ(t)ρ(t)aa].\dot{\rho}(t)=\Gamma[2a\rho(t)a^{\dagger}-a^{\dagger}a\rho(t)-\rho(t)a^{\dagger}a]. (3)

We introduce superoperators J^\hat{J} and L^\hat{L}, where we put hats on the symbols to emphasize that they are superoperators, as follows:

J^ρ(t)=aρ(t)a,\hat{J}\rho(t)=a\rho(t)a^{\dagger}, (4)
L^ρ(t)=(1/2)[aaρ(t)+ρ(t)aa].\hat{L}\rho(t)=-(1/2)[a^{\dagger}a\rho(t)+\rho(t)a^{\dagger}a]. (5)

Then, the master equation (3) can be rewritten in the form,

ρ˙(t)=2Γ(J^+L^)ρ(t).\dot{\rho}(t)=2\Gamma(\hat{J}+\hat{L})\rho(t). (6)

Using a commutation relation [J^,L^]=J^[\hat{J},\hat{L}]=-\hat{J}, we obtain a formal solution of Eq. (6),

ρ(t)\displaystyle\rho(t) =\displaystyle= exp[2Γt(J^+L^)]ρ(0)\displaystyle\exp[2\Gamma t(\hat{J}+\hat{L})]\rho(0) (7)
=\displaystyle= exp(2ΓtL^)exp{[1exp(2Γt)]J^}ρ(0).\displaystyle\exp(2\Gamma t\hat{L})\exp\{[1-\exp(-2\Gamma t)]\hat{J}\}\rho(0).

Now, we substitute an operator v(0)=|αβ|v(0)=|\alpha\rangle\langle\beta| for ρ(0)\rho(0), where |α|\alpha\rangle and |β|\beta\rangle are coherent states and α\alpha and β\beta are arbitrary complex numbers. Then, v(t)v(t) is given by

v(t)\displaystyle v(t) =\displaystyle= exp{(1/2)(|α|2+|β|22αβ)[1exp(2Γt)]}\displaystyle\exp\{-(1/2)(|\alpha|^{2}+|\beta|^{2}-2\alpha\beta^{*})[1-\exp(-2\Gamma t)]\} (8)
×|αexp(Γt)βexp(Γt)|.\displaystyle\times|\alpha\exp(-\Gamma t)\rangle\langle\beta\exp(-\Gamma t)|.

Here, we change the interaction picture of Eq. (8) into the Schrödinger picture. Finally, we attain

v(t)\displaystyle v(t) =\displaystyle= exp{(1/2)(|α|2+|β|22αβ)[1exp(2Γt)]}\displaystyle\exp\{-(1/2)(|\alpha|^{2}+|\beta|^{2}-2\alpha\beta^{*})[1-\exp(-2\Gamma t)]\} (9)
×|αexp(iΩt)βexp(iΩt)|,\displaystyle\times|\alpha\exp(-i\Omega t)\rangle\langle\beta\exp(-i\Omega t)|,

where Ω=ωiΓ\Omega=\omega-i\Gamma.

We pay attention to the following facts. If we put v(0)=|αβ|v(0)=|\alpha\rangle\langle\beta| and v~(0)=|βα|\tilde{v}(0)=|\beta\rangle\langle\alpha|, v~(t)=v(t)\tilde{v}(t)^{\dagger}=v(t) holds. Moreover, if we set ρ(0)=|αα|\rho(0)=|\alpha\rangle\langle\alpha|, we obtain
ρ(t)=|αexp(iΩt)αexp(iΩt)|\rho(t)=|\alpha\exp(-i\Omega t)\rangle\langle\alpha\exp(-i\Omega t)|.

The LGI is defined as follows. For the sake of simplicity, we assume that the dichotomic observables are given by projection operators. We consider an operator O^\hat{O} whose eigenvalues are equal to ±1\pm 1. To emphasize that it is an operator, we put a hat on the symbol. Next, we introduce equally spaced three times, t1=0t_{1}=0, t2=τt_{2}=\tau, and t3=2τt_{3}=2\tau, where τ>0\tau>0. We describe an observed value of the measurement with O^\hat{O} at time t1t_{1} as O1O_{1}. Obviously, O1=±1O_{1}=\pm 1 holds. We write a probability that observed values at times t1t_{1} and t2t_{2} are equal to O1O_{1} and O2O_{2} respectively as P21(O1,O2)P_{21}(O_{1},O_{2}). We define the correlation function C21C_{21} as

C21=O1,O2{1,+1}O2O1P21(O1,O2).C_{21}=\sum_{O_{1},O_{2}\in\{-1,+1\}}O_{2}O_{1}P_{21}(O_{1},O_{2}). (10)

Then, the LGI is given by

K3=C21+C32C31,K_{3}=C_{21}+C_{32}-C_{31}, (11)
3K31.-3\leq K_{3}\leq 1. (12)

We define projection operators which are called the displaced parity operators as

Π(+)(β)\displaystyle\Pi^{(+)}(\beta) =\displaystyle= D(β)n=0|2n2n|D(β),\displaystyle D(\beta)\sum_{n=0}^{\infty}|2n\rangle\langle 2n|D^{\dagger}(\beta),
Π()(β)\displaystyle\Pi^{(-)}(\beta) =\displaystyle= D(β)n=0|2n+12n+1|D(β),\displaystyle D(\beta)\sum_{n=0}^{\infty}|2n+1\rangle\langle 2n+1|D^{\dagger}(\beta), (13)

where

D(β)=exp(βaβa).D(\beta)=\exp(\beta a^{\dagger}-\beta^{*}a). (14)

In the present paper, hereafter, we choose the following operator for the orthogonal measurement of the LGI,

O^=Π(+)(β)Π()(β).\hat{O}=\Pi^{(+)}(\beta)-\Pi^{(-)}(\beta). (15)

3 The explicit form of the LGI for the system initially prepared in the coherent state

In the present section, putting a coherent state |α|\alpha\rangle as the initial state, we derive a closed-form expression of the LGI.

The probability that we obtain O1=1O_{1}=1 for the measurement of the initial state |α|\alpha\rangle at time t1=0t_{1}=0 is given by

P1+\displaystyle P_{1+} =\displaystyle= α|Π(+)(β)|α\displaystyle\langle\alpha|\Pi^{(+)}(\beta)|\alpha\rangle (16)
=\displaystyle= exp(|αβ|2)cosh|αβ|2.\displaystyle\exp(-|\alpha-\beta|^{2})\cosh|\alpha-\beta|^{2}.

Then, the state collapses from |α|\alpha\rangle to w1+(0)/P1+w_{1+}(0)/P_{1+}, where w1+(0)w_{1+}(0) is given by

w1+(0)=Π(+)(β)|αα|Π(+)(β),w_{1+}(0)=\Pi^{(+)}(\beta)|\alpha\rangle\langle\alpha|\Pi^{(+)}(\beta), (17)
Π(+)(β)|α=(1/2)[|α+exp(βαβα)|2βα].\Pi^{(+)}(\beta)|\alpha\rangle=(1/2)[|\alpha\rangle+\exp(\beta^{*}\alpha-\beta\alpha^{*})|2\beta-\alpha\rangle]. (18)

In the above derivation, we use a relation,

n=0|2n2n|αβ=(1/2)[|αβ+|(αβ)].\sum_{n=0}^{\infty}|2n\rangle\langle 2n|\alpha-\beta\rangle=(1/2)[|\alpha-\beta\rangle+|-(\alpha-\beta)\rangle]. (19)

Similarly, the probability that we obtain O1=1O_{1}=-1 for the measurement of the initial state |α|\alpha\rangle at time t1=0t_{1}=0 is written in the form,

P1\displaystyle P_{1-} =\displaystyle= α|Π()(β)|α\displaystyle\langle\alpha|\Pi^{(-)}(\beta)|\alpha\rangle (20)
=\displaystyle= exp(|αβ|2)sinh|αβ|2.\displaystyle\exp(-|\alpha-\beta|^{2})\sinh|\alpha-\beta|^{2}.

Then, the initial state |α|\alpha\rangle reduces to w1(0)/P1w_{1-}(0)/P_{1-}, where w1(0)w_{1-}(0) is given by

w1(0)=Π()(β)|αα|Π()(β),w_{1-}(0)=\Pi^{(-)}(\beta)|\alpha\rangle\langle\alpha|\Pi^{(-)}(\beta), (21)
Π()(β)|α=(1/2)[|αexp(βαβα)|2βα].\Pi^{(-)}(\beta)|\alpha\rangle=(1/2)[|\alpha\rangle-\exp(\beta^{*}\alpha-\beta\alpha^{*})|2\beta-\alpha\rangle]. (22)

In the above derivation, we use a relation,

n=0|2n+12n+1|αβ=(1/2)[|αβ|(αβ)].\sum_{n=0}^{\infty}|2n+1\rangle\langle 2n+1|\alpha-\beta\rangle=(1/2)[|\alpha-\beta\rangle-|-(\alpha-\beta)\rangle]. (23)

According to Eq. (9), w1±(0)w_{1\pm}(0), the unnormalized state at time t1=0t_{1}=0, evolves into the state w1±(τ)w_{1\pm}(\tau) at time t2=τt_{2}=\tau, whose explicit form is given by Eq. (74) in Appendix A. Describing the probability that we detect O2=1O_{2}=1 with the observation of w1±(τ)w_{1\pm}(\tau) at time t2=τt_{2}=\tau as p1±,2+p_{1\pm,2+}, we obtain it in the form of Eq. (75) in Appendix A. Similarly, writing down the probability that we have O2=1O_{2}=-1 with the observation of w1±(τ)w_{1\pm}(\tau) at time t2=τt_{2}=\tau as p1±,2p_{1\pm,2-}, we acquire it in the form of Eq. (76) in Appendix A.

From Eqs. (75) and (76), the correlation function C21C_{21} is given by

C21=p1+,2+p1+,2p1,2++p1,2.C_{21}=p_{1+,2+}-p_{1+,2-}-p_{1-,2+}+p_{1-,2-}. (24)

Here, we regard C21C_{21} as a function of multiple variables,

C21=C(α,β,ω,Γ,τ).C_{21}=C(\alpha,\beta,\omega,\Gamma,\tau). (25)

Then, C31C_{31} and C32C_{32} are given by

C31=C(α,β,ω,Γ,2τ),C_{31}=C(\alpha,\beta,\omega,\Gamma,2\tau), (26)
C32=C(αeiΩτ,β,ω,Γ,τ).C_{32}=C(\alpha e^{-i\Omega\tau},\beta,\omega,\Gamma,\tau). (27)

Thus, from Eqs. (24), (25), (26), and (27), we obtain K3K_{3} eventually.

4 The explicit form of the LGI for the system initialized in the cat state

In this section, preparing the cat state (|α+|α)(|\alpha\rangle+|-\alpha\rangle) as the initial state, we derive a mathematically rigorous form of the LGI.

We assume that the initial state of the system at time t1=0t_{1}=0 is given by the cat state,

ρ(t1)=|ψψ|,\rho(t_{1})=|\psi\rangle\langle\psi|, (28)
|ψ=q(α)1/2(|α+|α),|\psi\rangle=q(\alpha)^{-1/2}(|\alpha\rangle+|-\alpha\rangle), (29)
q(α)=2[1+exp(2|α|2)],q(\alpha)=2[1+\exp(-2|\alpha|^{2})], (30)

where |α|\alpha\rangle and |α-|\alpha\rangle are coherent states. It is convenient to use the following notation for calculations carried out hereafter:

ρ(t1)=q(α)1(K+L+L+M),\rho(t_{1})=q(\alpha)^{-1}(K+L+L^{\dagger}+M), (31)
K\displaystyle K =\displaystyle= |αα|,\displaystyle|\alpha\rangle\langle\alpha|,
L\displaystyle L =\displaystyle= |αα|,\displaystyle|\alpha\rangle\langle-\alpha|,
M\displaystyle M =\displaystyle= |αα|.\displaystyle|-\alpha\rangle\langle-\alpha|. (32)

The probability that we obtain O1=±1O_{1}=\pm 1 with the observation of ρ(t1)\rho(t_{1}) at time t1=0t_{1}=0 is written down in the form,

P1±=Tr[Π(±)(β)ρ(t1)].P_{1\pm}=\mbox{Tr}[\Pi^{(\pm)}(\beta)\rho(t_{1})]. (33)

Then, the state collapses from ρ(t1)\rho(t_{1}) to w1±(0)/P1±w_{1\pm}(0)/P_{1\pm}, where w1±(0)w_{1\pm}(0) is given by

w1±(0)\displaystyle w_{1\pm}(0) =\displaystyle= Π(±)(β)ρ(t1)Π(±)(β)\displaystyle\Pi^{(\pm)}(\beta)\rho(t_{1})\Pi^{(\pm)}(\beta) (34)
=\displaystyle= q(α)1Π(±)(β)(K+L+L+M)Π(±)(β).\displaystyle q(\alpha)^{-1}\Pi^{(\pm)}(\beta)(K+L+L^{\dagger}+M)\Pi^{(\pm)}(\beta).

Moreover, we divide operators KK, LL, and MM into operators as

Π(±)(β)KΠ(±)(β)\displaystyle\Pi^{(\pm)}(\beta)K\Pi^{(\pm)}(\beta) =\displaystyle= (1/4)[K(1)(0)±K(2)(0)±K(3)(0)+K(4)(0)],\displaystyle(1/4)[K^{(1)}(0)\pm K^{(2)}(0)\pm K^{(3)}(0)+K^{(4)}(0)],
Π(±)(β)LΠ(±)(β)\displaystyle\Pi^{(\pm)}(\beta)L\Pi^{(\pm)}(\beta) =\displaystyle= (1/4)[L(1)(0)±L(2)(0)±L(3)(0)+L(4)(0)],\displaystyle(1/4)[L^{(1)}(0)\pm L^{(2)}(0)\pm L^{(3)}(0)+L^{(4)}(0)],
Π(±)(β)MΠ(±)(β)\displaystyle\Pi^{(\pm)}(\beta)M\Pi^{(\pm)}(\beta) =\displaystyle= (1/4)[M(1)(0)±M(2)(0)±M(3)(0)+M(4)(0)],\displaystyle(1/4)[M^{(1)}(0)\pm M^{(2)}(0)\pm M^{(3)}(0)+M^{(4)}(0)], (35)

where K(3)(0)=K(2)(0)K^{(3)}(0)=K^{(2)}(0)^{\dagger} and M(3)(0)=M(2)(0)M^{(3)}(0)=M^{(2)}(0)^{\dagger}. We give explicit forms of {K(j)(0):j=1,2,4}\{K^{(j)}(0):j=1,2,4\}, {L(j)(0):j=1,2,3,4}\{L^{(j)}(0):j=1,2,3,4\}, and {M(j)(0):j=1,2,4}\{M^{(j)}(0):j=1,2,4\} in Eqs. (77), (78), and (79) in Appendix B. Closed-form expressions of time evolution from time t1=0t_{1}=0 to time t2=τt_{2}=\tau of these operators, {K(j)(τ):j=1,2,4}\{K^{(j)}(\tau):j=1,2,4\}, {L(j)(τ):j=1,2,3,4}\{L^{(j)}(\tau):j=1,2,3,4\}, and {M(j)(τ):j=1,2,4}\{M^{(j)}(\tau):j=1,2,4\}, are given by Eqs. (80), (81), and (82) in Appendix B explicitly.

