Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Two-mode squeezed states as Schrödinger cat-like states

Open Access Open Access

Abstract

In recent years, there has been an increased interest in the generation of superpositions of coherent states with opposite phases, the so-called photonic Schrödinger cat states. These experiments are challenging, and, so far, cats involving only small photon numbers have been implemented. Here, we propose to consider two-mode squeezed states as examples of Schrödinger cat-like states. For this, we apply criteria that aim to identify macroscopic superpositions in a more general sense. We extend some of these criteria to the two-mode continuous variable regime. Furthermore, we compare the size of states obtained in several experiments and discuss experimental challenges for further improvements. Our results not only promote two-mode squeezed states for exploring quantum effects at the macroscopic level but also provide direct measures to evaluate their usefulness for quantum metrology.

© 2015 Optical Society of America

1. INTRODUCTION

The question of what is a macroscopic quantum state has received a lot of attention over the last decade [117]. The motivation is not to address a new question, as it dates back from the early days of quantum theory [18], but rather comes from the experimental progress, which now allows one to harness large systems while highlighting their quantum nature. Quantum optics experiments reporting on squeezing operations provide a nice example. They are obtained from a χ2 nonlinearity and can result in largely entangled states. The entanglement can be further detected with homodyne detections by means of the Duan–Simon criterion [19,20]. When the χ2 nonlinearity is seeded by coherent states and/or embedded in a high finesse cavity, entanglement in squeezed states can be demonstrated with a large number of photons [2129]. This naturally raises the question of whether squeezed states have macroscopic quantum features, which is a question of deep relevance because, thus far, squeezed states have been combined with conditional detections [3036] for exploring quantum effects in many photon states.

In the literature, there exist different criteria for quantifying the macroscopic quantumness [314]. Typically, this includes a definition that assigns a number to a quantum state, which is here called effective size (or simply size). These criteria can be grouped into two categories. The first addresses the question of whether a two-component superposition |ϕ0+|ϕ1 is macroscopic, i.e., whether |ϕ0 and |ϕ1 are macroscopically distinct. For example, the proposal of [8] states that two spin states are macroscopically distinct if they can be distinguished from a small number of their spins, as a dead cat and a live cat can be distinguished from a small number of their cells. We also can refer to the proposals of [13,37] defining two states as being macroscopically distinct if they can be distinguished with a coarse-grained measurement, as a dead cat and a live cat can be distinguished with a detector having a very limited resolution. The second category aims to identify quantum states that are able to show some kind of macroscopic quantum effect. This term characterizes experimental evidence that cannot be explained by an accumulated quantum effect originated at the microscopic level of the system. For pure states, a large variance with respect to given observables and Hamiltonians is a sufficient signature for quantum fluctuations that are persistent on a macroscopic level. For mixed states, one typically uses a convex function that reduces to the variance for pure states. For example, the proposal of [11] shows how the notion of macroscopicity can be linked to the so-called quantum Fisher information [38].

Not all of these measures are able to correctly handle two-mode squeezed states for different reasons. Thus far, only the contributions [7,10,39] explicitly treat two-mode squeezed vacuum. This is valuable, but these criteria also leave open questions, which are discussed in more detail in the course of this paper. In short, Cavalcanti and Reid [7,39] propose a bound to measure the size of a two-mode squeezed state, which, however, does not give satisfactory answers to nonsqueezed states. This drawback is compensated by Lee and Jeong [10], but the mode additivity of their measure remains problematic.

In this paper, we aim to reinforce arguments in favor of assigning a macroscopic quantum nature to two-mode squeezed states. For this, we extend existing measures for macroscopic quantum states [13,40] to be able to investigate two-mode, continuous variable (cv) quantum states. Apart from avoiding problems that appeared in previous attempts [7,10,39], this broadens the range of arguments supporting the macroscopic quantumness of a two-mode squeezed vacuum. Importantly, we prove that the effective size of two-mode squeezed vacuum states (with N mean photons) is basically the same as superpositions of coherent states with opposite phases |α+|α and |α|2tanh|α|2=N but with the great advantage that they are much easier to create [41]. The tools we propose allow one to bound the size of states obtained experimentally as well as their usefulness for parameter estimation beyond the classical limit. Aside from a fundamental interest, our results ultimately have important applications for quantum metrology. Furthermore, we discuss the impact of noise and technical limitations that cumber the verification of macroscopic quantumness of two-mode squeezed states. Finally, we use data from some performed experiments [2129] to lower-bound the effective size achieved thus far.

2. TWO-MODE VACUUM SQUEEZED STATES

As an example of two-mode squeezed states, let us consider the two-mode squeezed vacuum, which is obtained from a parametric process in which photons from a pump laser decay spontaneously into photon pairs—one in mode 1, its twin in mode 2—while preserving energy and momentum. The corresponding propagator S¯(g)=eg(a1a2a1a2), with squeezing parameter g, applies straightforwardly on the vacuum if written in the normal order. This results in

|ψtms=(1tanh2g)12etanhga1a2|00.
The mean photon number in both modes is N=2tr(a1a1|ψtmsψtms|)=2sinh2g. Furthermore, the variance of the observable X¯1φX¯2ϕ where X¯iθ=12(aieiθ+aieiθ) is given by
Vψtms(X¯1φX¯2ϕ)=cosh2gsinh2gcos(φ+ϕ).
This indicates that the quadratures X¯10X¯20 are correlated whereas X¯1π/2X¯2π/2 are anticorrelated. The quantum nature of these correlations can be revealed through the Duan–Simon criterion [19,20], which states that, for any bipartite separable states and any real parameter a,
Vsep(|a|X¯1ϕ+1aX¯2Φ)+Vsep(|a|X¯1ϕ1aX¯2Φ)>a2[X¯1ϕ,X¯1ϕ]+1a2[X¯2Φ,X¯2Φ]2forϕϕ=ΦΦ=π2,
while for a two-mode squeezed state,
Vψtms(X¯10X¯20)+Vψtms(X¯1π/2+X¯2π/2)=2e2g.
The questions at the core of this paper are as follows: How do we evaluate the size of these kinds of states? Is their effective size comparable with other photonic states?