We consider that the state w1±(0)w_{1\pm}(0) at time t1=0t_{1}=0 evolves into w1±(τ)w_{1\pm}(\tau) at time t2=τt_{2}=\tau. Then, we can obtain the state w1±(τ)w_{1\pm}(\tau) by replacing {K(j)(0):j=1,2,4}\{K^{(j)}(0):j=1,2,4\}, {L(j)(0):j=1,2,3,4}\{L^{(j)}(0):j=1,2,3,4\}, and {M(j)(0):j=1,2,4}\{M^{(j)}(0):j=1,2,4\} in Eqs. (34) and (35) with {K(j)(τ):j=1,2,4}\{K^{(j)}(\tau):j=1,2,4\}, {L(j)(τ):j=1,2,3,4}\{L^{(j)}(\tau):j=1,2,3,4\}, and {M(j)(τ):j=1,2,4}\{M^{(j)}(\tau):j=1,2,4\} given by Eqs. (80), (81), and (82). Describing the probability that we detect O2=1O_{2}=1 with the observation of the state w1±(τ)w_{1\pm}(\tau) at time t2=τt_{2}=\tau as p1±,2+p_{1\pm,2+}, we obtain it in the form of Eq. (LABEL:p1pm2p) in Appendix B.

Letting the probability that we obtain O2=1O_{2}=-1 with the observation of w1±(τ)w_{1\pm}(\tau) at time t2=τt_{2}=\tau be equal to p1±,2p_{1\pm,2-}, we can write down it as Eq. (LABEL:p1pm2m) in Appendix B. Then, we can derive the correlation function C21C_{21} from Eq. (24). We can also obtain C31C_{31} from Eqs. (25) and (26).

Next, we think about the correlation function C32C_{32}. Here, we remark that we cannot obtain C32C_{32} from Eq. (27). The reason why is going to be clarified by Eqs.(36) and (37). The initial state at time t1=0t_{1}=0 is given by Eqs. (28), (29), and (30). The initial state ρ(t1)\rho(t_{1}) evolves into the following state at time t2=τt_{2}=\tau:

ρ(t2)\displaystyle\rho(t_{2}) =\displaystyle= q(α)1(K~+exp{2|α|2[1exp(2Γτ)]}(L~+L~)+M~),\displaystyle q(\alpha)^{-1}\Bigl{(}\tilde{K}+\exp\{-2|\alpha|^{2}[1-\exp(-2\Gamma\tau)]\}(\tilde{L}+\tilde{L}^{\dagger})+\tilde{M}\Bigr{)}, (36)
K~\displaystyle\tilde{K} =\displaystyle= |αexp(iΩτ)αexp(iΩτ)|,\displaystyle|\alpha\exp(-i\Omega\tau)\rangle\langle\alpha\exp(-i\Omega\tau)|,
L~\displaystyle\tilde{L} =\displaystyle= |αexp(iΩτ)αexp(iΩτ)|,\displaystyle|\alpha\exp(-i\Omega\tau)\rangle\langle-\alpha\exp(-i\Omega\tau)|,
M~\displaystyle\tilde{M} =\displaystyle= |αexp(iΩτ)αexp(iΩτ)|.\displaystyle|-\alpha\exp(-i\Omega\tau)\rangle\langle-\alpha\exp(-i\Omega\tau)|. (37)

Looking at Eqs. (36) and (37), we notice that ρ(t2)\rho(t_{2}) is different from ρ=|ψψ|\rho^{\prime}=|\psi^{\prime}\rangle\langle\psi^{\prime}| where |ψ=q(α)1/2(|α+|α)|\psi^{\prime}\rangle=q(\alpha^{\prime})^{-1/2}(|\alpha^{\prime}\rangle+|-\alpha^{\prime}\rangle) and α=αexp(iΩτ)\alpha^{\prime}=\alpha\exp(-i\Omega\tau). Thus, we cannot derive C32C_{32} from Eq. (27).

The probability that we obtain O2=±1O_{2}=\pm 1 with the observation of ρ(t2)\rho(t_{2}) at time t2=τt_{2}=\tau is given by

P2±=Tr[Π(±)(β)ρ(t2)].P_{2\pm}=\mbox{Tr}[\Pi^{(\pm)}(\beta)\rho(t_{2})]. (38)

Then, the state of the system ρ(t2)\rho(t_{2}) reduces to w2±(τ)/P2±w_{2\pm}(\tau)/P_{2\pm}, where w2±(τ)w_{2\pm}(\tau) is given by

w2±(τ)\displaystyle w_{2\pm}(\tau) =\displaystyle= Π(±)(β)ρ(t2)Π(±)(β)\displaystyle\Pi^{(\pm)}(\beta)\rho(t_{2})\Pi^{(\pm)}(\beta)
=\displaystyle= q(α)1Π(±)(β)(K~+exp{2|α|2[1exp(2Γτ)]}(L~+L~)+M~)Π(±)(β).\displaystyle q(\alpha)^{-1}\Pi^{(\pm)}(\beta)\Bigl{(}\tilde{K}+\exp\{-2|\alpha|^{2}[1-\exp(-2\Gamma\tau)]\}(\tilde{L}+\tilde{L}^{\dagger})+\tilde{M}\Bigr{)}\Pi^{(\pm)}(\beta).

We can divide Π(±)(β)K~Π(±)(β)\Pi^{(\pm)}(\beta)\tilde{K}\Pi^{(\pm)}(\beta), Π(±)(β)L~Π(±)(β)\Pi^{(\pm)}(\beta)\tilde{L}\Pi^{(\pm)}(\beta), and Π(±)(β)M~Π(±)(β)\Pi^{(\pm)}(\beta)\tilde{M}\Pi^{(\pm)}(\beta) into operators,

Π(±)(β)K~Π(±)(β)\displaystyle\Pi^{(\pm)}(\beta)\tilde{K}\Pi^{(\pm)}(\beta) =\displaystyle= (1/4)[K~(1)(0)±K~(2)(0)±K~(2)(0)+K~(4)(0)],\displaystyle(1/4)[\tilde{K}^{(1)}(0)\pm\tilde{K}^{(2)}(0)\pm\tilde{K}^{(2)}(0)^{\dagger}+\tilde{K}^{(4)}(0)], (40)
K~(j)(0)=K(j)(0)|ααexp(iΩτ)for j=1,2,3,4,\tilde{K}^{(j)}(0)=\left.K^{(j)}(0)\right|_{\alpha\rightarrow\alpha\exp(-i\Omega\tau)}\quad\mbox{for $j=1,2,3,4$}, (41)
Π(±)(β)L~Π(±)(β)\displaystyle\Pi^{(\pm)}(\beta)\tilde{L}\Pi^{(\pm)}(\beta) =\displaystyle= (1/4)[L~(1)(0)±L~(2)(0)±L~(2)(0)+L~(4)(0)],\displaystyle(1/4)[\tilde{L}^{(1)}(0)\pm\tilde{L}^{(2)}(0)\pm\tilde{L}^{(2)}(0)^{\dagger}+\tilde{L}^{(4)}(0)], (42)
L~(j)(0)=L(j)(0)|ααexp(iΩτ)for j=1,2,3,4,\tilde{L}^{(j)}(0)=\left.L^{(j)}(0)\right|_{\alpha\rightarrow\alpha\exp(-i\Omega\tau)}\quad\mbox{for $j=1,2,3,4$}, (43)
Π(±)(β)M~Π(±)(β)\displaystyle\Pi^{(\pm)}(\beta)\tilde{M}\Pi^{(\pm)}(\beta) =\displaystyle= (1/4)[M~(1)(0)±M~(2)(0)±M~(2)(0)+M~(4)(0)],\displaystyle(1/4)[\tilde{M}^{(1)}(0)\pm\tilde{M}^{(2)}(0)\pm\tilde{M}^{(2)}(0)^{\dagger}+\tilde{M}^{(4)}(0)], (44)
M~(j)(0)=M(j)(0)|ααexp(iΩτ)for j=1,2,3,4.\tilde{M}^{(j)}(0)=\left.M^{(j)}(0)\right|_{\alpha\rightarrow\alpha\exp(-i\Omega\tau)}\quad\mbox{for $j=1,2,3,4$}. (45)

For example, Eq. (41) implies the following. The operator K~(j)(0)\tilde{K}^{(j)}(0) is equal to the operator K(j)(0)K^{(j)}(0) with the replacement of α\alpha by αexp(iΩτ)\alpha\exp(-i\Omega\tau) in Eq. (77).

Referring to Eqs. (80), (81), and (82), closed-form expressions of time evolution from time t2=τt_{2}=\tau to time t3=2τt_{3}=2\tau of {K~(j)(0):j=1,2,3,4}\{\tilde{K}^{(j)}(0):j=1,2,3,4\}, {L~(j)(0):j=1,2,3,4}\{\tilde{L}^{(j)}(0):j=1,2,3,4\}, and {M~(j)(0):j=1,2,3,4}\{\tilde{M}^{(j)}(0):j=1,2,3,4\} are given by

K~(j)(τ)\displaystyle\tilde{K}^{(j)}(\tau) =\displaystyle= K(j)(τ)|ααexp(iΩτ),\displaystyle\left.K^{(j)}(\tau)\right|_{\alpha\rightarrow\alpha\exp(-i\Omega\tau)},
L~(j)(τ)\displaystyle\tilde{L}^{(j)}(\tau) =\displaystyle= L(j)(τ)|ααexp(iΩτ),\displaystyle\left.L^{(j)}(\tau)\right|_{\alpha\rightarrow\alpha\exp(-i\Omega\tau)},
M~(j)(τ)\displaystyle\tilde{M}^{(j)}(\tau) =\displaystyle= M(j)(τ)|ααexp(iΩτ)for j=1,2,3,4.\displaystyle\left.M^{(j)}(\tau)\right|_{\alpha\rightarrow\alpha\exp(-i\Omega\tau)}\quad\mbox{for $j=1,2,3,4$}. (46)

Thus, the state w2±(2τ)w_{2\pm}(2\tau), that is to say, the time evolution from time t2=τt_{2}=\tau to time t3=2τt_{3}=2\tau of the state w2±(τ)w_{2\pm}(\tau), is obtained by replacing {K~(j)(0):j=1,2,3,4}\{\tilde{K}^{(j)}(0):j=1,2,3,4\}, {L~(j)(0):j=1,2,3,4}\{\tilde{L}^{(j)}(0):j=1,2,3,4\}, and {M~(j)(0):j=1,2,3,4}\{\tilde{M}^{(j)}(0):j=1,2,3,4\} with Eq. (46) in Eqs. (LABEL:w-2pm-tau-definition), (40), (41), (42), (43), (44), and (45).

We describe the probability that we obtain O3=1O_{3}=1 with the observation of w2±(2τ)w_{2\pm}(2\tau) at time t3=2τt_{3}=2\tau as p2±,3+p_{2\pm,3+}. Similarly, we let p2±,3p_{2\pm,3-} represent the probability that we obtain O3=1O_{3}=-1 with the observation of w2±(2τ)w_{2\pm}(2\tau) at time t3=2τt_{3}=2\tau. Then, they are given by Eqs. (LABEL:p2pm2p-formula) and (LABEL:p2pm2m-formula) in Appendix B.

Thus, the correlation function C32C_{32} is given in the form,

C32=p2+,3+p2+,3p2,3++p2,3.C_{32}=p_{2+,3+}-p_{2+,3-}-p_{2-,3+}+p_{2-,3-}. (47)

Because we have just derived C21C_{21}, C31C_{31}, and C32C_{32}, finally we obtain K3K_{3}.

Here, we point out that K3K_{3} has the following symmetry about the variable β\beta. The value of K3K_{3} is invariant under a transformation ββ\beta\rightarrow-\beta. This fact is understood from relations,

Tr[K(1)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[K^{(1)}(\tau)\Pi^{(+)}(-\beta)] =\displaystyle= Tr[M(1)(τ)Π(+)(β)],\displaystyle\mbox{Tr}[M^{(1)}(\tau)\Pi^{(+)}(\beta)],
Tr[K(2)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[K^{(2)}(\tau)\Pi^{(+)}(-\beta)] =\displaystyle= Tr[M(2)(τ)Π(+)(β)],\displaystyle\mbox{Tr}[M^{(2)}(\tau)\Pi^{(+)}(\beta)],
Tr[K(4)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[K^{(4)}(\tau)\Pi^{(+)}(-\beta)] =\displaystyle= Tr[M(4)(τ)Π(+)(β)],\displaystyle\mbox{Tr}[M^{(4)}(\tau)\Pi^{(+)}(\beta)],
Tr[L(1)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[L^{(1)}(\tau)\Pi^{(+)}(-\beta)] =\displaystyle= Tr[L(1)(τ)Π(+)(β)],\displaystyle\mbox{Tr}[L^{(1)}(\tau)\Pi^{(+)}(\beta)]^{*},
Tr[L(2)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[L^{(2)}(\tau)\Pi^{(+)}(-\beta)] =\displaystyle= Tr[L(3)(τ)Π(+)(β)],\displaystyle\mbox{Tr}[L^{(3)}(\tau)\Pi^{(+)}(\beta)]^{*},
Tr[L(3)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[L^{(3)}(\tau)\Pi^{(+)}(-\beta)] =\displaystyle= Tr[L(2)(τ)Π(+)(β)],\displaystyle\mbox{Tr}[L^{(2)}(\tau)\Pi^{(+)}(\beta)]^{*},
Tr[L(4)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[L^{(4)}(\tau)\Pi^{(+)}(-\beta)] =\displaystyle= Tr[L(4)(τ)Π(+)(β)].\displaystyle\mbox{Tr}[L^{(4)}(\tau)\Pi^{(+)}(\beta)]^{*}. (48)

Hence, putting β=rexp(iθ)\beta=r\exp(i\theta) and examining a value of K3K_{3} with variation of β\beta, we only have to try θ\theta in a range of 0θπ0\leq\theta\leq\pi. In other words, K3K_{3} is a periodic function of θ\theta and its period is equal to π\pi.

5 Numerical analyses of the LGI and its optimization with initially putting the system in a coherent state |α|\alpha\rangle

In the current section, preparing the system initially in a coherent state |α|\alpha\rangle for the LGI, we compute K3K_{3} numerically and examine its properties. Moreover, we go into the optimization problem of the displaced parity operators. In order to let discussion of the current section be simple, we assume

α=1/2,ω=1.\alpha=1/2,\quad\omega=1. (49)

Because we set ω=1\omega=1, that is we let the angular velocity be equal to unity, we can adopt the following system of units. Because of ω=1\omega=1, we can rewrite a dimensionless quantity ωτ\omega\tau as τ\tau. Thus, we can regard the time variable τ\tau as dimensionless. Similarly, we can rewrite a dimensionless quantity Γ/ω\Gamma/\omega as Γ\Gamma, so that we can consider Γ\Gamma to be dimensionless, as well. Hereafter, we use this system of units.