3. MACROSCOPIC QUANTUMNESS OF TWO-MODE CV STATES

A. Macroscopic Distinctness for cv States

While several definitions have been proposed to identify states that are macroscopically distinct [5,8,9,13], here we focus on the proposal of [13] based on coarse-grained measurements. This choice is arbitrary to some extent. Note, however, that the extension that we propose below easily applies to the measure of [8]. The extension of measures of [9] to two-mode squeezed states is less obvious, as they primarily address spin systems, but the link between measures for spins and photons presented in [17] might be the way to proceed.

The basic principle of the measure of macroscopicity based on coarse-grained measurement is simple. It can be seen as a game where Alice chooses a state in the set {|ϕ0,|ϕ1} with equal a priori probabilities and sends it to Bob. Bob has to guess which one has been sent using a coarse-grained measurement only. It can be any measurement, provided that its resolution is limited. The quantum superposition state |ϕ0+|ϕ1 is qualified macroscopic if Bob wins the game with a detector having no microscopic resolution. Concretely, if one focuses on a noisy photon-counting detector, for example, the size of |ϕ0+|ϕ1 is characterized by the noise that one can tolerate to distinguish |ϕ0 and |ϕ1.

To extend this measure to cv states, we can mimic its original idea by introducing a 50/50 binning of measurement outcomes. For a two-mode squeezed vacuum state in particular, Alice measures her mode with a given quadrature and bins the result with respect to its sign. As Alice’s measurement is assumed to be accurate, this binning corresponds to equiprobable projections onto two orthogonal subspaces of the measured state. Bob has to guess whether she received a positive or negative outcome by measuring his mode with a noisy measurement. The distinguishability of components that Bob receives is again given by the noise that can be tolerated to win the game. Note that the measurement of correlated quadratures maximizes the probability to correctly guess Alice’s outcome. Concretely, the probability that Alice obtains the result x1 and Bob x2 knowing that they measure the quadratures X¯10 and X¯20 is given by |p(x1,x2,σ)|2=tr(|ψtmsψtms|δ(X¯10x1)gσ(X¯20x2)), where gσ stands for the noise of Bob’s measurement device. We assume that gσ is a Gaussian with spread σ and zero mean. Hence, the probability that Bob correctly guesses the sign of Alice’s result is given by Pσguess=0+|p(x1,x2,σ)|2dx1dx2+0|p(x1,x2,σ)|2dx1dx2. We find

Pσguess=12+1πarctan(sinh2g1+2σ2cosh2g).
We can access the maximum noise σmax that Bob can tolerate to win the game with a fixed probability Pσguess by inverting the previous formula:
σmax=1+N(12+N)cotan2(12Pσguess)2+2N.
For comparison, the noise that can be tolerated to win a similar game with the optical Schrödinger cat state (||α||α) is given by
σmax=|α|2(erf1(Pσguess))212.
In both cases, the noise scales like the square root of the photon number. Thus, we claim that two-mode squeezed vacuum and Schrödinger cat states exhibit comparable macroscopic quantumness. In some sense, they belong to the same class of macroscopic quantum states.

Let us now focus on practical considerations. The observation that Alice and Bob’s x quadratures of the two-mode squeezed vacuum state are “macroscopically” correlated (correlated at a large scale, larger than the detector’s resolution) is at the heart of our generalization of the coarse-grained measure. These correlations can be revealed by measuring the joint probability distribution |p(x1,x2,0)|2 with accurate quadrature measurements. (For simplicity, we introduce p(x1,x2)=p(x1,x2,0), which stands for the probability amplitudes without noise.) Although this approach is sufficient to measure the size of a given state in theory, one also has to ensure that those correlations are truly quantum in practice. In mathematical terms, we can always write the state that is shared by Alice and Bob in the x basis:

ρ=p(x1,x2)p*(x¯1,x¯2)f(x1,x¯1,x2,x¯2)|x1,x2x¯1,x¯2|dx1dx2dx¯1dx¯2,
with |p(x1,x2)|2dx1dx2=1 and f(x,x,x,x)=1 x,x. If the shared state is pure, we have f(x1,x¯1,x2,x¯2)=1 x1,x¯1,x2,x¯2, and the correlations revealed through the probability distribution |p(x1,x2)|2 are fully quantum. The violation of the Duan–Simon criterion is then sufficient to attest to the quantum nature of the state for which the size is evaluated through σmax. But how do we certify in practice that the function f(x1,x¯1,x2,x¯2) is close to one, at least in a certain range?

To do so, we consider the effect of imperfect coherences (decoherence) f(x1,x¯1,x2,x¯2)1 on the observed violation of the Duan–Simon witness. Note first that the variance V(X¯10X¯20) can be directly obtained from |p(x1,x2)|2. For the second term required in Eq. (3), we can show that the variance in the presence of decoherence (see Appendix A),

V(X¯1π2+X¯2π2)=V(X¯1π2+X¯2π2)|ideal(x1x¯1+x2x¯2)2f,
equals the ideal-case variance V(X¯1π2+X¯2π2)|ideal plus a factor containing the crossed and second derivatives of f (x1x¯1f(x1,x¯1,x2,x¯2)=dx1dx2|p(x1,x2)|2(x1x¯1f(x1,x¯1,x2,x2))|x1=x¯1, etc. Because V(X¯1π2+X¯2π2)|ideal is positive, we obtain the following upper bound on the observed variance:
(x1x¯1+x2x¯2)2fV(X¯1π2+X¯2π2).
Note that, without further assumptions, we cannot bound the range δ for which f(x1,x1+δ,x2,x2+δ) stays close to one. In other words, even if the state of Alice and Bob largely violates the Duan–Simon witness, the state can be arbitrarily close to a separable one, and p(x1,x2) essentially correspond to classical correlations [42]. However, under the assumption of a Gaussian decay of coherence f(x1,x¯1,x2,x¯2)=e(x1x¯1)2/(2γ12)e(x2x¯2)2/(2γ22), Eq. (7) becomes 1γ12+1γ22V(X¯1π2+X¯2π2). This implies min(γ1,γ2)1/V(X¯1π2+X¯2π2), i.e., if one observes the variance Vψtms(X¯1π2+X¯2π2) of the total momentum, we can certify that the correlations |p(x1,x2)|2 are quantum at least in the range
xC=1Vψtms(X¯1π2+X¯2π2).
Accordingly, if the coherence range xC is lower than the correlation range, as witnessed by σmax, one can only claim that the state exhibits quantum correlations within the range xC, which is then the true size of the state. Revealing the size of large quantum states thus requires us to reveal narrow variances, which becomes increasingly difficult as the size increases (see Section 4).