Here, we define a simple notation as follows. Fixing the variables α\alpha and ω\omega at α=1/2\alpha=1/2 and ω=1\omega=1 respectively, we describe K3K_{3} as a function of Γ\Gamma, β=reiθ\beta=re^{i\theta}, and τ\tau,

K3(Γ;θ,r,τ).K_{3}(\Gamma;\theta,r,\tau). (50)
Refer to caption
Figure 1: Graphs of K3(Γ;0,1/2,τ)K_{3}(\Gamma;0,1/2,\tau) as a function of τ\tau with β=1/2\beta=1/2, that is θ=0\theta=0 and r=1/2r=1/2, where the system is initialized in the coherent state. The thick solid red, thin solid blue, and thin dashed purple curves represent plots of Γ=0\Gamma=0, 0.050.05, and 0.20.2, respectively. The thick solid red curve has a period 2π2\pi. We notice that amplitudes of the curves shrink as Γ\Gamma becomes larger.

In Fig. 1, we plot K3(Γ;0,1/2,τ)K_{3}(\Gamma;0,1/2,\tau) as a function of τ\tau with β=1/2\beta=1/2, that is θ=0\theta=0 and r=1/2r=1/2. The thick solid red, thin solid blue, and thin dashed purple curves represent graphs with Γ=0\Gamma=0, 0.050.05, and 0.20.2, respectively. The graph of the thick solid red curve has a period 2π2\pi. As Γ\Gamma becomes larger, amplitudes of the curves shrink.

We can confirm numerically that the curve of Γ=0.2\Gamma=0.2 converges to 0.367 8790.367{\,}879... as τ\tau increases. We can also validate this fact in the manner of analytical mathematics. Fixing Γ\Gamma at a finite positive value Γ0(>0)\Gamma_{0}(>0) and letting τ\tau approach infinity, τ\tau\to\infty, we obtain

limτK3(Γ0;0,1/2,τ)\displaystyle\lim_{\tau\rightarrow\infty}K_{3}(\Gamma_{0};0,1/2,\tau) =\displaystyle= 1/e\displaystyle 1/e (51)
\displaystyle\simeq 0.367 879.\displaystyle 0.367{\,}879....

From the above result, we understand that K3K_{3} converges to 1/e1/e for Γ0>0\Gamma_{0}>0 and τ\tau\rightarrow\infty.

In general, we can show the following. α\forall\alpha, ω\forall\omega, Γ(>0)\forall\Gamma(>0), and β=reiθ\forall\beta=re^{i\theta}, we obtain

limτK3=exp(4r2).\lim_{\tau\rightarrow\infty}K_{3}=\exp(-4r^{2}). (52)

We can explain that Eq. (52) does not depend on ω\omega from the following consideration. According to Eq. (9), the angular velocity ω\omega appears in the explicit expression of K3K_{3} as the form exp(iΩτ)\exp(-i\Omega\tau) with Ω=ωiΓ\Omega=\omega-i\Gamma. Thus, setting Γ>0\Gamma>0, we obtain exp(iΩτ)0\exp(-i\Omega\tau)\rightarrow 0 for τ\tau\rightarrow\infty and dependency of ω\omega disappears.

From Eq. (52), we become aware that K31K_{3}\leq 1 for τ\tau\rightarrow\infty. Thus, we understand that the violation of the LGI vanishes as τ\tau approaches infinity.

Refer to caption
Figure 2: Graphs of the optimum θ\theta that maximizes K3K_{3} for each τ(>0)\tau(>0) as a function of τ\tau with letting the system be initialized in the coherent state. The thick solid red, thin solid blue, and thin dashed purple curves represent plots of Γ=0\Gamma=0, 0.10.1, and 11, respectively. For all the three curves, we can find discontinuity points.
Refer to caption
Figure 3: A thick red curve represents a plot of the optimum θ\theta that maximizes K3K_{3} for each τ(>0)\tau(>0) as a function of τ\tau with Γ=0\Gamma=0, where the system is prepared initially in the coherent state. Three parallel thin blue lines represent graphs of θ=2πτ\theta=2\pi-\tau, θ=πτ\theta=\pi-\tau, and θ=τ\theta=-\tau. The optimized θ\theta moves and jumps on the lines of θ=πτ\theta=\pi-\tau, θ=τ\theta=-\tau, θ=π\theta=-\pi, θ=2πτ\theta=2\pi-\tau, and θ=πτ\theta=\pi-\tau in order as τ\tau increases from zero to 6.56.5. Because K3K_{3} is periodic about θ\theta and its period is given by 2π2\pi, the lines θ=2πτ\theta=2\pi-\tau and θ=τ\theta=-\tau are essentially equivalent to each other.
Refer to caption
Figure 4: Graphs of the optimum rr that maximizes K3K_{3} for each τ(>0)\tau(>0) as a function of τ\tau, where the system is initialized in the coherent state. The thick solid red, thin solid blue, and thin dashed purple curves represent plots of Γ=0\Gamma=0, 0.10.1, and 11, respectively. All the three curves seemingly diverge to infinity at τ=0\tau=0. Moreover, the thick red curve of Γ=0\Gamma=0 apparently diverges to infinity at τ=2π\tau=2\pi.
Refer to caption
Figure 5: Graphs of the maximized K3K_{3} with adjusting θ\theta and rr for each τ(>0)\tau(>0) as a function of τ\tau with letting the system be prepared initially in the coherent state. The thick solid red, thin solid blue, and thin dashed purple curves represent plots of Γ=0\Gamma=0, 0.10.1, and 11, respectively. As Γ\Gamma increases, amplitudes of the graphs are suppressed.

Next, we consider the following optimization problem. At given arbitrary τ(>0)\tau(>0), we look for values of θ\theta and rr which maximize K3K_{3}. In Figs. 2 and 4, we plot the optimum θ\theta and rr that maximize K3K_{3} for each τ(>0)\tau(>0) as functions of τ\tau. In Fig. 5, we draw a curve of the maximized K3K_{3} versus τ\tau. In Figs. 2, 4, and 5, the thick solid red, thin solid blue, and thin dashed purple curves represent plots for Γ=0\Gamma=0, 0.10.1, and 11, respectively. Moreover, in Fig. 3, we plot the optimum θ\theta that maximizes K3K_{3} for each τ(>0)\tau(>0) and Γ=0\Gamma=0 as a function of τ\tau with a thick red curve and draw θ=2πτ\theta=2\pi-\tau, θ=πτ\theta=\pi-\tau, and θ=τ\theta=-\tau with thin blue lines. Drawing graphs in Figs. 2, 3, 4, and 5, we divide a range 0<τ6.50<\tau\leq 6.5 into equal spaces as τ=nΔτ\tau=n\Delta\tau, Δτ=0.025\Delta\tau=0.025, and n=1,2,,260n=1,2,...,260 and optimize θ\theta and rr at each time τ=nΔτ\tau=n\Delta\tau.

Here, we concentrate on Fig. 3. The optimum θ\theta that maximizes K3K_{3} moves and jumps on three parallel lines, θ=2πτ\theta=2\pi-\tau, θ=πτ\theta=\pi-\tau, and θ=τ\theta=-\tau. In the following paragraphs, we analyse this fact in detail.

Carrying out slightly tough calculations, we can show a relation,

θK3(0;θ,r,τ)|θ=πτ\displaystyle\left.\frac{\partial}{\partial\theta}K_{3}(0;\theta,r,\tau)\right|_{\theta=\pi-\tau} =\displaystyle= θK3(0;θ,r,τ)|θ=τ\displaystyle\left.\frac{\partial}{\partial\theta}K_{3}(0;\theta,r,\tau)\right|_{\theta=-\tau} (53)
=\displaystyle= 0.\displaystyle 0.

Because K3(0;θ,r,τ)K_{3}(0;\theta,r,\tau) is a periodic function about θ\theta and its period is equal to 2π2\pi, it is obvious that (/θ)K3(0;θ,r,τ)|θ=2πτ=0\left.(\partial/\partial\theta)K_{3}(0;\theta,r,\tau)\right|_{\theta=2\pi-\tau}=0 holds. Hence, for three cases where θ=2πτ\theta=2\pi-\tau, θ=πτ\theta=\pi-\tau, and θ=τ\theta=-\tau, we can expect that K3(0;θ,r,τ)K_{3}(0;\theta,r,\tau) takes extreme values. Now, we examine this expectation concretely below.

Refer to caption
Figure 6: A graph of f(r)=(/r)K3(0;πτ,r,τ)|τ=0.05f(r)=\left.(\partial/\partial r)K_{3}(0;\pi-\tau,r,\tau)\right|_{\tau=0.05} versus rr. For r=0r=0 and r=1.989 01r=1.989{\,}01..., f(r)=0f(r)=0 holds.

Figure 6 shows a plot of f(r)f(r) versus rr, where f(r)f(r) is given by

f(r)=rK3(0;πτ,r,τ)|τ=0.05.f(r)=\left.\frac{\partial}{\partial r}K_{3}(0;\pi-\tau,r,\tau)\right|_{\tau=0.05}. (54)

Looking at Fig. 6, we notice that f(r)=0f(r)=0 holds at points, r=0r=0 and r=1.989 01r=1.989{\,}01.... For τ=0.05\tau=0.05, we can maximize K3K_{3} at r=1.989 01r=1.989{\,}01....

Because of these circumstances, the optimum θ\theta that maximizes K3K_{3} walks and jumps around three lines, θ=2πτ\theta=2\pi-\tau, θ=πτ\theta=\pi-\tau, and θ=τ\theta=-\tau. In Fig. 3, in a range of 2.150τ4.1502.150\leq\tau\leq 4.150, θ=π\theta=-\pi holds. In this range with Γ=0\Gamma=0, we can confirm that r=0r=0 holds by looking at Fig. 4. Thus, in the range of 2.150τ4.1502.150\leq\tau\leq 4.150, we can consider that the value of θ\theta is meaningless.

Around a neighbourhood of τ=0\tau=0, the function of the maximized K3K_{3} exhibits strong singularity. According to Eq. (11), the definition of K3K_{3}, we can naively suppose

K3(0;θ,r,0)=1.K_{3}(0;\theta,r,0)=1. (55)

However, inspecting Fig. 5, we recognize that the maximized K3K_{3} approaches 1.51.5 in the limit τ+0\tau\rightarrow+0. By numerical calculations, we can verify that the maximized K3K_{3} attains 1.498 481.498{\,}48... for τ=0.001\tau=0.001. From these careful looks, we grasp that it is very difficult to estimate the maximum value of K3K_{3} in the limit τ+0\tau\to+0. In fact, Fig. 4 shows that the optimized rr apparently diverges to infinity, rr\rightarrow\infty, as τ\tau approaches zero, τ+0\tau\rightarrow+0. This fact makes the problem be very intractable. In the following paragraphs, we consider this problem carefully.

Refer to caption
Figure 7: A plot of g(r,τ)g(r,\tau) given by Eq. (56) on a two-dimensional plane of (τ,r)(\tau,r). We can find strong singularity in the limits, τ+0\tau\rightarrow+0 and rr\rightarrow\infty.

Paying our attention to Fig. 3, we can recognize that the optimum θ\theta that maximizes K3K_{3} is given by θ=πτ\theta=\pi-\tau. Thus, we consider a function,

g(r,τ)\displaystyle g(r,\tau) =\displaystyle= K3(0;πτ,r,τ)\displaystyle K_{3}(0;\pi-\tau,r,\tau) (56)
=\displaystyle= 2exp[4r2(1+cosτ)]cos[2r(1+2r)sinτ]\displaystyle 2\exp[4r^{2}(-1+\cos\tau)]\cos[2r(1+2r)\sin\tau]
exp(8r2sin2τ)cos[4r(1+2rcosτ)sinτ].\displaystyle-\exp(-8r^{2}\sin^{2}\tau)\cos[4r(1+2r\cos\tau)\sin\tau].

Figure 7 shows a graph of g(r,τ)g(r,\tau) plotted in a two-dimensional plane of (τ,r)(\tau,r). In Fig. 7, we can find strong singularity in the limits, τ+0\tau\rightarrow+0 and rr\rightarrow\infty, and we cannot estimate the maximum value of g(r,τ)g(r,\tau) under these conditions.

Here, we apply the following approximations to g(r,τ)g(r,\tau). Taking the limit τ+0\tau\rightarrow+0, we set sinττ\sin\tau\simeq\tau and cosτ1τ2/2\cos\tau\simeq 1-\tau^{2}/2. Moreover, we let r1r\gg 1. Then, we obtain

g(r,τ)2exp(2r2τ2)cos(4r2τ)exp(8r2τ2)cos(8r2τ).g(r,\tau)\simeq 2\exp(-2r^{2}\tau^{2})\cos(4r^{2}\tau)-\exp(-8r^{2}\tau^{2})\cos(8r^{2}\tau). (57)

Because g(r,τ)g(r,\tau) given by Eq. (57) is very unmanageable, it is difficult to compute its maximum value under τ+0\tau\rightarrow+0 and rr\rightarrow\infty. Thus, we utilize the following special technique for obtaining the limit of g(r,τ)g(r,\tau).

Here, we put x=r2τx=r^{2}\tau and assume that xx takes a finite value. Then, we can rewrite Eq. (57) as follows:

g(τ,x)=2exp(2xτ)cos4xexp(8xτ)cos8x.g(\tau,x)=2\exp(-2x\tau)\cos 4x-\exp(-8x\tau)\cos 8x. (58)

For the above equation, first we fix xx at a finite value and second we let τ\tau approach zero, that is τ+0\tau\to+0. As a result of these operations, we obtain the following function as the limit of g(τ,x)g(\tau,x):

g(x)=2cos4xcos8x.g(x)=2\cos 4x-\cos 8x. (59)

The smallest positive value of xx that maximizes g(x)g(x) is given by x=π/12x=\pi/12 and the maximum of g(x)g(x) is equal to 3/23/2. Thus, we can suppose that the maximum value of K3(0;πτ,r,τ)K_{3}(0;\pi-\tau,r,\tau) converges to 3/23/2 under τ+0\tau\to+0 and rr\to\infty.

Here, we are confronted with a problem whether or not r2τ=π/12r^{2}\tau=\pi/12 actually holds for rr and τ\tau that maximize K3(0;πτ,r,τ)K_{3}(0;\pi-\tau,r,\tau). We have confirmed that r2τπ/12r^{2}\tau\simeq\pi/12 holds around τ0\tau\simeq 0 for Γ=0\Gamma=0 by numerical calculations. To be more precise, we have obtained that r2τ=Const.(π/12)r^{2}\tau=\mbox{Const.}(\pi/12) with 0.9068Const.0.96830.9068\leq\mbox{Const.}\leq 0.9683 for τ=0.001×n\tau=0.001\times n and n=1,2,,8n=1,2,...,8. Moreover, we have verified that Const. increases and gets closer to unity as τ\tau approaches zero. Thus, this evidence suggests that the maximum K3K_{3} converges to 3/23/2 for τ+0\tau\to+0.

In Refs. [4, 9], the following results were shown. We consider a two-level atom which does not interact with an environment and has independent time evolution. We assume that its Hamiltonian is given by H=(ω0/2)σzH=(\omega_{0}/2)\sigma_{z}. Then, we obtain the time-symmetrized correlation function,

C(τ)\displaystyle C(\tau) =\displaystyle= {σx(t),σx(t+τ)}/2\displaystyle\langle\{\sigma_{x}(t),\sigma_{x}(t+\tau)\}\rangle/2 (60)
=\displaystyle= cos(ω0τ).\displaystyle\cos(\omega_{0}\tau).