B. General Measures for Multimode cv States

Beside measures for macroscopic distinguishability, there have been recent proposals that aim to go beyond the basic structure |ϕ0+|ϕ1 [4,6,10,11]. While the measures of [4,6,11] were originally defined for spin systems, the definition of [10] is directly suitable for cv photonic states. For pure states, these three proposals are comparable because a state |ψ is called macroscopically quantum if it shows a large variance with respect to a restricted class of operators. In the spin case, the proposals [4,6,11] focus on sums of local operators (henceforth simply called “local operators”), whereas Lee and Jeong [10] define their measure for pure states proportional to V(X¯0)+V(X¯π/2). In [17], it was argued that local operators in the spin case play to some extent the same role as quadrature operators in mono-mode photonic systems.

The common feature of the proposals for mixed states is that the measures [4,6,10,11] are convex in the state, which is an important and natural feature for the present purpose. There are no clear arguments in favor of one of the proposals. Nevertheless, we focus here on the quantum Fisher information (QFI) [38], which is denoted as Fρ(X¯) for the state ρ and the operator X¯. Importantly, the QFI is the convex roof of the variance [43,44] (up to a factor four); that is, it is the largest convex function that reduces to the variance for pure states. For experiments, it is interesting to note that there exist lower bounds on the QFI based on measurable quantities [45].

The extension to photonic states with n>1 modes is not straightforward. Indeed, a multimode version for the measure of Lee and Jeong was proposed [10]. However, it is additive and, hence, a bunch of “kitten states” |ψαn(|α+|α)n (with potentially small α but large n) is as macroscopically quantum as a “big” single cat state |ψnα|nα+|nα. Here, we propose instead to use a similar account that has been successfully applied in the spin case [4,6,11]. The idea is that the effective size of a product state is the average value of its components, while entangled states should be able to profit from quantum correlations between the modes. Both requirements are achieved by defining the effective size for ρ as

Neff(ρ)=12nmaxθFρ(Xθ),
where Xθ=Σi=1nX¯iθi. In other words, one maximizes the QFI (or the variance for pure states) with respect to sums of local quadrature operators parametrized by θ=(θ1,,θn). The examples from above then lead to Neff(|ψαn)=4|α|2/[1+exp(2|α|2)] and Neff(|ψnα)=4n|α|2/[1+exp(2n|α|2)] (compare to [46]).

We now come to the evaluation of the effective size for the two-mode squeezed vacuum state. It is simple to see that the variance is largest for the quadratures that are maximally correlated. For the state in Eq. (1), these are the operators X¯10+X¯20 and X¯1π/2X¯2π/2. The effective size for each of these choices reads Neff(ψtms)=V(X¯10+X¯20)=e2g2N, which is approximately half of the value as for the cat state with the same photon number, Neff(|ψα)4N. Again, we conclude that two-mode squeezed states are compatible with photonic cat states.

In principle, the effective size of a pure state could be determined by witnessing a large variance for sums of quadrature operators. However, for mixed states, a large variance is not sufficient. Instead, one has to verify a large value of a convex function like the QFI. Since this quantity is typically only accessible through a full state tomography, one has to find other means to estimate it. Recently, a general lower bound on the QFI has been found [45]. It was shown that, for any quantum state ρ and any pair of operators A, B, it holds that Vρ(A)Fρ(B)i[A,B]ρ2, which is a tighter version of the Heisenberg uncertainty relation. Here, we use this inequality to bound the QFI from below. For B=X¯10+X¯20, we set A=X¯1π/2+X¯2π/2 and find i[A,B]=2. Hence, one has

Neff(ρ)1Vρ(X¯1π/2+X¯2π/2).
For the two-mode squeezed state, the anticorrelations between X¯1π/2 and X¯2π/2 lead to a reduced variance and therefore to a potentially large value of Neff.

Note that Eq. (10) [as well as Eq. (8)] resembles the ideas of [7,39] for a generalized notion of macroscopic quantum coherences. However, in these works, solely the bound in Eq. (10) is used, which restricts one to investigate squeezed states. In contrast, we use this expression only to give a bound on the definition in Eq. (9), which also can deal with nonsqueezed states. Hence, our measure is general enough to compare two-mode squeezed states with other states such as cat states.

4. ON THE DIFFICULTY TO CERTIFY THE QUANTUM NATURE OF TWO-MODE SQUEEZED STATES

The common feature of measures for macroscopicity presented before is the requirement to reveal narrow variances, especially when dealing with large size states. How difficult is it in practice? To answer this question, we consider the effect of various experimental imperfections on the observed variance Vψtms(X¯1π/2+X¯2π/2).

  • 1. Consider first a noise along X¯0 that acts on a state ρ as ρdλh(λ)eiX^0λρeiX^0λ with characteristic function (noise distribution) h(λ) of variance Δ2h. The effect of this noise can be directly absorbed in the statistics of the momentum distribution and leads to the following modification of the variance V(X¯1π/2+X¯2π/2)V(X¯1π/2+X¯2π/2)+Δ2h1+Δ2h2. Therefore, if the experimental setup suffers from such a noise, we cannot certify the state of an effective size larger than Neffmax=1Δ2h1+Δ2h2.
  • 2. Similarly, consider a loss channel with transmission η. It leads to V(X¯1π/2+X¯2π/2)ηV(X¯1π/2+X¯2π/2)+(1η), and the maximal certifiable size is given by Neffmax=11η.
  • 3. Now consider a phase noise characterized by the variance Δφ2=p(φ)φ2dφ. It increases the observed variance according to V(X¯1π/2+X¯2π/2)Δφ2(X¯10+X¯20). Specifically, for the two-mode squeezed state, one has Neffmax(ψtms)=1Δφ2(2sinh2(g)+1), which decays exponentially with the squeezing parameter (in the limit of large enough g).