Thus, K3K_{3} of the LGI of equally spaced measurements with separation τ\tau is given by

K3\displaystyle K_{3} =\displaystyle= 2C(τ)C(2τ)\displaystyle 2C(\tau)-C(2\tau) (61)
=\displaystyle= 2cos(ω0τ)cos(2ω0τ).\displaystyle 2\cos(\omega_{0}\tau)-\cos(2\omega_{0}\tau).

This equation is similar to Eq. (59) and they correspond to each other with putting ω0τ=4x\omega_{0}\tau=4x. The reason why we can find such a resemblance in Eqs. (59) and (61) is as follows.

If we fix α\alpha at a finite complex value and let β\beta diverge to infinity, β=reiθ\beta=re^{i\theta} and rr\rightarrow\infty, we obtain

α|2β\displaystyle\langle\alpha|2\beta\rangle =\displaystyle= exp[12(|α|2+4|β|2)+2αβ]\displaystyle\exp[-\frac{1}{2}(|\alpha|^{2}+4|\beta|^{2})+2\alpha^{*}\beta] (62)
\displaystyle\rightarrow 0,\displaystyle 0,

so that we can consider that |α|\alpha\rangle and |2β|2\beta\rangle are approximately orthogonal to each other. Because of Eqs. (18) and (22), putting |β||\beta|\rightarrow\infty, we obtain

Π(±)(β)|α\displaystyle\Pi^{(\pm)}(\beta)|\alpha\rangle =\displaystyle= 12[|α±exp(βαβα)|2βα]\displaystyle\frac{1}{2}[|\alpha\rangle\pm\exp(\beta^{*}\alpha-\beta\alpha^{*})|2\beta-\alpha\rangle] (63)
\displaystyle\rightarrow 12(|α±eiφ|2β),\displaystyle\frac{1}{2}(|\alpha\rangle\pm e^{i\varphi}|2\beta\rangle),

where α=seiμ\alpha=se^{i\mu} and φ=2irssin(μθ)\varphi=2irs\sin(\mu-\theta).

These facts imply that Π(±)(β)\Pi^{(\pm)}(\beta) are equivalent to a projection measurement with σx\sigma_{x} if we regard the system of interest as a two-level system {|α,eiφ|2β}\{|\alpha\rangle,e^{i\varphi}|2\beta\rangle\}, that is |0|α|0\rangle\equiv|\alpha\rangle and |1eiφ|2β|1\rangle\equiv e^{i\varphi}|2\beta\rangle. Defining the Hamiltonian of the system as H=ωaaH=\omega a^{\dagger}a with ω=1\omega=1, we evaluate energies of the two states |α|\alpha\rangle and eiφ|2βe^{i\varphi}|2\beta\rangle roughly,

α|H|α=|α|2,\langle\alpha|H|\alpha\rangle=|\alpha|^{2}, (64)
2β|eiφHeiφ|2β=4r2.\langle 2\beta|e^{-i\varphi}He^{i\varphi}|2\beta\rangle=4r^{2}. (65)

Because of |α|24r2|\alpha|^{2}\ll 4r^{2}, we can rewrite the Hamiltonian as

H=12(4r2)σz.H=\frac{1}{2}(4r^{2})\sigma_{z}. (66)

Here, setting ω0=4r2\omega_{0}=4r^{2}, we can obtain K3K_{3} of the LGI for this system as

K3=2cos(4r2τ)cos(8r2τ).K_{3}=2\cos(4r^{2}\tau)-\cos(8r^{2}\tau). (67)

If we substitute r2τ=xr^{2}\tau=x into Eq. (67), we reach Eq. (59).

During the above discussion, first we take the limit τ+0\tau\to+0, and second we let rr approach infinity as rr\to\infty. Here, we consider these two processes in reverse. First of all, we choose Π(±)(β)\Pi^{(\pm)}(\beta) for the projection measurement of the system. Although Π(±)(β)|α\Pi^{(\pm)}(\beta)|\alpha\rangle are orthogonal to each other, they are not normalized.

To maximize K3K_{3}, we had better make Π(±)(β)|α\Pi^{(\pm)}(\beta)|\alpha\rangle be normalized and orthogonal to each other. Thus, because of Eqs. (62) and (63), we need to let |β||\beta| diverge to infinity as |β||\beta|\rightarrow\infty. Then, we can regard the system of interest as a two-level system and its K3K_{3} is given by Eq. (67). In Eq. (67), in order to maximize K3K_{3}, we have to let a relation r2τ=(π/12)+(n/2)πr^{2}\tau=(\pi/12)+(n/2)\pi holds for n=0,1,2,n=0,1,2,.... Here, we choose the smallest value as r2τ=π/12r^{2}\tau=\pi/12 for determining it uniquely. Then, due to r=|β|r=|\beta|\rightarrow\infty, we come to a conclusion τ+0\tau\rightarrow+0. Therefore, we have to employ the two processes of approaching the limits, rr\rightarrow\infty and τ+0\tau\rightarrow+0, and the singularity emerges.

In the above arguments, the limit rr\to\infty seems to cause the limit τ+0\tau\to+0. However, from a viewpoint of a practical procedure of physics, we have to first decide the time τ\tau of the measurement and second determine the parameter rr for the projection operators, so that we cannot help feeling that the order of taking limits are not normal. Here, we calm ourselves and follow the discussion in a different manner. If we take the limit rr\to\infty, the optimum τ\tau that maximizes K3K_{3} is uniquely determined as τ+0\tau\to+0 because of r2τ=π/12r^{2}\tau=\pi/12. Hence, thinking about the optimization problem, we have to admit that the two operations, taking the limits as rr\to\infty and τ+0\tau\to+0, are commutes with each other.

6 Numerical analyses of the LGI and its optimization for the system initially prepared in the cat state (|α+|α)(|\alpha\rangle+|-\alpha\rangle)

In the present section, letting the initial state be given by the cat state, we examine K3K_{3} of the LGI numerically. Furthermore, we consider the optimization problem of the displaced parity operators. In a similar fashion to Sect. 5, to make discussion be simple, we put α=1/2\alpha=1/2 and ω=1\omega=1. Moreover, we use the notation K3(Γ;θ,r,τ)K_{3}(\Gamma;\theta,r,\tau).

Refer to caption
Figure 8: Graphs of K3(Γ;0,1/2,τ)K_{3}(\Gamma;0,1/2,\tau) as a function of τ\tau with letting the system be given by the cat state initially and putting β=1/2\beta=1/2, that is θ=0\theta=0 and r=1/2r=1/2. The thick solid red, thin solid blue, and thin dashed purple curves represent plots of Γ=0\Gamma=0, 0.050.05, and 0.20.2, respectively. The thick solid red plot is periodic about τ\tau and its period is equal to 2π2\pi. As Γ\Gamma becomes larger, amplitudes of the curves decrease.

In Fig. 8, we draw graphs of K3(Γ;0,1/2,τ)K_{3}(\Gamma;0,1/2,\tau) as a function of τ\tau with β=1/2\beta=1/2, that is θ=0\theta=0 and r=1/2r=1/2. The thick solid red, thin solid blue, and thin dashed purple curves represent plots of Γ=0\Gamma=0, 0.050.05, and 0.20.2, respectively. The thick solid red graph is periodic about τ\tau and its period is given by 2π2\pi. As Γ\Gamma increases, amplitudes of the curves shrink.

We can numerically verify that the curve of Γ=0.2\Gamma=0.2 converges to 0.376 7420.376{\,}742... as τ\tau becomes larger. This fact can be confirmed in the manner of mathematical analysis. Fixing Γ\Gamma at a finite value Γ0(>)\Gamma_{0}(>) and taking the limit τ\tau\to\infty, we obtain

limτK3(Γ0;0,1/2,τ)\displaystyle\lim_{\tau\to\infty}K_{3}(\Gamma_{0};0,1/2,\tau) =\displaystyle= 1+2e+5e4e3/2(1+e)\displaystyle\frac{1+2\sqrt{e}+5e}{4e^{3/2}(1+\sqrt{e})} (68)
\displaystyle\simeq 0.376 472.\displaystyle 0.376{\,}472....

From Eq. (68), we understand that K3K_{3} converges to the above value for Γ0>0\Gamma_{0}>0 and τ\tau\to\infty.

In general, α=seiμ\forall\alpha=se^{i\mu}, ω\forall\omega, Γ(>0)\forall\Gamma(>0), and β=reiθ\forall\beta=re^{i\theta}, we obtain

limτK3\displaystyle\lim_{\tau\rightarrow\infty}K_{3} =\displaystyle= exp{2r[2r+scos(θμ)]}4[1+exp(2s2)]\displaystyle\frac{\exp\{-2r[2r+s\cos(\theta-\mu)]\}}{4[1+\exp(2s^{2})]} (69)
×{exp[2rscos(θμ)][3+exp(2r2)+4exp(2s2)]\displaystyle\times\Biggl{\{}\exp[2rs\cos(\theta-\mu)][3+\exp(2r^{2})+4\exp(2s^{2})]
+[1exp(2r2)]cos[2rssin(θμ)]}.\displaystyle+[1-\exp(2r^{2})]\cos[2rs\sin(\theta-\mu)]\Biggr{\}}.
Refer to caption
Figure 9: Graphs of the optimum θ\theta that maximizes K3K_{3} for each τ(>0)\tau(>0) as a function of τ\tau with letting the system be given by the cat state initially. The thick solid red, thin solid blue, and thin dashed purple curves represent plots of Γ=0\Gamma=0, 0.10.1, and 11, respectively. There are discontinuity points in all the three curves.
Refer to caption
Figure 10: A thick red curve represents the optimum θ\theta that maximizes K3K_{3} for each τ(>0)\tau(>0) with Γ=0\Gamma=0 as a function of τ\tau, where the system is initialized in the cat state. Four parallel thin blue lines represent θ=(5π/2)τ\theta=(5\pi/2)-\tau, θ=2πτ\theta=2\pi-\tau, θ=πτ\theta=\pi-\tau, and θ=(π/2)τ\theta=(\pi/2)-\tau. The optimum θ\theta moves and jumps on the lines, θ=(π/2)τ\theta=(\pi/2)-\tau, θ=πτ\theta=\pi-\tau, θ=π\theta=-\pi, θ=2πτ\theta=2\pi-\tau, and θ=(5π/2)τ\theta=(5\pi/2)-\tau, in order as τ\tau increases. As explained in Sect. 4, K3K_{3} is a periodic function about θ\theta and its period is given by π\pi, so that θ=2πτ\theta=2\pi-\tau and θ=πτ\theta=\pi-\tau are essentially equivalent to each other, and so do θ=(5π/2)τ\theta=(5\pi/2)-\tau and θ=(π/2)τ\theta=(\pi/2)-\tau.
Refer to caption
Figure 11: Plots of the optimum rr that maximizes K3K_{3} for each τ(>0)\tau(>0) as a function of τ\tau with preparing the system initially in the cat state. The thick solid red, thin solid blue, and thin dashed purple curves represent graphs of Γ=0\Gamma=0, 0.10.1, and 11, respectively. All the three curves apparently diverge to infinity at τ=0\tau=0. Moreover, the curve of Γ=0\Gamma=0 seemingly diverges to infinity at τ=2π\tau=2\pi.
Refer to caption
Figure 12: Graphs of the maximized K3K_{3}, with adjusting θ\theta and rr, for each τ(>0)\tau(>0) as a function of τ\tau, where the system is initialized in the cat state. The thick solid red, thin solid blue, and thin dashed purple curves represent plots of Γ=0\Gamma=0, 0.10.1, and 11, respectively. As Γ\Gamma increases, amplitudes of the curves diminish.

Next, in a similar way to Sect. 5, we consider the following optimization problem. We look for the optimum values of θ\theta and rr that maximize K3K_{3} for given arbitrary τ(>0)\tau(>0). In Figs. 9 and 11, we plot the optimum θ\theta and rr that maximize K3K_{3} for each τ(>0)\tau(>0) as functions of τ\tau. In Fig. 12, we plot the maximized K3K_{3} with adjusting θ\theta and rr for each τ(>0)\tau(>0) as a function of τ\tau. In Figs. 9, 11, and 12, the thick solid red, thin solid blue, and thin dashed purple curves represent plots of Γ=0\Gamma=0, 0.10.1, and 11, respectively. In Fig. 10, a thick red curve represents the maximized K3K_{3} for each τ(>0)\tau(>0) and Γ=0\Gamma=0 with adjustments of θ\theta and rr as a function of τ\tau, and thin blue lines consist of θ=(5π/2)τ\theta=(5\pi/2)-\tau, θ=2πτ\theta=2\pi-\tau, θ=πτ\theta=\pi-\tau, and θ=(π/2)τ\theta=(\pi/2)-\tau.

Here, we focus on Fig. 10. The optimized θ\theta that makes K3K_{3} be maximum moves and jumps on the four parallel blue lines. In the following, we examine this fact in detail.

From slightly tough calculations, we obtain

θK3(0;θ,r,τ)|θ=(π/2)τ\displaystyle\left.\frac{\partial}{\partial\theta}K_{3}(0;\theta,r,\tau)\right|_{\theta=(\pi/2)-\tau} =\displaystyle= θK3(0;θ,r,τ)|θ=πτ\displaystyle\left.\frac{\partial}{\partial\theta}K_{3}(0;\theta,r,\tau)\right|_{\theta=\pi-\tau} (70)
=\displaystyle= 0.\displaystyle 0.

Because K3(0;θ,r,τ)K_{3}(0;\theta,r,\tau) is a periodic function about θ\theta and its period is given by π\pi,
(/θ)K3(0;θ,r,τ)|θ=2πτ=(/θ)K3(0;θ,r,τ)|θ=(5π/2)τ=0\left.(\partial/\partial\theta)K_{3}(0;\theta,r,\tau)\right|_{\theta=2\pi-\tau}=\left.(\partial/\partial\theta)K_{3}(0;\theta,r,\tau)\right|_{\theta=(5\pi/2)-\tau}=0 holds obviously. Thus, in the cases of θ=(π/2)τ\theta=(\pi/2)-\tau, θ=πτ\theta=\pi-\tau, θ=2πτ\theta=2\pi-\tau, and θ=(5π/2)τ\theta=(5\pi/2)-\tau, we can expect that K3(0;θ,r,τ)K_{3}(0;\theta,r,\tau) takes extreme values.

In Fig. 10, θ=0\theta=0 holds for 2.050τ4.2502.050\leq\tau\leq 4.250. Looking at the curve of Γ=0\Gamma=0 in Fig. 11, we can confirm that r=0r=0 holds for this range of τ\tau. Thus, we can regard θ\theta as meaningless for 2.050τ4.2502.050\leq\tau\leq 4.250.

In a similar fashion to the case where the initial state is given by a coherent state |α|\alpha\rangle, letting the initial state be the cat state, we can find strong singularity of K3K_{3} around τ=0\tau=0. Due to the definition of K3K_{3} in Eq. (11), we suppose that K3(0;θ,r,0)=1K_{3}(0;\theta,r,0)=1 seemingly holds. However, the maximum value of K3K_{3} approaches 1.51.5 under the limit τ+0\tau\rightarrow+0 in Fig. 12. From numerical calculations, we can confirm that the maximum value of K3K_{3} attains 1.499 021.499{\,}02... at τ=0.001\tau=0.001. We investigate these facts in detail in the following paragraphs.

Looking at Fig. 10, we suppose that the maximized K3K_{3} is given by θ=(π/2)τ\theta=(\pi/2)-\tau around τ=0\tau=0. Thus, we consider a function:

g(r,τ)=K3(0;(π/2)τ,r,τ).g(r,\tau)=K_{3}(0;(\pi/2)-\tau,r,\tau). (71)

We apply the following approximations to g(r,τ)g(r,\tau) in the above equation. Taking the limit τ+0\tau\rightarrow+0, we put sinττ\sin\tau\simeq\tau and cosτ1τ2/2\cos\tau\simeq 1-\tau^{2}/2. Moreover, we set r1r\gg 1. Then, we obtain

g(r,τ)\displaystyle g(r,\tau) \displaystyle\simeq exp(2rτ)1+e{[1+exp(4rτ)+2exp(12+2rτ)]cos(4r2τ)\displaystyle\frac{\exp(-2r\tau)}{1+\sqrt{e}}\Biggl{\{}[1+\exp(4r\tau)+2\exp(\frac{1}{2}+2r\tau)]\cos(4r^{2}\tau) (72)
[1+exp(12+16r2τ2)]cos(8r2τ)}.\displaystyle-[1+\exp(\frac{1}{2}+16r^{2}\tau^{2})]\cos(8r^{2}\tau)\Biggr{\}}.

Being similar to that in Eq. (57), the above g(r,τ)g(r,\tau) is very intractable, so that we cannot compute g(r,τ)g(r,\tau) for rr\to\infty and τ+0\tau\to+0 with ease. Thus, we use the same technique for estimation of the limit as discussed in Sect. 5.

First, we put x=r2τx=r^{2}\tau and we assume that xx is fixed at a finite value. Second, putting xx at the finite value, we take the limits, τ+0\tau\to+0 and rr\to\infty. Then, because of rτ=x/r0r\tau=x/r\to 0 and r2τ2=xτ0r^{2}\tau^{2}=x\tau\to 0, we obtain the following function for the limit of g(r,τ)g(r,\tau):

g(x)=2cos4xcos8x.g(x)=2\cos 4x-\cos 8x. (73)

This equation has appeared in Sect. 5 already and g(x)g(x) attains the maximum value 3/23/2 at x=π/12x=\pi/12. Thus, we can suppose that the maximum value of K3(0;(π/2)τ,r,τ)K_{3}(0;(\pi/2)-\tau,r,\tau) converges to 3/23/2 in the limits, rr\to\infty and τ+0\tau\to+0.

We have verified that r2τπ/12r^{2}\tau\simeq\pi/12 holds around τ0\tau\simeq 0 for Γ=0\Gamma=0 numerically. To be more precise, we have obtained that r2τ=Const.(π/12)r^{2}\tau=\mbox{Const.}(\pi/12) with 0.9934Const.0.99910.9934\leq\mbox{Const.}\leq 0.9991 for τ=0.001×n\tau=0.001\times n and n=1,2,,8n=1,2,...,8. Furthermore, we have made sure that Const. increases and approaches unity as τ\tau gets closer to zero. Therefore, we can guess that the maximum value of K3K_{3} converges to 3/23/2 in the limit τ+0\tau\to+0.

7 Comparisons of the LGIs for cases where the systems are initially prepared in the coherent and cat states

Refer to caption
Figure 13: Graphs of the maximized K3K_{3} as a function of τ\tau. The thick and thin curves represent plots of K3K_{3} for the systems initially set in the coherent state |α|\alpha\rangle and the cat state (|α+|α)(|\alpha\rangle+|-\alpha\rangle) respectively, where α=1/2\alpha=1/2 and ω=1\omega=1. The solid red, long dashed blue, and short dashed purple curves represent plots of Γ=0\Gamma=0, 0.10.1, 11, respectively. We notice that the coherent state shows a larger violation of the LGI than the cat state for τ\tau in a specific range.
Refer to caption
Figure 14: Graphs of the maximized K3K_{3} as a function of τ\tau. The thick and thin curves represent plots of K3K_{3} for the systems prepared initially in the coherent state |α|\alpha\rangle and the cat state (|α+|α)(|\alpha\rangle+|-\alpha\rangle) respectively, where α=1\alpha=1 and ω=1\omega=1. The solid red, long dashed blue, and short dashed purple curves represent plots of Γ=0\Gamma=0, 0.10.1, and 11, respectively. We become aware that the coherent state reveals a larger violation of the LGI than the cat state for τ\tau in a specific range.

In Fig. 13, we plot the maximized K3K_{3} as a function of τ\tau with α=1/2\alpha=1/2 and ω=1\omega=1. The thick and thin curves represent plots for the systems initialized in the coherent and cat states, respectively. The solid red, long dashed blue, and short dashed purple curves represent graphs of Γ=0\Gamma=0, 0.10.1, and 11, respectively. In Fig. 14, the same graphs are drawn as Fig. 13 but α=1\alpha=1.

We focus on the curves of Γ=0\Gamma=0 in Fig. 13. In the ranges of 0.175τ2.1250.175\leq\tau\leq 2.125 and 4.175τ6.1254.175\leq\tau\leq 6.125, the value of the maximized K3K_{3} for the coherent state is larger than that for the cat state. By contrast, in the ranges of 0τ<0.1750\leq\tau<0.175 and 6.125<τ2π6.125<\tau\leq 2\pi, the value of the maximized K3K_{3} for the cat state is larger than that for the coherent state. Except for the above ranges, the values of the maximized K3K_{3} are equal to each other for both the states.

In Fig. 14 for α=1\alpha=1, we can observe the same facts. We concentrate on the curves of Γ=0\Gamma=0. In the ranges of 0.1τ2.550.1\leq\tau\leq 2.55 and 3.725τ6.1753.725\leq\tau\leq 6.175, the values of the maximized K3K_{3} for the coherent state is larger than that for the cat state. Contrastingly, in the ranges of 0τ<0.10\leq\tau<0.1 and 6.175<τ2π6.175<\tau\leq 2\pi, the value of the maximized K3K_{3} for the cat state is larger than that for the coherent state. Except for the above ranges, the values of the maximized K3K_{3} for both the states are equal to each other.

In Figs. 13 and 14, for both Γ=0.1\Gamma=0.1 and Γ=1\Gamma=1, the value of the maximized K3K_{3} for coherent state is larger than that for the cat state in a wide range of τ\tau.

From the above results, we can conclude that the coherent state exhibits a characteristic of quantum nature more strongly than the cat state in the specific ranges of τ\tau.

8 Discussion

In the current paper, we demonstrate that the coherent state shows a characteristic of the quantum nature more intensely than the cat state for the specific ranges of the time difference concerning to the violation of the LGI. In Ref. [6], it has been already pointed out that the coherent state is able to violate the LGI. In Ref. [6], Chevalier et al. constructed the LGI by using the Mach-Zehnder interferometer for the measurements. Chevalier et al. argued how to evaluate the LGI by injecting the coherent light into the Mach-Zehnder interferometer and applying the negative measurement, that was so-called interaction-free measurement, to it. In contrast, Ref. [11] showed that the cat state was able to violate the LGI.

In general, the coherent state has the balanced minimum uncertainty ΔX=ΔP=1/2\Delta X=\Delta P=1/\sqrt{2} and it is regarded as one of the most classical-like states. By contrast, Refs. [18, 19] mentioned that the cat state was able to exhibit sub-Poisson photon statistics. From these viewpoints, we can regard the coherent and cat states as pseudoclassical and nonclassical, respectively. However, as shown in Sect. 7, the violation of the Leggett-Garg inequality of the coherent state is stronger than that of the cat state for time differences of τ\tau in some ranges. This fact suggests to us that the violation of the Leggett-Garg inequality can be a witness of non-classicality of wave functions but does not work as a quantitative measure of it.

It is certain that the violation disproves the macroscopic realism of the physical system. However, we cannot define the concept of the macroscopicity explicitly. In actual fact, Ref. [31] showed that a particular time evolution with coarse-grained measurements caused the macrorealism in a quantum system. In Ref. [32], Moreira et al. considered the macroscopic realism to be a model dependent notion and provided a toy model in which the invasiveness was controlled by physical parameters. Because of these circumstances, we have not obtained quantitative measure of the macrorealism yet.

It is possible that a choice of observables affects the degree of non-macroscopicity revealed by the violation of the LGI. In the present paper, we choose the displaced parity operators for the measurements of the boson system in the LGI. We can suppose that this choice lets the coherent state exhibit a characteristic of the quantum nature more strongly than the cat state.

In Ref. [33], it was reported that two- and four-qubit cat states violated the LGI but a six-qubit cat state did not violate the bound of clumsy-macrorealistic. (The experimental solution of the clumsy loophole, in other words clumsy measurement process inducing invasive one and causing a violation rather than quantum effect, was addressed by Refs. [34, 35].) In the current paper, we cannot determine what kind of quantity of the macroscopic realism the displaced parity operators reflect. We may be able to let the cat state exhibit a characteristic of quantum nature more strongly than the coherent state for the violation of the LGI by opting the other operators for the measurements, rather than the displaced parity operators.

In order to realize the tests the current paper considers in a laboratory, we have to perform measurements of quantum states using the displaced parity operators without destroying them. In other words, quantum nondemolition measurements with the displaced parity operators are essential. Reference [36] has reviewed measurements with the parity operators comprehensively. In Ref. [37], an experimental proposition to detect the parity of the field of one specified mode in a high-QQ resonator is explained. Reference [38] reported experiments for photon-number parity measurements of coherent states using a photon-number resolving detector and a polarization version of the Mach-Zehnder interferometer. However, in these experiments, quantum states were destroyed after the observations and the quantum nondemolition measurements were not executed. The quantum nondemolition measurement with the displaced parity operators is one of the most difficult challenges in the field of experimental quantum optics.

Recently, as another path to demonstrate a quantum nondemolition measurement on a bosonic mode, opto-electro-mechanical and nanomechanical systems have been examined theoretically. In Ref. [39], Lambert et al. showed that an unambiguous violation of the LGI was given using an opto-electo-mechanical system with an additional circuit-QED measurement device. In Ref. [40], Johansson et al. considered how to generate entangled states with a multimode nanomechanical resonator and observe violations of the Bell inequality. Although these set-ups do not utilize the displaced parity operators, they give us some suggestions for a realization of the quantum nondemolition measurements of the LGI.

In Figs. 13 and 14, we become aware that K31K_{3}\neq 1 holds in the limit τ+0\tau\to+0 for Γ=0\Gamma=0. This phenomenon can be observed for a positive decay rate, that is Γ>0\Gamma>0, as well. It is a novel discovery that the displaced parity operators reveal the strong singularity in the LGI in the limit τ+0\tau\rightarrow+0. Referring to Refs. [4, 5], in general, K3=1K_{3}=1 holds at τ=0\tau=0 for a two-level system using projection operators as observables. We can attribute the singularity found in the current paper to the fact that the dimension of the Hilbert space of the system is infinite.

In the current paper, we examine the LGI for a single boson mode that lies on an infinite dimensional Hilbert space. In contrast, there are some works concerning the LGI for a multi-level system and an ensemble of qubits. In Ref. [41], Budroni and Emary showed that K3>3/2K_{3}>3/2 was able to hold for an NN-level system such as a large spin with projection operators measuring the spin in the zz direction. It is proved in general that the maximum value of K3K_{3} is equal to 3/23/2 for a two-level system [2], so that their results are interesting. In Ref. [42], Lambert et al. investigated the violation of the LGI for a large ensemble of qubits and showed the following. When a parity of the projection of the spin in the zz direction was chosen as the dichotomic variable for measurement like Ref. [41], the violation of the LGI occurred at τ\tau that approached zero (τ+0\tau\to+0) as the number of qubits NN became larger. This observation matches our result of Figs. 13 and 14 that the maximized K3K_{3} appears at τ=0\tau=0. However, the maximized K3K_{3} shrank to unity as NN became larger in Ref. [42] although K3K_{3} of Figs. 13 and 14 attain 3/23/2 at τ=0\tau=0. We can suppose that our optimization of measurement operators causes this difference.

In the current paper, under the optimization of the displaced parity operators, we obtain the restriction K33/2K_{3}\leq 3/2. In Ref. [43], it was proven thoroughly that K3K_{3} must be equal to or less than 3/23/2 if the observables Π+\Pi_{+} and Π\Pi_{-} are projection operators onto eigenspaces of Q=±1Q=\pm 1. Hence, in the case of the current paper, the relation K33/2K_{3}\leq 3/2 is valid. Contrastingly, Ref. [41] showed examples which demonstrated K3>3/2K_{3}>3/2.

In the present paper, we study the boson system coupled to the zero-temperature environment. The simplest method for analysing time evolution of a boson system that interacts with a thermal reservoir is solving the master equation with the perturbative approach, for instance, low temperature expansion. However, calculations of this method tend to be complicated, so that it is not practical. In Ref. [9], Friedenberger and Lutz derived time evolution of a qubit coupled to the thermal reservoir by using the quantum regression theorem. Because this approach gives us a clear perspective for solving the time evolution of the qubit, we may apply it to a problem of the thermal boson system, as well.

Appendix A The mathematical forms used in Sect. 3

An explicit form of w1±(τ)w_{1\pm}(\tau), the time evolution of w1±(0)w_{1\pm}(0) given by Eqs. (17) and (21), is written down as follows:

w1±(τ)\displaystyle w_{1\pm}(\tau) =\displaystyle= 14{|αeiΩταeiΩτ|\displaystyle\frac{1}{4}\Biggl{\{}|\alpha e^{-i\Omega\tau}\rangle\langle\alpha e^{-i\Omega\tau}| (74)
±exp{βαβα(1/2)[|2βα|2+|α|22(2βα)α](1e2Γτ)}\displaystyle\pm\exp\{\beta^{*}\alpha-\beta\alpha^{*}-(1/2)[|2\beta-\alpha|^{2}+|\alpha|^{2}-2(2\beta-\alpha)\alpha^{*}](1-e^{-2\Gamma\tau})\}
×|(2βα)eiΩταeiΩτ|\displaystyle\times|(2\beta-\alpha)e^{-i\Omega\tau}\rangle\langle\alpha e^{-i\Omega\tau}|
±(the hermitian conjugate of the above term)\displaystyle\pm(\mbox{the hermitian conjugate of the above term})
+|(2βα)eiΩτ(2βα)eiΩτ|}.\displaystyle+|(2\beta-\alpha)e^{-i\Omega\tau}\rangle\langle(2\beta-\alpha)e^{-i\Omega\tau}|\Biggr{\}}.