An experimental issue for the detection of highly squeezed states is the size of the local oscillator |α (LO) used for homodyne detection. On the one hand, the mean total photon number N+|α|2 has to be smaller than some value Dth in such a way that the detectors are not saturated. On the other hand, α has to be large enough in order to be a good phase reference because a limited size of the LO is similar to a phase noise. Let us show this quantitatively. In a homodyne measurement, the system is mixed with the LO on a balanced beam splitter with the phase θ. The quadrature is accessed via the difference in the mean photon number of the two outputs of the beam splitter I1I2|α|. While for the mean of any quadrature, this is exactly given by I1I2|α|=Xθ, for the second moment (I1I2)2|α|2=Xθ2+aa|α|2, this correspondence with the measured result is not exact. Accordingly, for a finite size of the local oscillator, the observed variance is bounded by V(X¯1π/2+X¯2π/2)4sinh2(g)|α|2. Because the size of the LO is limited by the detector saturation threshold |α|2Dth2sinh2(g), one has Neffmax(ψtms)=Dth2sinh2(g)4sinh2(g).

In each case, we clearly see that it becomes increasingly difficult to observe narrow variances with two-mode squeezed states as their size increases. This is in agreement with recent results [4750] stating that it is difficult to observe the quantum nature of macroscopic states.

This naturally raises the question of the size of states that can be observed in practice. Note that Eqs. (8) and (10) are general, i.e., the variance along the conjugate of quadratures that are maximally correlated gives a bound on the size of the measured state. We use experimental data obtained in various setups in which the χ2 nonlinearity is either seeded or embedded in a cavity [2124,27,28] to bound the effective size of the produced states (see Fig. 1). All these experiments have in common that the Duan–Simon criterion is used to reveal entanglement and the photon number is large. We clearly see that their effective size cannot be compared with their mean photon number. In the seeded case, the reason is that the seed increases the photon number but does not change the variance. Similarly with a cavity, the photon number can be large even if the gain slightly dominates the loss provided that the cavity finesse is large but the variance of interest is limited by the ratio between the gain and the loss only (see Appendix B). Interestingly, the results presented in Fig. 1 can be used directly to quantitatively estimate the metrological usefulness of states realized experimentally, as the size Neff gives the QFI through the formula in Eq. (9).

 figure: Fig. 1.

Fig. 1. Bounds on the effective size Neff (blue squares) of two-mode squeezed states obtained from experimental data reported in [2129] using the inequality in Eq. (10). The red triangles indicate the minimal photon number N necessary for a cat state |α+|α to have the same effective size according to Eq. (9). For example, the state reported in [21] has a size Neff1.2 for which one needs at least a cat state with N0.2 for the same size.

Download Full Size | PDF

5. CONCLUSION

We propose to consider a two-mode squeezed vacuum as being macroscopically quantum. Our conclusion is based on the results from two measures for macroscopic quantum states [11,13], which have been extended for the purpose of studying two-mode cv states. It is straightforward to generalize our argument to general two-mode squeezed states. This study is in line with former works [7,10,39], but it overcomes some of their problems. On a quantitative level, we see that two-mode squeezed vacuums are comparable with superpositions of coherent states, which are generally considered as archetypal Schrödinger cat states. However, one should acknowledge two advantages of the former compared with the latter: Two-mode squeezed states are not only relatively easy to generate. The experimental verification of macroscopic quantumness can be simply achieved by demonstrating strong (anti-)correlations between the quadratures of the two modes. Nevertheless, in agreement with previous findings [48,50], the precise control of the entire experiment, especially of the measurement, is imperative for showing large quantumness.

APPENDIX A: COHERENCE LENGTH OF CORRELATION

Every two-mode state ρ can be expressed in the joint x basis:

ρ=F(x1,x2,x¯1,x¯2)|x1,x2x¯1,x¯2|dx⃗.
However, for our purpose, it is useful to make the decomposition
F(x1,x2,x¯1,x¯2)=p(x1,x2)p*(x¯1,x¯2)f(x1,x2,x¯1,x¯2),
where we can enforce that |p(x1,x2)|2dx1dx2=1 and f(x1=x¯1,x2=x¯2)=1, and consequently |f(x1x¯1,x2x¯2)|1. The later inequality is ensured by positivity of ρ, i.e., if it does not hold, then there is a state α|x1,x2+β|x¯1,x¯2 that has a negative overlap with ρ. The decomposition in Eq. (A2) is useful because the function f(x1,x2,x¯1,x¯2) can be simply interpreted as characterizing the lack of purity of ρ because, for a pure state |ψ=p(x1,x2)|x1,x2dx1dx2, it satisfies f(x1,x2,x¯1,x¯2)1.

Let us now consider the mean values of p1(2), p1(2)2, and p1p2 on the state ρ. In this appendix, we denote X¯0=x and X¯π/2=p. Using the representation of momenta eigenstate in the x basis |p=12πeixp|x, one obtains

p1=F(x1,x2,x¯1,x¯2)peip(x1x¯1)2πδ(x2x¯2)dpdx⃗,p12=F(x1,x2,x¯1,x¯2)p2eip(x1x¯1)2πδ(x2x¯2)dpdx⃗,p1p2=F(x1,x2,x¯1,x¯2)p1p2eip1(x1x¯1)2π×eip2(x2x¯2)2πdp1dp2dx⃗.
Using pneip(Δx)=(iΔx)neip(Δx) and 12πeip(xx¯)dp=δ(xx¯), a simple integration by parts allows us to rewrite the above expressions as
p1=(ix1x¯1)F(x1,x2,x¯1,x¯2)|x1=x¯1dx1dx2,p12=(ix1x¯1)2F(x1,x2,x¯1,x¯2)|x1=x¯1dx1dx2,p1p2=(ix2x¯2)(ix1x¯1),F(x1,x2,x¯1,x¯2)|x1=x¯1,x2=x¯2dx1dx2.
Those expressions allow us to use the decomposition in Eq. (A2) to its full advantage, leading to
p1=p1f1,p12=p12f1x1x¯12f,p12=p22f1x2x¯22f,p1p2=p1p2f1x1x¯1x2x¯2f,
with the averages ·f1 being taken over the pure state |ψ=p(x1,x2)|x1,x2dx1dx2 and
D[f]=|p(x1,x2)|2D[f](x1,x2)dx1dx2.
Note that, to derive these expressions, we used the fact that the first derivatives of f are zero (because ρ is Hermitian).

The expressions in Eq. (A4) allow us to rewrite the variance of p1+p2 in a form where the contributions of p(x1,x2) and f(x1,x2,x¯1,x¯2) are separated:

V(p1+p2)=V(p1+p2)f1(x1x¯1+x2x¯2)2f.
Keep in mind that, for pure states, the variance V(p1+p2)f1 is always positive. This allows us to upper bound the decay of coherences in the x basis:
(x1x¯1+x2x¯2)2fV(p1+p2).
Without supplementary assumptions, local derivatives of f at x1=x¯1 and x2=x¯2 are not sufficient to determine global properties, such that the variances V(x1x¯1) and V(x2x¯2) of f, as one can imagine irregular functions f that have zero derivatives but arbitrarily small variance V(x1x¯1) (e.g., the step function of arbitrarily small width). But assuming a Gaussian profile for the decay of coherence allows us to draw conclusions on the coherence width of f from the upper bound in Eq. (A7), as we show in the main text.

APPENDIX B: QUADRATURE CORRELATIONS FOR AN AMPLIFIER WITH LOSS

In this appendix, we derive a simple model for a two-mode optical parametric amplification in a cavity with loss. The amplification Hamiltonian is given by

HA=iχ(abab),
with χ>0. The loss is described by a beam splitter operating on each mode a and b. The global process can be seen as a sequence of alternating infinitesimal amplifiers with gain χdt and losses with intensity transmission 12λdt. Consider an operator of the form
O(η,μ,κ)=eκei(ηa+μb)ei(η*a+μ*b)
and propagate it through an elementary step of our process (amplification + loss). It is easy to see that, after an infinitesimal time step dt, the operator becomes
trlossUdtO(η,μ,κ)Udt=O(η+(μ*χηλ)dt,μ+(η*χμλ)dt,κ(μ*η*+μη)χdt),
where we omit terms of higher order in dt in the exponent, and the trace is there to remind us that the loss is not a unitary evolution (the trace is taken over the vacuum modes of the environment). Thus, during the evolution, the operator Ot=O(η(t),μ(t),κ(t)) keeps its form while the scalar functions satisfy the system of differential equations:
{η˙(t)=χμ*(t)λη(t)μ˙(t)=χη*(t)λμ(t)κ(t)=χ0t(η*(s)μ*(s)+η(s)μ(s))ds+κ(0).
The solution is straightforward:
(η(t)μ*(t))=exp((λχχλ)t)(η0μ0*),(μ(t)η*(t))=exp((λχχλ)t)(μ0η0*),κ(t)=(η0μ0*)·(0te(λχχλ)2sds)·(μ0η0*)+κ0.
Given the expression of the propagated Ot operator, one can evaluate the quadrature statistics of an evolved state. Let us calculate the following probability p(xθa,yξb)=|xθa,yξbxθa,yξb| on an evolved state. Using
|xθaxθa|=12πdζeiζ(x^θax)dζ
x^θa=12(aeiθ+aeiθ),
the projector on the quadrature eigenstates can be expressed as
|xθa,yξbxθa,yξb|=dζdγ(2π)2eiζxiγy×eζ2+γ24ei(ζeiθ2a+γeiξ2b)ei(ζeiθ2a+γeiξ2b)O0,
where the nontrivial part has the form of the operator O in Eq. (B2) with η0=ζeiγ2, μ0=γeiξ2 and κ0=ζ2+γ24. Resolving the time evolution of Ot using Eq. (B5), one obtains that, for the final probability (at time t),
p(xθa,yξb)=dζdγ(2π)2eiζxiγyOt.
For a coherent input states (seeds), the mean value Ot=α,β|Ot|α,β=eκ(t)ei(α*η(t)+β*μ(t))ei(αη*(t)+βμ*(t)) is particularly simple.

A direct calculation using the formulas above gives

p(xθa,yξb)=dζdγ(2π)2ei(ζγ)·(xZ(α,β)yΓ(α,γ))eζ2+γ24exp[18(ζγ)T·((1+e2t(λ+χ)λ+χ1+e2t(χλ)λχ)ei(θ+ξ)(1+e2t(χλ)λχ1+e2t(λ+χ)λ+χ)ei(θ+ξ)(1+e2t(χλ)λχ1+e2t(λ+χ)λ+χ)(1+e2t(λ+χ)λ+χ1+e2t(χλ)λχ))·(ζγ)].
The Fourier transform yields a Gaussian joint probability,
p(xθa,yξb)=rr+4πe14(xZ(α,β)yΓ(α,β))·M·(xZ(α,β)yΓ(α,β)),
where r+ and r are the two eigenvalues of the matrix M given by
r=(λ+χe2t(λ+χ)4(λ+χ))1,
r+=(λχe2t(χλ)4(λχ))1.
Accordingly, the joint probability p(xθa,yξb) decomposes in a product of two Gaussians with variance Δ=2/r in the squeezed direction (decreasing with time) and Δ+=2/r+ in the antisqueezed direction (increasing with time). Let us comment on their asymptotic values for t (limit of high finesse) for the two different regimes:
  • Below threshold λ>χ,
    Δ+λ2(λχ)Δλ2(λ+χ),
    both variances saturate at constant values.
  • Above threshold λ<χ,
    Δ+χ2(χλ)e2t(χλ)Δλ2(λ+χ),
    while the variance in the antisqueezed direction increases exponentially with time (finesse), the squeezed width cannot be decreased below a constant 1211+χ/λ set by the quality factor of the amplification process χλ.

Funding

Austrian Science Fund (FWF) (J3462, P24273-N16); European Research Council (ERC) (ERC MEC); National Swiss Science Foundation (SNSF) (P2GEP2-151964, PP00P2-150579).

Acknowledgment

We thank M. Mitchell for sharing several discussions with us; one of them initiated this work. We also warmly thank W. Dür, J. Laurat, and M. Skotiniotis for many discussions.