Explicit forms of p1±,2+p_{1\pm,2+} and p1±,2p_{1\pm,2-}, the probabilities that O2=1O_{2}=1 and O2=1O_{2}=-1 are observed with the measurement on the above w1±(τ)w_{1\pm}(\tau) at time t2t_{2} respectively, are given by

p1±,2+\displaystyle p_{1\pm,2+} =\displaystyle= Tr[Π(+)(β)w1±(τ)]\displaystyle\mbox{Tr}[\Pi^{(+)}(\beta)w_{1\pm}(\tau)] (75)
=\displaystyle= 14{exp(|αeiΩτβ|2)cosh|αeiΩτβ|2\displaystyle\frac{1}{4}\Biggl{\{}\exp(-|\alpha e^{-i\Omega\tau}-\beta|^{2})\cosh|\alpha e^{-i\Omega\tau}-\beta|^{2}
±2Re[exp{βαβα(1/2)[|2βα|2+|α|22(2βα)α](1e2Γτ)\displaystyle\pm 2\mbox{Re}\Biggr{[}\exp\{\beta^{*}\alpha-\beta\alpha^{*}-(1/2)[|2\beta-\alpha|^{2}+|\alpha|^{2}-2(2\beta-\alpha)\alpha^{*}](1-e^{-2\Gamma\tau})
β(βα)exp(iΩτ)+β(βα)exp(iΩτ)\displaystyle-\beta(\beta^{*}-\alpha^{*})\exp(i\Omega^{*}\tau)+\beta^{*}(\beta-\alpha)\exp(-i\Omega\tau)
(1/2)[|αeiΩτβ|2+|(2βα)eiΩτβ|2]}\displaystyle-(1/2)[|\alpha e^{-i\Omega\tau}-\beta|^{2}+|(2\beta-\alpha)e^{-i\Omega\tau}-\beta|^{2}]\}
×cosh[(αexp(iΩτ)β)((2βα)exp(iΩτ)β)]]\displaystyle\times\cosh[(\alpha^{*}\exp(i\Omega^{*}\tau)-\beta^{*})((2\beta-\alpha)\exp(-i\Omega\tau)-\beta)]\Biggl{]}
+exp[|(2βα)eiΩτβ|2]cosh|(2βα)eiΩτβ|2},\displaystyle+\exp[-|(2\beta-\alpha)e^{-i\Omega\tau}-\beta|^{2}]\cosh|(2\beta-\alpha)e^{-i\Omega\tau}-\beta|^{2}\Biggr{\}},
p1±,2\displaystyle p_{1\pm,2-} =\displaystyle= Tr[Π()(β)w1±(τ)]\displaystyle\mbox{Tr}[\Pi^{(-)}(\beta)w_{1\pm}(\tau)] (76)
=\displaystyle= 14{exp(|αeiΩτβ|2)sinh|αeiΩτβ|2\displaystyle\frac{1}{4}\Biggl{\{}\exp(-|\alpha e^{-i\Omega\tau}-\beta|^{2})\sinh|\alpha e^{-i\Omega\tau}-\beta|^{2}
±2Re[exp{βαβα(1/2)[|2βα|2+|α|22(2βα)α](1e2Γτ)\displaystyle\pm 2\mbox{Re}\Biggr{[}\exp\{\beta^{*}\alpha-\beta\alpha^{*}-(1/2)[|2\beta-\alpha|^{2}+|\alpha|^{2}-2(2\beta-\alpha)\alpha^{*}](1-e^{-2\Gamma\tau})
β(βα)exp(iΩτ)+β(βα)exp(iΩτ)\displaystyle-\beta(\beta^{*}-\alpha^{*})\exp(i\Omega^{*}\tau)+\beta^{*}(\beta-\alpha)\exp(-i\Omega\tau)
(1/2)[|αeiΩτβ|2+|(2βα)eiΩτβ|2]}\displaystyle-(1/2)[|\alpha e^{-i\Omega\tau}-\beta|^{2}+|(2\beta-\alpha)e^{-i\Omega\tau}-\beta|^{2}]\}
×sinh[(αexp(iΩτ)β)((2βα)exp(iΩτ)β)]]\displaystyle\times\sinh[(\alpha^{*}\exp(i\Omega^{*}\tau)-\beta^{*})((2\beta-\alpha)\exp(-i\Omega\tau)-\beta)]\Biggl{]}
+exp[|(2βα)eiΩτβ|2]sinh|(2βα)eiΩτβ|2}.\displaystyle+\exp[-|(2\beta-\alpha)e^{-i\Omega\tau}-\beta|^{2}]\sinh|(2\beta-\alpha)e^{-i\Omega\tau}-\beta|^{2}\Biggr{\}}.

Appendix B The mathematical forms used in Sect. 4

Explicit forms of {K(j)(0):j=1,2,4}\{K^{(j)}(0):j=1,2,4\}, {L(j)(0):j=1,2,3,4}\{L^{(j)}(0):j=1,2,3,4\}, and {M(j)(0):j=1,2,4}\{M^{(j)}(0):j=1,2,4\} appearing in Eq. (35) are given by

K(1)(0)\displaystyle K^{(1)}(0) =\displaystyle= |αα|,\displaystyle|\alpha\rangle\langle\alpha|,
K(2)(0)\displaystyle K^{(2)}(0) =\displaystyle= exp(βαβα)|αα+2β|,\displaystyle\exp(\beta\alpha^{*}-\beta^{*}\alpha)|\alpha\rangle\langle-\alpha+2\beta|,
K(4)(0)\displaystyle K^{(4)}(0) =\displaystyle= |α+2βα+2β|,\displaystyle|-\alpha+2\beta\rangle\langle-\alpha+2\beta|, (77)
L(1)(0)\displaystyle L^{(1)}(0) =\displaystyle= |αα|,\displaystyle|\alpha\rangle\langle-\alpha|,
L(2)(0)\displaystyle L^{(2)}(0) =\displaystyle= exp(βα+βα)|αα+2β|,\displaystyle\exp(-\beta\alpha^{*}+\beta^{*}\alpha)|\alpha\rangle\langle\alpha+2\beta|,
L(3)(0)\displaystyle L^{(3)}(0) =\displaystyle= exp(βα+βα)|α+2βα|,\displaystyle\exp(-\beta\alpha^{*}+\beta^{*}\alpha)|-\alpha+2\beta\rangle\langle-\alpha|,
L(4)(0)\displaystyle L^{(4)}(0) =\displaystyle= exp[2(βα+βα)]|α+2βα+2β|,\displaystyle\exp[2(-\beta\alpha^{*}+\beta^{*}\alpha)]|-\alpha+2\beta\rangle\langle\alpha+2\beta|, (78)
M(1)(0)\displaystyle M^{(1)}(0) =\displaystyle= |αα|,\displaystyle|-\alpha\rangle\langle-\alpha|,
M(2)(0)\displaystyle M^{(2)}(0) =\displaystyle= exp(βα+βα)|αα+2β|,\displaystyle\exp(-\beta\alpha^{*}+\beta^{*}\alpha)|-\alpha\rangle\langle\alpha+2\beta|,
M(4)(0)\displaystyle M^{(4)}(0) =\displaystyle= |α+2βα+2β|.\displaystyle|\alpha+2\beta\rangle\langle\alpha+2\beta|. (79)

Because of Eq. (9), time evolution from time t1=0t_{1}=0 to time t2=τt_{2}=\tau of the above operators, {K(j)(τ):j=1,2,4}\{K^{(j)}(\tau):j=1,2,4\}, {L(j)(τ):j=1,2,3,4}\{L^{(j)}(\tau):j=1,2,3,4\}, and {M(j)(τ):j=1,2,4}\{M^{(j)}(\tau):j=1,2,4\}, are described in the forms,

K(1)(τ)\displaystyle K^{(1)}(\tau) =\displaystyle= |αexp(iΩτ)αexp(iΩτ)|,\displaystyle|\alpha\exp(-i\Omega\tau)\rangle\langle\alpha\exp(-i\Omega\tau)|,
K(2)(τ)\displaystyle K^{(2)}(\tau) =\displaystyle= exp{βαβα\displaystyle\exp\{\beta\alpha^{*}-\beta^{*}\alpha
(1/2)[|α|2+|α+2β|22α(α+2β)][1exp(2Γτ)]}\displaystyle-(1/2)[|\alpha|^{2}+|-\alpha+2\beta|^{2}-2\alpha(-\alpha^{*}+2\beta^{*})][1-\exp(-2\Gamma\tau)]\}
×|αexp(iΩτ)(α+2β)exp(iΩτ)|,\displaystyle\times|\alpha\exp(-i\Omega\tau)\rangle\langle(-\alpha+2\beta)\exp(-i\Omega\tau)|,
K(4)(τ)\displaystyle K^{(4)}(\tau) =\displaystyle= |(α+2β)exp(iΩτ)(α+2β)exp(iΩτ)|,\displaystyle|(-\alpha+2\beta)\exp(-i\Omega\tau)\rangle\langle(-\alpha+2\beta)\exp(-i\Omega\tau)|, (80)
L(1)(τ)\displaystyle L^{(1)}(\tau) =\displaystyle= exp{2|α|2[1exp(2Γτ)]}|αexp(iΩτ)αexp(iΩτ)|,\displaystyle\exp\{-2|\alpha|^{2}[1-\exp(-2\Gamma\tau)]\}|\alpha\exp(-i\Omega\tau)\rangle\langle-\alpha\exp(-i\Omega\tau)|,
L(2)(τ)\displaystyle L^{(2)}(\tau) =\displaystyle= exp{βα+βα(1/2)[|α|2+|α+2β|22α(α+2β)][1exp(2Γτ)]}\displaystyle\exp\{-\beta\alpha^{*}+\beta^{*}\alpha-(1/2)[|\alpha|^{2}+|\alpha+2\beta|^{2}-2\alpha(\alpha^{*}+2\beta^{*})][1-\exp(-2\Gamma\tau)]\}
×|αexp(iΩτ)(α+2β)exp(iΩτ)|,\displaystyle\times|\alpha\exp(-i\Omega\tau)\rangle\langle(\alpha+2\beta)\exp(-i\Omega\tau)|,
L(3)(τ)\displaystyle L^{(3)}(\tau) =\displaystyle= exp{βα+βα\displaystyle\exp\{-\beta\alpha^{*}+\beta^{*}\alpha
(1/2)[|α+2β|2+|α|2+2(α+2β)α][1exp(2Γτ)]}\displaystyle-(1/2)[|-\alpha+2\beta|^{2}+|\alpha|^{2}+2(-\alpha+2\beta)\alpha^{*}][1-\exp(-2\Gamma\tau)]\}
×|(α+2β)exp(iΩτ)αexp(iΩτ)|,\displaystyle\times|(-\alpha+2\beta)\exp(-i\Omega\tau)\rangle\langle-\alpha\exp(-i\Omega\tau)|,
L(4)(τ)\displaystyle L^{(4)}(\tau) =\displaystyle= exp{2(βα+βα)\displaystyle\exp\{2(-\beta\alpha^{*}+\beta^{*}\alpha) (81)
(1/2)[|α+2β|2+|α+2β|22(α+2β)(α+2β)][1exp(2Γτ)]}\displaystyle-(1/2)[|-\alpha+2\beta|^{2}+|\alpha+2\beta|^{2}-2(-\alpha+2\beta)(\alpha^{*}+2\beta^{*})][1-\exp(-2\Gamma\tau)]\}
×|(α+2β)exp(iΩτ)(α+2β)exp(iΩτ)|,\displaystyle\times|(-\alpha+2\beta)\exp(-i\Omega\tau)\rangle\langle(\alpha+2\beta)\exp(-i\Omega\tau)|,
M(1)(τ)\displaystyle M^{(1)}(\tau) =\displaystyle= |αexp(iΩτ)αexp(iΩτ)|,\displaystyle|-\alpha\exp(-i\Omega\tau)\rangle\langle-\alpha\exp(-i\Omega\tau)|,
M(2)(τ)\displaystyle M^{(2)}(\tau) =\displaystyle= exp{βα+βα(1/2)[|α|2+|α+2β|2+2α(α+2β)][1exp(2Γτ)]}\displaystyle\exp\{-\beta\alpha^{*}+\beta^{*}\alpha-(1/2)[|\alpha|^{2}+|\alpha+2\beta|^{2}+2\alpha(\alpha^{*}+2\beta^{*})][1-\exp(-2\Gamma\tau)]\}
×|αexp(iΩτ)(α+2β)exp(iΩτ)|,\displaystyle\times|-\alpha\exp(-i\Omega\tau)\rangle\langle(\alpha+2\beta)\exp(-i\Omega\tau)|,
M(4)(τ)\displaystyle M^{(4)}(\tau) =\displaystyle= |(α+2β)exp(iΩτ)(α+2β)exp(iΩτ)|.\displaystyle|(\alpha+2\beta)\exp(-i\Omega\tau)\rangle\langle(\alpha+2\beta)\exp(-i\Omega\tau)|. (82)

A mathematically rigorous form of p1±,2+p_{1\pm,2+}, the probability that O2=1O_{2}=1 is obtained with the measurement on w1±(τ)w_{1\pm}(\tau) at time t2t_{2}, is given by

p1±,2+\displaystyle p_{1\pm,2+} =\displaystyle= (1/4)q(α)1\displaystyle(1/4)q(\alpha)^{-1}
×(Tr[K(1)(τ)Π(+)(β)]±2Re{Tr[K(2)(τ)Π(+)(β)]}+Tr[K(4)(τ)Π(+)(β)]\displaystyle\times\Bigl{(}\mbox{Tr}[K^{(1)}(\tau)\Pi^{(+)}(\beta)]\pm 2\mbox{Re}\{\mbox{Tr}[K^{(2)}(\tau)\Pi^{(+)}(\beta)]\}+\mbox{Tr}[K^{(4)}(\tau)\Pi^{(+)}(\beta)]
+2Re{Tr[L(1)(τ)Π(+)(β)]±Tr[L(2)(τ)Π(+)(β)]±Tr[L(3)(τ)Π(+)(β)]\displaystyle+2\mbox{Re}\{\mbox{Tr}[L^{(1)}(\tau)\Pi^{(+)}(\beta)]\pm\mbox{Tr}[L^{(2)}(\tau)\Pi^{(+)}(\beta)]\pm\mbox{Tr}[L^{(3)}(\tau)\Pi^{(+)}(\beta)]
+Tr[L(4)(τ)Π(+)(β)]}\displaystyle+\mbox{Tr}[L^{(4)}(\tau)\Pi^{(+)}(\beta)]\}
+Tr[M(1)(τ)Π(+)(β)]±2Re{Tr[M(2)(τ)Π(+)(β)]}+Tr[M(4)(τ)Π(+)(β)]),\displaystyle+\mbox{Tr}[M^{(1)}(\tau)\Pi^{(+)}(\beta)]\pm 2\mbox{Re}\{\mbox{Tr}[M^{(2)}(\tau)\Pi^{(+)}(\beta)]\}+\mbox{Tr}[M^{(4)}(\tau)\Pi^{(+)}(\beta)]\Bigr{)},

where explicit forms of {Tr[K(j)(τ)Π(+)(β)]:j=1,2,4}\{\mbox{Tr}[K^{(j)}(\tau)\Pi^{(+)}(\beta)]:j=1,2,4\}, {Tr[L(j)(τ)Π(+)(β)]:j=1,2,3,4}\{\mbox{Tr}[L^{(j)}(\tau)\Pi^{(+)}(\beta)]:j=1,2,3,4\}, and {Tr[M(j)(τ)Π(+)(β)]:j=1,2,4}\{\mbox{Tr}[M^{(j)}(\tau)\Pi^{(+)}(\beta)]:j=1,2,4\} are written as