REFERENCES AND NOTES

1. A. J. Leggett, “Macroscopic quantum systems and the quantum theory of measurement,” Prog. Theor. Phys. Suppl. 69, 80–100 (1980). [CrossRef]  

2. A. J. Leggett, “Testing the limits of quantum mechanics: motivation, state of play, prospects,” J. Phys. Condens. Matter 14, R415–R451 (2002). [CrossRef]  

3. W. Dür, C. Simon, and J. I. Cirac, “Effective size of certain macroscopic quantum superpositions,” Phys. Rev. Lett. 89, 210402 (2002). [CrossRef]  

4. A. Shimizu and T. Miyadera, “Stability of quantum states of finite macroscopic systems against classical noises, perturbations from environments, and local measurements,” Phys. Rev. Lett. 89, 270403 (2002). [CrossRef]  

5. G. Björk and P. G. L. Mana, “A size criterion for macroscopic superposition states,” J. Opt. B 6, 429–436 (2004). [CrossRef]  

6. A. Shimizu and T. Morimae, “Detection of macroscopic entanglement by correlation of local observables,” Phys. Rev. Lett. 95, 090401 (2005). [CrossRef]  

7. E. G. Cavalcanti and M. D. Reid, “Signatures for generalized macroscopic superpositions,” Phys. Rev. Lett. 97, 170405 (2006). [CrossRef]  

8. J. I. Korsbakken, K. B. Whaley, J. Dubois, and J. I. Cirac, “Measurement-based measure of the size of macroscopic quantum superpositions,” Phys. Rev. A 75, 042106 (2007). [CrossRef]  

9. F. Marquardt, B. Abel, and J. Von Delft, “Measuring the size of a quantum superposition of many-body states,” Phys. Rev. A 78, 012109 (2008). [CrossRef]  

10. C.-W. Lee and H. Jeong, “Quantification of macroscopic quantum superpositions within phase space,” Phys. Rev. Lett. 106, 220401 (2011). [CrossRef]  

11. F. Fröwis and W. Dür, “Measures of macroscopicity for quantum spin systems,” New J. Phys. 14, 093039 (2012). [CrossRef]  

12. S. Nimmrichter and K. Hornberger, “Macroscopicity of mechanical quantum superposition states,” Phys. Rev. Lett. 110, 160403 (2013). [CrossRef]  

13. P. Sekatski, N. Sangouard, and N. Gisin, “Size of quantum superpositions as measured with classical detectors,” Phys. Rev. A 89, 012116 (2014). [CrossRef]  

14. A. Laghaout, J. S. Neergaard-Nielsen, and U. L. Andersen, “Assessments of macroscopicity for quantum optical states,” Opt. Commun. 337, 96–101 (2015).

15. H. Jeong, M. Kang, and H. Kwon, “Characterizations and quantifications of macroscopic quantumness and its implementations using optical fields,” Opt. Commun. 337, 12–21 (2015).

16. T. Farrow and V. Vedral, “Classification of macroscopic quantum effects,” Opt. Commun. 337, 22–26 (2015).

17. F. Fröwis, N. Sangouard, and N. Gisin, “Linking measures for macroscopic quantum states via photon–spin mapping,” Opt. Commun. 337, 2–11 (2015).

18. E. Schrödinger, “Die gegenwärtige Situation in der Quantenmechanik,” Naturwissenschaften 23, 807–812 (1935). [CrossRef]  

19. L.-M. Duan, G. Giedke, J. I. Cirac, and P. Zoller, “Inseparability criterion for continuous variable systems,” Phys. Rev. Lett. 84, 2722–2725 (2000). [CrossRef]  

20. R. Simon, “Peres-Horodecki separability criterion for continuous variable systems,” Phys. Rev. Lett. 84, 2726–2729 (2000). [CrossRef]  

21. Y. Zhang, H. Wang, X. Li, J. Jing, C. Xie, and K. Peng, “Experimental generation of bright two-mode quadrature squeezed light from a narrow-band nondegenerate optical parametric amplifier,” Phys. Rev. A 62, 023813 (2000). [CrossRef]  

22. C. Silberhorn, P. K. Lam, O. Weiß, F. König, N. Korolkova, and G. Leuchs, “Generation of continuous variable Einstein-Podolsky-Rosen entanglement via the Kerr nonlinearity in an optical fiber,” Phys. Rev. Lett. 86, 4267–4270 (2001). [CrossRef]  

23. W. P. Bowen, R. Schnabel, P. K. Lam, and T. C. Ralph, “Experimental investigation of criteria for continuous variable entanglement,” Phys. Rev. Lett. 90, 043601 (2003). [CrossRef]  

24. A. S. Villar, L. S. Cruz, K. N. Cassemiro, M. Martinelli, and P. Nussenzveig, “Generation of bright two-color continuous variable entanglement,” Phys. Rev. Lett. 95, 243603 (2005). [CrossRef]  

25. J. Jing, S. Feng, R. Bloomer, and O. Pfister, “Experimental continuous-variable entanglement from a phase-difference-locked optical parametric oscillator,” Phys. Rev. A 74, 041804 (2006). [CrossRef]  

26. X. Su, A. Tan, X. Jia, Q. Pan, C. Xie, and K. Peng, “Experimental demonstration of quantum entanglement between frequency-nondegenerate optical twin beams,” Opt. Lett. 31, 1133–1135 (2006). [CrossRef]  

27. G. Keller, V. D’Auria, N. Treps, T. Coudreau, J. Laurat, and C. Fabre, “Experimental demonstration of frequency-degenerate bright EPR beams with a self-phase-locked OPO,” Opt. Express 16, 9351–9356 (2008). [CrossRef]  

28. Y. Wang, H. Shen, X. Jin, X. Su, C. Xie, and K. Peng, “Experimental generation of 6 dB continuous variable entanglement from a nondegenerate optical parametric amplifier,” Opt. Express 18, 6149–6155 (2010). [CrossRef]  

29. T. Eberle, V. Händchen, and R. Schnabel, “Stable control of 10 dB two-mode squeezed vacuum states of light,” Opt. Express 21, 11546–11553 (2013). [CrossRef]  

30. A. Ourjoumtsev, R. Tualle-Brouri, J. Laurat, and P. Grangier, “Generating optical Schrödinger kittens for quantum information processing,” Science 312, 83–86 (2006). [CrossRef]  

31. J. S. Neergaard-Nielsen, B. M. Nielsen, C. Hettich, K. Mølmer, and E. S. Polzik, “Generation of a superposition of odd photon number states for quantum information networks,” Phys. Rev. Lett. 97, 083604 (2006). [CrossRef]  

32. K. Wakui, H. Takahashi, A. Furusawa, and M. Sasaki, “Photon subtracted squeezed states generated with periodically poled KTiOPO4,” Opt. Express 15, 3568–3574 (2007). [CrossRef]  