Tr[K(1)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[K^{(1)}(\tau)\Pi^{(+)}(\beta)] =\displaystyle= exp[|αexp(iΩτ)β|2]cosh|αexp(iΩτ)β|2,\displaystyle\exp[-|\alpha\exp(-i\Omega\tau)-\beta|^{2}]\cosh|\alpha\exp(-i\Omega\tau)-\beta|^{2},
Tr[K(2)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[K^{(2)}(\tau)\Pi^{(+)}(\beta)] =\displaystyle= exp{βαβα\displaystyle\exp\{\beta\alpha^{*}-\beta^{*}\alpha
(1/2)[|α|2+|α+2β|22α(α+2β)][1exp(2Γτ)]\displaystyle-(1/2)[|\alpha|^{2}+|-\alpha+2\beta|^{2}-2\alpha(-\alpha^{*}+2\beta^{*})][1-\exp(-2\Gamma\tau)]
[βαexp(iΩτ)βαexp(iΩτ)]+2i|β|2sinωτexp(Γτ)\displaystyle-[\beta\alpha^{*}\exp(i\Omega^{*}\tau)-\beta^{*}\alpha\exp(-i\Omega\tau)]+2i|\beta|^{2}\sin\omega\tau\exp(-\Gamma\tau)
(1/2)[|αexp(iΩτ)β|2+|(α+2β)exp(iΩτ)β|2]}\displaystyle-(1/2)[|\alpha\exp(-i\Omega\tau)-\beta|^{2}+|(-\alpha+2\beta)\exp(-i\Omega\tau)-\beta|^{2}]\}
×cosh{[(α+2β)exp(iΩτ)β][αexp(iΩτ)β]},\displaystyle\times\cosh\{[(-\alpha^{*}+2\beta^{*})\exp(i\Omega^{*}\tau)-\beta^{*}][\alpha\exp(-i\Omega\tau)-\beta]\},
Tr[K(4)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[K^{(4)}(\tau)\Pi^{(+)}(\beta)] =\displaystyle= exp[|(α+2β)exp(iΩτ)β|2]\displaystyle\exp[-|(-\alpha+2\beta)\exp(-i\Omega\tau)-\beta|^{2}] (84)
×cosh|(α+2β)exp(iΩτ)β|2,\displaystyle\times\cosh|(-\alpha+2\beta)\exp(-i\Omega\tau)-\beta|^{2},
Tr[L(1)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[L^{(1)}(\tau)\Pi^{(+)}(\beta)] =\displaystyle= exp{2|α|2[1exp(2Γτ)][βαexp(iΩτ)βαexp(iΩτ)]\displaystyle\exp\{-2|\alpha|^{2}[1-\exp(-2\Gamma\tau)]-[\beta\alpha^{*}\exp(i\Omega^{*}\tau)-\beta^{*}\alpha\exp(-i\Omega\tau)]
(1/2)[|αexp(iΩτ)β|2+|αexp(iΩτ)+β|2]}\displaystyle-(1/2)[|\alpha\exp(-i\Omega\tau)-\beta|^{2}+|\alpha\exp(-i\Omega\tau)+\beta|^{2}]\}
×cosh{[αexp(iΩτ)+β][αexp(iΩτ)β]},\displaystyle\times\cosh\{-[\alpha^{*}\exp(i\Omega^{*}\tau)+\beta^{*}][\alpha\exp(-i\Omega\tau)-\beta]\},
Tr[L(2)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[L^{(2)}(\tau)\Pi^{(+)}(\beta)] =\displaystyle= exp{βα+βα\displaystyle\exp\{-\beta\alpha^{*}+\beta^{*}\alpha
(1/2)[|α|2+|α+2β|22α(α+2β)][1exp(2Γτ)]\displaystyle-(1/2)[|\alpha|^{2}+|\alpha+2\beta|^{2}-2\alpha(\alpha^{*}+2\beta^{*})][1-\exp(-2\Gamma\tau)]
+2i|β|2sinωτexp(Γτ)\displaystyle+2i|\beta|^{2}\sin\omega\tau\exp(-\Gamma\tau)
(1/2)[|αexp(iΩτ)β|2+|(α+2β)exp(iΩτ)β|2]}\displaystyle-(1/2)[|\alpha\exp(-i\Omega\tau)-\beta|^{2}+|(\alpha+2\beta)\exp(-i\Omega\tau)-\beta|^{2}]\}
×cosh{[(α+2β)exp(iΩτ)β][αexp(iΩτ)β]},\displaystyle\times\cosh\{[(\alpha^{*}+2\beta^{*})\exp(i\Omega^{*}\tau)-\beta^{*}][\alpha\exp(-i\Omega\tau)-\beta]\},
Tr[L(3)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[L^{(3)}(\tau)\Pi^{(+)}(\beta)] =\displaystyle= exp{βα+βα\displaystyle\exp\{-\beta\alpha^{*}+\beta^{*}\alpha
(1/2)[|α+2β|2+|α|2+2(α+2β)α][1exp(2Γτ)]\displaystyle-(1/2)[|-\alpha+2\beta|^{2}+|\alpha|^{2}+2(-\alpha+2\beta)\alpha^{*}][1-\exp(-2\Gamma\tau)]
2i|β|2sinωτexp(Γτ)\displaystyle-2i|\beta|^{2}\sin\omega\tau\exp(-\Gamma\tau)
(1/2)[|(α+2β)exp(iΩτ)β|2+|αexp(iΩτ)+β|2]}\displaystyle-(1/2)[|(-\alpha+2\beta)\exp(-i\Omega\tau)-\beta|^{2}+|\alpha\exp(-i\Omega\tau)+\beta|^{2}]\}
×cosh{[αexp(iΩτ)+β][(α+2β)exp(iΩτ)β]},\displaystyle\times\cosh\{-[\alpha^{*}\exp(i\Omega^{*}\tau)+\beta^{*}][(-\alpha+2\beta)\exp(-i\Omega\tau)-\beta]\},
Tr[L(4)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[L^{(4)}(\tau)\Pi^{(+)}(\beta)] =\displaystyle= exp{2(βα+βα)\displaystyle\exp\{2(-\beta\alpha^{*}+\beta^{*}\alpha)
(1/2)[|α+2β|2+|α+2β|22(α+2β)(α+2β)]\displaystyle-(1/2)[|-\alpha+2\beta|^{2}+|\alpha+2\beta|^{2}-2(-\alpha+2\beta)(\alpha^{*}+2\beta^{*})]
×[1exp(2Γτ)]\displaystyle\times[1-\exp(-2\Gamma\tau)]
[βαexp(iΩτ)+βαexp(iΩτ)]\displaystyle-[-\beta\alpha^{*}\exp(i\Omega^{*}\tau)+\beta^{*}\alpha\exp(-i\Omega\tau)]
(1/2)[|(α+2β)exp(iΩτ)β|2\displaystyle-(1/2)[|(-\alpha+2\beta)\exp(-i\Omega\tau)-\beta|^{2}
+|(α+2β)exp(iΩτ)β|2]}\displaystyle+|(\alpha+2\beta)\exp(-i\Omega\tau)-\beta|^{2}]\}
×cosh{[(α+2β)exp(iΩτ)β][(α+2β)exp(iΩτ)β]},\displaystyle\times\cosh\{[(\alpha^{*}+2\beta^{*})\exp(i\Omega^{*}\tau)-\beta^{*}][(-\alpha+2\beta)\exp(-i\Omega\tau)-\beta]\},
Tr[M(1)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[M^{(1)}(\tau)\Pi^{(+)}(\beta)] =\displaystyle= exp[|αexp(iΩτ)+β|2]cosh|αexp(iΩτ)+β|2,\displaystyle\exp[-|\alpha\exp(-i\Omega\tau)+\beta|^{2}]\cosh|\alpha\exp(-i\Omega\tau)+\beta|^{2},
Tr[M(2)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[M^{(2)}(\tau)\Pi^{(+)}(\beta)] =\displaystyle= exp{βα+βα(1/2)[|α|2+|α+2β|2+2α(α+2β)]\displaystyle\exp\{-\beta\alpha^{*}+\beta^{*}\alpha-(1/2)[|\alpha|^{2}+|\alpha+2\beta|^{2}+2\alpha(\alpha^{*}+2\beta^{*})]
×[1exp(2Γτ)]\displaystyle\times[1-\exp(-2\Gamma\tau)]
+[βαexp(iΩτ)βαexp(iΩτ)]+2i|β|2sinωτexp(Γτ)\displaystyle+[\beta\alpha^{*}\exp(i\Omega^{*}\tau)-\beta^{*}\alpha\exp(-i\Omega\tau)]+2i|\beta|^{2}\sin\omega\tau\exp(-\Gamma\tau)
(1/2)[|αexp(iΩτ)+β|2+|(α+2β)exp(iΩτ)β|2]}\displaystyle-(1/2)[|\alpha\exp(-i\Omega\tau)+\beta|^{2}+|(\alpha+2\beta)\exp(-i\Omega\tau)-\beta|^{2}]\}
×cosh{[(α+2β)exp(iΩτ)β][αexp(iΩτ)+β]},\displaystyle\times\cosh\{-[(\alpha^{*}+2\beta^{*})\exp(i\Omega^{*}\tau)-\beta^{*}][\alpha\exp(-i\Omega\tau)+\beta]\},
Tr[M(4)(τ)Π(+)(β)]\displaystyle\mbox{Tr}[M^{(4)}(\tau)\Pi^{(+)}(\beta)] =\displaystyle= exp[|(α+2β)exp(iΩτ)β|2]\displaystyle\exp[-|(\alpha+2\beta)\exp(-i\Omega\tau)-\beta|^{2}] (86)
×cosh|(α+2β)exp(iΩτ)β|2.\displaystyle\times\cosh|(\alpha+2\beta)\exp(-i\Omega\tau)-\beta|^{2}.

An explicit form of p1±,2p_{1\pm,2-}, the probability that O2=1O_{2}=-1 is obtained with the measurement on w1±(τ)w_{1\pm}(\tau) at time t2t_{2}, is described in the form,

p1±,2\displaystyle p_{1\pm,2-} =\displaystyle= (1/4)q(α)1\displaystyle(1/4)q(\alpha)^{-1}
×(Tr[K(1)(τ)Π()(β)]±2Re{Tr[K(2)(τ)Π()(β)]}+Tr[K(4)(τ)Π()(β)]\displaystyle\times\Bigl{(}\mbox{Tr}[K^{(1)}(\tau)\Pi^{(-)}(\beta)]\pm 2\mbox{Re}\{\mbox{Tr}[K^{(2)}(\tau)\Pi^{(-)}(\beta)]\}+\mbox{Tr}[K^{(4)}(\tau)\Pi^{(-)}(\beta)]
+2Re{Tr[L(1)(τ)Π()(β)]±Tr[L(2)(τ)Π()(β)]±Tr[L(3)(τ)Π()(β)]\displaystyle+2\mbox{Re}\{\mbox{Tr}[L^{(1)}(\tau)\Pi^{(-)}(\beta)]\pm\mbox{Tr}[L^{(2)}(\tau)\Pi^{(-)}(\beta)]\pm\mbox{Tr}[L^{(3)}(\tau)\Pi^{(-)}(\beta)]
+Tr[L(4)(τ)Π()(β)]}\displaystyle+\mbox{Tr}[L^{(4)}(\tau)\Pi^{(-)}(\beta)]\}
+Tr[M(1)(τ)Π()(β)]±2Re{Tr[M(2)(τ)Π()(β)]}+Tr[M(4)(τ)Π()(β)]),\displaystyle+\mbox{Tr}[M^{(1)}(\tau)\Pi^{(-)}(\beta)]\pm 2\mbox{Re}\{\mbox{Tr}[M^{(2)}(\tau)\Pi^{(-)}(\beta)]\}+\mbox{Tr}[M^{(4)}(\tau)\Pi^{(-)}(\beta)]\Bigr{)},

where {Tr[K(j)(τ)Π()(β)]:j=1,2,4}\{\mbox{Tr}[K^{(j)}(\tau)\Pi^{(-)}(\beta)]:j=1,2,4\}, {Tr[L(j)(τ)Π()(β)]:j=1,2,3,4}\{\mbox{Tr}[L^{(j)}(\tau)\Pi^{(-)}(\beta)]:j=1,2,3,4\}, and
{Tr[M(j)(τ)Π()(β)]:j=1,2,4}\{\mbox{Tr}[M^{(j)}(\tau)\Pi^{(-)}(\beta)]:j=1,2,4\} are given by {Tr[K(j)(τ)Π(+)(β)]:j=1,2,4}\{\mbox{Tr}[K^{(j)}(\tau)\Pi^{(+)}(\beta)]:j=1,2,4\},
{Tr[L(j)(τ)Π(+)(β)]:j=1,2,3,4}\{\mbox{Tr}[L^{(j)}(\tau)\Pi^{(+)}(\beta)]:j=1,2,3,4\}, and {Tr[M(j)(τ)Π(+)(β)]:j=1,2,4}\{\mbox{Tr}[M^{(j)}(\tau)\Pi^{(+)}(\beta)]:j=1,2,4\} in Eqs. (84), (LABEL:L-formula), and (86) with replacing cosh\cosh with sinh\sinh, that is substitution of hyperbolic sines for hyperbolic cosines.