33. N. Bruno, A. Martin, P. Sekatski, N. Sangouard, R. T. Thew, and N. Gisin, “Displacement of entanglement back and forth between the micro and macro domains,” Nat. Phys. 9, 545–548 (2013). [CrossRef]  

34. A. I. Lvovsky, R. Ghobadi, A. Chandra, A. S. Prasad, and C. Simon, “Observation of micro-macro entanglement of light,” Nat. Phys. 9, 541–544 (2013). [CrossRef]  

35. O. Morin, K. Huang, J. Liu, H. Le Jeannic, C. Fabre, and J. Laurat, “Remote creation of hybrid entanglement between particle-like and wave-like optical qubits,” Nat. Photonics 8, 570–574 (2014). [CrossRef]  

36. H. Jeong, A. Zavatta, M. Kang, S.-W. Lee, L. S. Costanzo, S. Grandi, T. C. Ralph, and M. Bellini, “Generation of hybrid entanglement of light,” Nat. Photonics 8, 564–569 (2014). [CrossRef]  

37. P. Sekatski, N. Sangouard, M. Stobińska, F. Bussières, M. Afzelius, and N. Gisin, “Proposal for exploring macroscopic entanglement with a single photon and coherent states,” Phys. Rev. A 86, 060301 (2012). [CrossRef]  

38. S. L. Braunstein and C. M. Caves, “Statistical distance and the geometry of quantum states,” Phys. Rev. Lett. 72, 3439–3443 (1994). [CrossRef]  

39. E. G. Cavalcanti and M. D. Reid, “Criteria for generalized macroscopic and mesoscopic quantum coherence,” Phys. Rev. A 77, 062108 (2008). [CrossRef]  

40. F. Fröwis and W. Dür, “Are cloned quantum states macroscopic?” Phys. Rev. Lett. 109, 170401 (2012). [CrossRef]  

41. While superpositions of coherent states with opposite phases have a negative Wigner representation, here we show that the size of two-mode vacuum squeezed states is basically the same. This shows that states with a positive Wigner representation can be macroscopically quantum.

42. To illustrate this, consider a state ρϵ in Eq. (6) with f1(x,x¯)=f2(x,x¯)=1 for |xx¯|<ϵ and zero otherwise. The peculiarity of this state is that f satisfies f=f=0 in such a way that the violation of the Duan–Simon criterion by ρϵ is independent of ϵ unless it is strictly equal to zero. Therefore, ρϵ with ϵ approaching zero is an example of a state that can give an arbitrarily high violation of the Duan criteria, while being arbitrarily close to a separable state.

43. G. Tóth and D. Petz, “Extremal properties of the variance and the quantum Fisher information,” Phys. Rev. A 87, 032324 (2013). [CrossRef]  

44. S. Yu, “Quantum fisher information as the convex roof of variance,” arXiv:1302.5311 (2013).

45. F. Fröwis and N. Gisin, “Tighter quantum uncertainty relations follow from a general probabilistic bound,” Phys. Rev. A 92, 012102 (2015).

46. T. J. Volkoff and K. B. Whaley, “Measurement- and comparison-based sizes of Schrödinger cat states of light,” Phys. Rev. A 89, 012122 (2014). [CrossRef]  

47. S. Raeisi, P. Sekatski, and C. Simon, “Coarse graining makes it hard to see micro-macro entanglement,” Phys. Rev. Lett. 107, 250401 (2011). [CrossRef]  

48. F. Fröwis, M. Van den Nest, and W. Dür, “Certifiability criterion for large-scale quantum systems,” New J. Phys. 15, 113011 (2013). [CrossRef]  

49. T. Wang, R. Ghobadi, S. Raeisi, and C. Simon, “Precision requirements for observing macroscopic quantum effects,” Phys. Rev. A 88, 062114 (2013). [CrossRef]  

50. P. Sekatski, N. Gisin, and N. Sangouard, “How difficult is it to prove the quantumness of macroscropic states?” Phys. Rev. Lett. 113, 090403 (2014). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (1)

Fig. 1.
Fig. 1. Bounds on the effective size N eff (blue squares) of two-mode squeezed states obtained from experimental data reported in [2129] using the inequality in Eq. (10). The red triangles indicate the minimal photon number N necessary for a cat state | α + | α to have the same effective size according to Eq. (9). For example, the state reported in [21] has a size N eff 1.2 for which one needs at least a cat state with N 0.2 for the same size.

Equations (36)

Equations on this page are rendered with MathJax. Learn more.