Explicit forms of p2±,3+p_{2\pm,3+} and p2±,3p_{2\pm,3-}, the probabilities that O3=1O_{3}=1 and O3=1O_{3}=-1 are obtained with the measurement on w2±(2τ)w_{2\pm}(2\tau) at t3t_{3} respectively, are given by

p2±,3+\displaystyle p_{2\pm,3+} =\displaystyle= (1/4)q(α)1\displaystyle(1/4)q(\alpha)^{-1}
×(Tr[K~(1)(τ)Π(+)(β)]±2Re{Tr[K~(2)(τ)Π(+)(β)]}+Tr[K~(4)(τ)Π(+)(β)]\displaystyle\times\Bigl{(}\mbox{Tr}[\tilde{K}^{(1)}(\tau)\Pi^{(+)}(\beta)]\pm 2\mbox{Re}\{\mbox{Tr}[\tilde{K}^{(2)}(\tau)\Pi^{(+)}(\beta)]\}+\mbox{Tr}[\tilde{K}^{(4)}(\tau)\Pi^{(+)}(\beta)]
+2exp[2|α|2(1e2Γτ)]\displaystyle+2\exp[-2|\alpha|^{2}(1-e^{-2\Gamma\tau})]
×Re{Tr[L~(1)(τ)Π(+)(β)]±Tr[L~(2)(τ)Π(+)(β)]±Tr[L~(3)(τ)Π(+)(β)]\displaystyle\times\mbox{Re}\{\mbox{Tr}[\tilde{L}^{(1)}(\tau)\Pi^{(+)}(\beta)]\pm\mbox{Tr}[\tilde{L}^{(2)}(\tau)\Pi^{(+)}(\beta)]\pm\mbox{Tr}[\tilde{L}^{(3)}(\tau)\Pi^{(+)}(\beta)]
+Tr[L~(4)(τ)Π(+)(β)]}\displaystyle+\mbox{Tr}[\tilde{L}^{(4)}(\tau)\Pi^{(+)}(\beta)]\}
+Tr[M~(1)(τ)Π(+)(β)]±2Re{Tr[M~(2)(τ)Π(+)(β)]}+Tr[M~(4)(τ)Π(+)(β)]),\displaystyle+\mbox{Tr}[\tilde{M}^{(1)}(\tau)\Pi^{(+)}(\beta)]\pm 2\mbox{Re}\{\mbox{Tr}[\tilde{M}^{(2)}(\tau)\Pi^{(+)}(\beta)]\}+\mbox{Tr}[\tilde{M}^{(4)}(\tau)\Pi^{(+)}(\beta)]\Bigr{)},
p2±,3\displaystyle p_{2\pm,3-} =\displaystyle= (1/4)q(α)1\displaystyle(1/4)q(\alpha)^{-1}
×(Tr[K~(1)(τ)Π()(β)]±2Re{Tr[K~(2)(τ)Π()(β)]}+Tr[K~(4)(τ)Π()(β)]\displaystyle\times\Bigl{(}\mbox{Tr}[\tilde{K}^{(1)}(\tau)\Pi^{(-)}(\beta)]\pm 2\mbox{Re}\{\mbox{Tr}[\tilde{K}^{(2)}(\tau)\Pi^{(-)}(\beta)]\}+\mbox{Tr}[\tilde{K}^{(4)}(\tau)\Pi^{(-)}(\beta)]
+2exp[2|α|2(1e2Γτ)]\displaystyle+2\exp[-2|\alpha|^{2}(1-e^{-2\Gamma\tau})]
×Re{Tr[L~(1)(τ)Π()(β)]±Tr[L~(2)(τ)Π()(β)]±Tr[L~(3)(τ)Π()(β)]\displaystyle\times\mbox{Re}\{\mbox{Tr}[\tilde{L}^{(1)}(\tau)\Pi^{(-)}(\beta)]\pm\mbox{Tr}[\tilde{L}^{(2)}(\tau)\Pi^{(-)}(\beta)]\pm\mbox{Tr}[\tilde{L}^{(3)}(\tau)\Pi^{(-)}(\beta)]
+Tr[L~(4)(τ)Π()(β)]}\displaystyle+\mbox{Tr}[\tilde{L}^{(4)}(\tau)\Pi^{(-)}(\beta)]\}
+Tr[M~(1)(τ)Π()(β)]±2Re{Tr[M~(2)(τ)Π()(β)]}+Tr[M~(4)(τ)Π()(β)]),\displaystyle+\mbox{Tr}[\tilde{M}^{(1)}(\tau)\Pi^{(-)}(\beta)]\pm 2\mbox{Re}\{\mbox{Tr}[\tilde{M}^{(2)}(\tau)\Pi^{(-)}(\beta)]\}+\mbox{Tr}[\tilde{M}^{(4)}(\tau)\Pi^{(-)}(\beta)]\Bigr{)},

where

Tr[K~(j)(τ)Π(±)(β)]\displaystyle\mbox{Tr}[\tilde{K}^{(j)}(\tau)\Pi^{(\pm)}(\beta)] =\displaystyle= Tr[K(j)(τ)Π(±)(β)]|ααexp(iΩτ),\displaystyle\left.\mbox{Tr}[K^{(j)}(\tau)\Pi^{(\pm)}(\beta)]\right|_{\alpha\rightarrow\alpha\exp(-i\Omega\tau)},
Tr[L~(j)(τ)Π(±)(β)]\displaystyle\mbox{Tr}[\tilde{L}^{(j)}(\tau)\Pi^{(\pm)}(\beta)] =\displaystyle= Tr[L(j)(τ)Π(±)(β)]|ααexp(iΩτ),\displaystyle\left.\mbox{Tr}[L^{(j)}(\tau)\Pi^{(\pm)}(\beta)]\right|_{\alpha\rightarrow\alpha\exp(-i\Omega\tau)},
Tr[M~(j)(τ)Π(±)(β)]\displaystyle\mbox{Tr}[\tilde{M}^{(j)}(\tau)\Pi^{(\pm)}(\beta)] =\displaystyle= Tr[M(j)(τ)Π(±)(β)]|ααexp(iΩτ)for j=1,2,3,4.\displaystyle\left.\mbox{Tr}[M^{(j)}(\tau)\Pi^{(\pm)}(\beta)]\right|_{\alpha\rightarrow\alpha\exp(-i\Omega\tau)}\quad\mbox{for $j=1,2,3,4$}. (90)

References

  • [1] J.S. Bell, ‘On the Einstein Podolsky Rosen paradox’, Physics 1(3), 195–200 (1964).
    doi:10.1103/PhysicsPhysiqueFizika.1.195
  • [2] A.J. Leggett and A. Garg, ‘Quantum mechanics versus macroscopic realism: Is the flux there when nobody looks?’, Phys. Rev. Lett. 54(9), 857–860 (1985). doi:10.1103/PhysRevLett.54.857
  • [3] Č. Brukner, S. Taylor, S. Cheung, and V. Vedral, ‘Quantum entanglement in time’, arXiv:quant-ph/0402127.
  • [4] C. Emary, N. Lambert, and F. Nori, ‘Leggett-Garg inequalities’, Rep. Prog. Phys. 77(1), 016001 (2014). doi:10.1088/0034-4885/77/1/016001
  • [5] S.F. Huelga, T.W. Marshall, and E. Santos, ‘Proposed test for realist theories using Rydberg atoms coupled to a high-QQ resonator’, Phys. Rev. A 52(4), R2497–R2500 (1995). doi:10.1103/PhysRevA.52.R2497
  • [6] H. Chevalier, A.J. Paige, H. Kwon, and M.S. Kim, ‘Violating the Leggett-Garg inequalities with classical light’, arXiv:2009.02219.
  • [7] Y.-N. Chen, C.-M. Li, N. Lambert, S.-L. Chen, Y. Ota, G.-Y. Chen, and F. Nori, ‘Temporal steering inequality’, Phys. Rev. A 89(3), 032112 (2014).
    doi:10.1103/PhysRevA.89.032112
  • [8] M. Łobejko, J. Łuczka, and J. Dajka,
    ‘Leggett-Garg inequality for qubits coupled to thermal environment’, Phys. Rev. A 91(4), 042113 (2015). doi:10.1103/PhysRevA.91.042113
  • [9] A. Friedenberger and E. Lutz, ‘Assessing the quantumness of a damped two-level system’, Phys. Rev. A 95(2), 022101 (2017). doi:10.1103/PhysRevA.95.022101
  • [10] H. Azuma and M. Ban, ‘The Leggett-Garg inequalities and the relative entropy of coherence in the Bixon-Jortner model’, Eur. Phys. J. D 72(10), 187 (2018). doi: 10.1140/epjd/e2018-90275-7
  • [11] M. Thenabadu and M.D. Reid, ‘Leggett-Garg tests of macrorealism for dynamical cat states evolving in a nonlinear medium’, Phys. Rev. A 99(3), 032125 (2019). doi:10.1103/PhysRevA.99.032125
  • [12] A. Palacios-Laloy, F. Mallet, F. Nguyen, P. Bertet, D. Vion, D. Esteve, and A.N. Korotkov, ‘Experimental violation of a Bell’s inequality in time with weak measurement’, Nat. Phys. 6, 442–447 (2010). doi:10.1038/nphys1641
  • [13] M.E. Goggin, M.P. Almeida, M. Barbieri, B.P. Lanyon, J.L. O’Brien, A.G. White, and G.J. Pryde, ‘Violation of the Leggett-Garg inequality with weak
    measurements of photons’, Proc. Natl. Acad. Sci. USA 108(4), 1256–1261 (2011). doi:10.1073/pnas.1005774108
  • [14] G.C. Knee, S. Simmons, E.M. Gauger, J.J.L. Morton, H. Riemann, N.V. Abrosimov, P. Becker, H.-J. Pohl, K.M. Itoh, M.L.W. Thewalt, G.A.D. Briggs, and S.C. Benjamin, ‘Violation of a Leggett-Garg inequality with ideal non-invasive measurements’, Nat. Commun. 3, 606 (2012). doi:10.1038/ncomms1614
  • [15] D.F. Walls and G.J. Milburn, Quantum optics (Springer-Verlag, Berlin, 1994).
  • [16] S.M. Barnett and P.M. Radmore, Methods in theoretical quantum optics (Oxford University Press, Oxford, 1997) Sect. 5.6.
  • [17] B. Yurke and D. Stoler, ‘Generating quantum mechanical superpositions of macroscopically distinguishable states via amplitude dispersion’, Phys. Rev. Lett. 57(1), 13–16 (1986). doi:10.1103/PhysRevLett.57.13
  • [18] W. Schleich, M. Pernigo, and F.L. Kien, ‘Nonclassical state from two pseudoclassical states’, Phys. Rev. A 44(3), 2172–2187 (1991). doi:10.1103/PhysRevA.44.2172
  • [19] V. Bužek, A. Vidiella-Barranco, and P.L. Knight, ‘Superpositions of
    coherent states: squeezing and dissipation’, Phys. Rev. A 45(9), 6570–6585 (1992). doi:10.1103/PhysRevA.45.6570
  • [20] M.S. Kim and V. Bužek, ‘Schrödinger-cat states at finite temperature: influence of a finite-temperature heat bath on quantum interferences’, Phys. Rev. A 46(7), 4239–4251 (1992). doi:10.1103/PhysRevA.46.4239
  • [21] R.F. Bishop and A. Vourdas, ‘Displaced and squeezed parity operator: its role in classical mappings of quantum theories’, Phys. Rev. A 50(6), 4488–4501 (1994). doi:10.1103/PhysRevA.50.4488
  • [22] K. Banaszek and K. Wódkiewicz, ‘Nonlocality of the Einstein-Podolsky-Rosen state in the Wigner representation’, Phys. Rev. A 58(6), 4345–4347 (1998).
    doi:10.1103/PhysRevA.58.4345
  • [23] K. Banaszek and K. Wódkiewicz, ‘Testing quantum nonlocality in phase space’, Phys. Rev. Lett. 82(10), 2009–2013 (1999). doi:10.1103/PhysRevLett.82.2009
  • [24] M.S. Kim and J. Lee, ‘Test of quantum nonlocality for cavity fields’, Phys. Rev. A 61(4), 042102 (2000). doi:10.1103/PhysRevA.61.042102
  • [25] H. Jeong, W. Son, M.S. Kim, D. Ahn, and Č. Brukner, ‘Quantum nonlocality test for continuous-variable states with dichotomic observables’, Phys. Rev. A 67(1), 012106 (2003). doi:10.1103/PhysRevA.67.012106
  • [26] N. Lambert, C. Emary, Y.-N. Chen, and F. Nori, ‘Distinguishing quantum and classical transport through nanostructures’, Phys. Rev. Lett. 105(17), 176801 (2010). doi:10.1103/PhysRevLett.105.176801
  • [27] N. Lambert, Y.-N. Chen, and F. Nori, ‘Unified single-photon and single-electron counting statistics: from cavity QED to electron transport’, Phys. Rev. A 82(6), 063840 (2010). doi:10.1103/PhysRevA.82.063840
  • [28] G.-Y. Chen, N. Lambert, C.-M. Li, Y.-N. Chen, and F. Nori, ‘Delocalized single-photon Dicke states and the Leggett-Garg inequality in solid state systems’, Sci. Rep. 2, 869 (2012). doi:10.1038/srep00869
  • [29] C.-M. Li, N. Lambert, Y.-N. Chen, G.-Y. Chen, and F. Nori, ‘Witnessing quantum coherence: from solid-state to biological systems’, Sci. Rep. 2, 885 (2012). doi:10.1038/srep00885
  • [30] C. Emary, N. Lambert, and F. Nori, ‘Leggett-Garg inequality in electron interferometers’, Phys. Rev. B 86(23), 235447 (2012). doi:10.1103/PhysRevB.86.235447
  • [31] J. Kofler and Č. Brukner, ‘Classical world arising out of quantum physics under the restriction of coarse-grained measurements’, Phys. Rev. Lett. 99(18), 180403 (2007). doi:10.1103/PhysRevLett.99.180403
  • [32] S.V. Moreira, A. Keller, T. Coudreau, and P. Milman,
    ‘Modeling Leggett-Garg-inequality violation’, Phys. Rev. A 92(6), 062132 (2015). doi:10.1103/PhysRevA.92.062132
  • [33] H.-Y. Ku, N. Lambert, F.-J. Chan, C. Emary, Y.-N. Chen, and F. Nori, ‘Experimental test of non-macrorealistic cat states in the cloud’, npj Quantum Inf. 6, 98 (2020). doi:10.1038/s41534-020-00321-x
  • [34] M.M. Wilde and A. Mizel, ‘Addressing the clumsiness loophole in a Leggett-Garg test of macrorealism’, Found. Phys. 42(2), 256–265 (2012).
    doi:10.1007/s10701-011-9598-4
  • [35] G.C. Knee, K. Kakuyanagi, M.-C. Yeh, Y. Matsuzaki, H. Toida, H. Yamaguchi, S. Saito, A.J. Leggett, and W.J. Munro, ‘A strict experimental test of macroscopic realism in a superconducting flux qubit’, Nat. Commun. 7, 13253 (2016). doi:10.1038/ncomms13253
  • [36] R.J. Birrittella, P.M. Alsing, and C.C. Gerry, ‘The parity operator: applications in quantum metrology featured’, AVS Quantum Sci. 3(1), 014701 (2021). doi:10.1116/5.0026148
  • [37] B.-G. Englert, N. Sterpi, and H. Walther, ‘Parity states in the one-atom maser’, Opt. Commun. 100(5-6), 526–535 (1993). doi:10.1016/0030-4018(93)90254-3
  • [38] L. Cohen, D. Istrati, L. Dovrat, and H.S. Eisenberg, ‘Super-resolved phase measurements at the shot noise limit by parity measurement’, Opt. Express 22(10), 11945–11953 (2014). doi:10.1364/OE.22.011945
  • [39] N. Lambert, R. Johansson, and F. Nori, ‘Macrorealism inequality for
    optoelectromechanical systems’, Phys. Rev. B 84(24), 245421 (2011).
    doi:10.1103/PhysRevB.84.245421
  • [40] J.R. Johansson, N. Lambert, I. Mahboob, H. Yamaguchi, and F. Nori, ‘Entangled-state generation and Bell inequality violations in nanomechanical resonators’, Phys. Rev. B 90(17), 174307 (2014). doi:10.1103/PhysRevB.90.174307
  • [41] C. Budroni and C. Emary, ‘Temporal quantum correlations and Leggett-Garg inequalities in multilevel systems’, Phys. Rev. Lett. 113(5), 050401 (2014).
    doi:10.1103/PhysRevLett.113.050401
  • [42] N. Lambert, K. Debnath, A.F. Kockum, G.C. Knee, W.J. Munro, and F. Nori, ‘Leggett-Garg inequality violations with a large ensemble of qubits’, Phys. Rev. A 94(1), 012105 (2016). doi:10.1103/PhysRevA.94.012105
  • [43] C. Budroni, T. Moroder, M. Kleinmann, and O. Guhne, ‘Bounding
    temporal quantum correlations’, Phys. Rev. Lett. 111(2), 020403 (2013).
    doi:10.1103/PhysRevLett.111.020403