| ψ tms = ( 1 tanh 2 g ) 1 2 e tanh g a 1 a 2 | 00 .
V ψ tms ( X ¯ 1 φ X ¯ 2 ϕ ) = cosh 2 g sinh 2 g cos ( φ + ϕ ) .
V sep ( | a | X ¯ 1 ϕ + 1 a X ¯ 2 Φ ) + V sep ( | a | X ¯ 1 ϕ 1 a X ¯ 2 Φ ) > a 2 [ X ¯ 1 ϕ , X ¯ 1 ϕ ] + 1 a 2 [ X ¯ 2 Φ , X ¯ 2 Φ ] 2 for ϕ ϕ = Φ Φ = π 2 ,
V ψ tms ( X ¯ 1 0 X ¯ 2 0 ) + V ψ tms ( X ¯ 1 π / 2 + X ¯ 2 π / 2 ) = 2 e 2 g .
P σ guess = 1 2 + 1 π arctan ( sinh 2 g 1 + 2 σ 2 cosh 2 g ) .
σ max = 1 + N ( 1 2 + N ) cotan 2 ( 1 2 P σ guess ) 2 + 2 N .
σ max = | α | 2 ( erf 1 ( P σ guess ) ) 2 1 2 .
ρ = p ( x 1 , x 2 ) p * ( x ¯ 1 , x ¯ 2 ) f ( x 1 , x ¯ 1 , x 2 , x ¯ 2 ) | x 1 , x 2 x ¯ 1 , x ¯ 2 | d x 1 d x 2 d x ¯ 1 d x ¯ 2 ,
V ( X ¯ 1 π 2 + X ¯ 2 π 2 ) = V ( X ¯ 1 π 2 + X ¯ 2 π 2 ) | ideal ( x 1 x ¯ 1 + x 2 x ¯ 2 ) 2 f ,
( x 1 x ¯ 1 + x 2 x ¯ 2 ) 2 f V ( X ¯ 1 π 2 + X ¯ 2 π 2 ) .
x C = 1 V ψ tms ( X ¯ 1 π 2 + X ¯ 2 π 2 ) .
N eff ( ρ ) = 1 2 n max θ F ρ ( X θ ) ,
N eff ( ρ ) 1 V ρ ( X ¯ 1 π / 2 + X ¯ 2 π / 2 ) .
ρ = F ( x 1 , x 2 , x ¯ 1 , x ¯ 2 ) | x 1 , x 2 x ¯ 1 , x ¯ 2 | d x⃗ .
F ( x 1 , x 2 , x ¯ 1 , x ¯ 2 ) = p ( x 1 , x 2 ) p * ( x ¯ 1 , x ¯ 2 ) f ( x 1 , x 2 , x ¯ 1 , x ¯ 2 ) ,
p 1 = F ( x 1 , x 2 , x ¯ 1 , x ¯ 2 ) p e i p ( x 1 x ¯ 1 ) 2 π δ ( x 2 x ¯ 2 ) d p d x⃗ , p 1 2 = F ( x 1 , x 2 , x ¯ 1 , x ¯ 2 ) p 2 e i p ( x 1 x ¯ 1 ) 2 π δ ( x 2 x ¯ 2 ) d p d x⃗ , p 1 p 2 = F ( x 1 , x 2 , x ¯ 1 , x ¯ 2 ) p 1 p 2 e i p 1 ( x 1 x ¯ 1 ) 2 π × e i p 2 ( x 2 x ¯ 2 ) 2 π d p 1 d p 2 d x⃗ .
p 1 = ( i x 1 x ¯ 1 ) F ( x 1 , x 2 , x ¯ 1 , x ¯ 2 ) | x 1 = x ¯ 1 d x 1 d x 2 , p 1 2 = ( i x 1 x ¯ 1 ) 2 F ( x 1 , x 2 , x ¯ 1 , x ¯ 2 ) | x 1 = x ¯ 1 d x 1 d x 2 , p 1 p 2 = ( i x 2 x ¯ 2 ) ( i x 1 x ¯ 1 ) , F ( x 1 , x 2 , x ¯ 1 , x ¯ 2 ) | x 1 = x ¯ 1 , x 2 = x ¯ 2 d x 1 d x 2 .
p 1 = p 1 f 1 , p 1 2 = p 1 2 f 1 x 1 x ¯ 1 2 f , p 1 2 = p 2 2 f 1 x 2 x ¯ 2 2 f , p 1 p 2 = p 1 p 2 f 1 x 1 x ¯ 1 x 2 x ¯ 2 f ,
D [ f ] = | p ( x 1 , x 2 ) | 2 D [ f ] ( x 1 , x 2 ) d x 1 d x 2 .
V ( p 1 + p 2 ) = V ( p 1 + p 2 ) f 1 ( x 1 x ¯ 1 + x 2 x ¯ 2 ) 2 f .
( x 1 x ¯ 1 + x 2 x ¯ 2 ) 2 f V ( p 1 + p 2 ) .
H A = i χ ( a b a b ) ,
O ( η , μ , κ ) = e κ e i ( η a + μ b ) e i ( η * a + μ * b )
tr loss U d t O ( η , μ , κ ) U d t = O ( η + ( μ * χ η λ ) d t , μ + ( η * χ μ λ ) d t , κ ( μ * η * + μ η ) χ d t ) ,
{ η ˙ ( t ) = χ μ * ( t ) λ η ( t ) μ ˙ ( t ) = χ η * ( t ) λ μ ( t ) κ ( t ) = χ 0 t ( η * ( s ) μ * ( s ) + η ( s ) μ ( s ) ) d s + κ ( 0 ) .
( η ( t ) μ * ( t ) ) = exp ( ( λ χ χ λ ) t ) ( η 0 μ 0 * ) , ( μ ( t ) η * ( t ) ) = exp ( ( λ χ χ λ ) t ) ( μ 0 η 0 * ) , κ ( t ) = ( η 0 μ 0 * ) · ( 0 t e ( λ χ χ λ ) 2 s d s ) · ( μ 0 η 0 * ) + κ 0 .
| x θ a x θ a | = 1 2 π d ζ e i ζ ( x ^ θ a x ) d ζ
x ^ θ a = 1 2 ( a e i θ + a e i θ ) ,
| x θ a , y ξ b x θ a , y ξ b | = d ζ d γ ( 2 π ) 2 e i ζ x i γ y × e ζ 2 + γ 2 4 e i ( ζ e i θ 2 a + γ e i ξ 2 b ) e i ( ζ e i θ 2 a + γ e i ξ 2 b ) O 0 ,
p ( x θ a , y ξ b ) = d ζ d γ ( 2 π ) 2 e i ζ x i γ y O t .
p ( x θ a , y ξ b ) = d ζ d γ ( 2 π ) 2 e i ( ζ γ ) · ( x Z ( α , β ) y Γ ( α , γ ) ) e ζ 2 + γ 2 4 exp [ 1 8 ( ζ γ ) T · ( ( 1 + e 2 t ( λ + χ ) λ + χ 1 + e 2 t ( χ λ ) λ χ ) e i ( θ + ξ ) ( 1 + e 2 t ( χ λ ) λ χ 1 + e 2 t ( λ + χ ) λ + χ ) e i ( θ + ξ ) ( 1 + e 2 t ( χ λ ) λ χ 1 + e 2 t ( λ + χ ) λ + χ ) ( 1 + e 2 t ( λ + χ ) λ + χ 1 + e 2 t ( χ λ ) λ χ ) ) · ( ζ γ ) ] .
p ( x θ a , y ξ b ) = r r + 4 π e 1 4 ( x Z ( α , β ) y Γ ( α , β ) ) · M · ( x Z ( α , β ) y Γ ( α , β ) ) ,
r = ( λ + χ e 2 t ( λ + χ ) 4 ( λ + χ ) ) 1 ,
r + = ( λ χ e 2 t ( χ λ ) 4 ( λ χ ) ) 1 .
Δ + λ 2 ( λ χ ) Δ λ 2 ( λ + χ ) ,
Δ + χ 2 ( χ λ ) e 2 t ( χ λ ) Δ λ 2 ( λ + χ ) ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.