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

Generation of tripartite entanglement from cascaded four-wave mixing processes

Open Access Open Access

Abstract

We investigate the possibility of an experimentally feasible cascaded four-wave mixing (FWM) system [Phys. Rev. Lett. 113, 023602 (2014)] to generate tripartite entanglement. We verify that genuine tripartite entanglement is present in this system by calculating the covariances of three output beams and then considering the violations of the inequalities of the three-mode entanglement criteria, such as two-condition criterion, single-condition criterion, optimal single-condition criterion and the positivity under partial transposition (PPT) criterion. We also consider the possibilities of the bipartite entanglement of any pair of the three output beams using the Duan-Giedke-Cirac-Zoller criterion and PPT criterion. We find that the tripartite entanglement and the bipartite entanglement for the two pairs are present in the whole gain region. The entanglement characteristics under different entanglement criteria are also considered. Our results pave the way for the realization and application of multipartite entanglement based on the cascaded FWM processes.

© 2016 Optical Society of America

1. Introduction

Multipartite entanglement is important not only for fundamental tests of quantum effects [1, 2] but also for their numerous possible applications in quantum technologies [3]. A number of different schemes for the generation of multipartite entangled states have been already theoretically proposed and experimentally implemented. The most well established technique consists of mixing several independently generated single mode squeezed states in a linear optical interferometer made of beamsplitters and propagation over well-controlled distances [4–7]. Recent progress towards scalable quantum optical networks includes the experimental generation of ultra-large-scale continuous variable (CV) cluster state multiplexed in both the time domain [8] and the frequency domain [9–11]. Meanwhile, there are many different kinds of criteria for detecting multipartite entanglement [12–16]. Among them, positivity under partial transposition (PPT) criterion is a necessary and sufficient criterion for Gaussian state under certain conditions while most of them are the sufficient criteria.

Four-wave mixing (FWM) is an effective χ(3) nonlinear technique to generate twin quantum correlated beams and bipartite entangled beams [17] in the CV domain. Recently, it has been reported that by cascading two FWM processes, tunable delay of Einstein–Podolsky–Rosen entangled states [18], low-noise amplification of an entangled states [19], SU(1,1) interferometer [20, 21], and quantum mutual information [22] have been experimentally realized.

Very recently, our group has theoretically proposed and experimentally demonstrated a cascaded FWM processes to produce multiple quantum correlated beams in hot atomic vapor [23]. In order to experimentally realize the tripartite entanglement in the cascaded FWM system, it is necessary to theoretically investigate its system parameter dependence. There are also some other cascaded systems studied by other groups for generating multipartite entanglement. According to the different ways of cascading, we could classify the related works into three categories, including cascading pump, cascading seed&pump and cascading seed. For the case of cascading pump, in the work [24], by successively employing the pump reflected by the first optical parametric oscillator (OPO) to pump the second one, the cascading pump act as an entanglement distributor among two OPOs, the existence of pentapartite entanglement (two signal beams+two idler beams+one pump beam) in this system is demonstrated using the PPT criterion. For the case of cascading seed&pump, one of signal and idler beams produced by the first nondegenerate optical parametric amplifiers (NOPO) can be used for the pump light of the second NOPO. The three-color entanglement among signal and idler beams produced by the second NOPO and the retained another beam of the first NOPO is theoretically proposed [25] and experimentally demonstrated [26]. The similar works [27, 28] use the cascaded parametric down-conversion processes to generate and observe the photon triplet and the three-photon energy-time entanglement, respectively. For the case of cascading seed, different from the above related works, in our work, one of signal and idler beams produced by the first FWM is used for the seed light of the second FWM. The tripartite entanglement among one signal and two idler beams produced by the two cascaded FWM processes is theoretically demonstrated using the three-mode entanglement criteria. In addition, our system has the following advantages that ensures its scalability. Firstly, as the number of quantum modes increases, so does the total degree of quantum correlations. Secondly, this method has advantage of being spatially separated due to the spatial multimode nature of the FWM system. Finally, this method is phase insensitive without the need of complicated phase locking technique. Due to the above advantages, this cascaded proposal is a promising candidate for tripartite entanglement generation and has many potential applications in the fields of quantum networks. Inspired by this, in this paper we theoretically analyze the dependence of genuine tripartite entanglement on the power gains of the system in the cascaded FWM processes using the three-mode entanglement criteria [13, 15, 16], such as two-condition criterion, single-condition criterion, optimal single-condition criterion and the PPT criterion.

This article is organized as follows. In Sec. 2 we describe the physical system of the cascaded FWM processes under consideration. In Sec. 3 we use the Duan-Giedke-Cirac-Zoller (DGCZ) criterion [12], and PPT criterion [13, 16, 29, 30] to characterize bipartite entanglement potentially existed in this system. In Sec. 4 we use the two-condition criterion, the single-condition criterion, the optimal single-condition criterion [15], and PPT criterion [13, 16, 31] to characterize the tripartite entanglement. In Sec. 5 we give a brief summary of this paper.

2. Cascaded FWM processes

Firstly, we give the optical quadrature definitions in our analysis, as the criteria will depend on these quadratures. For the three modes described by the bosonic annihilation oparators âk involved in our system, where k=1, 2, 3, we define quadrature operators for each mode as

X^k=a^k+a^k,Y^k=i(a^ka^k),
such that [k, Ŷk]=2i, k and Ŷk are respectively the amplitude and phase quadratures.

The energy level diagram of a single FWM process is shown in Fig. 1(b), where two pump photons can convert to one signal photon and one idler photon, or vice versa. In the cell1 of Fig. 1(a), an intense pump beam and a much weaker signal beam are crossed in the center of the Rb vapor cell with a slight angle. Then the signal beam is amplified and a new beam called idler beam is generated on the other side of the pump beam at the same time. The signal beam and idler beam have different frequencies. We then construct two similar cascaded FWM processes based on the single FWM process in Fig. 1(a) and Fig. 1(c). We take the signal beam from the first FWM process (cell1) as the seed for the second FWM process (cell2) in Fig. 1(a) and the idler beam from the first FWM process (cell1) as the seed for the second FWM process (cell2) in Fig. 1(c). â1, â2 and â3 are three newly-generated beams in the output stage of the cascaded processes. For convenience, here we will focus on the genuine tripartite entanglement (â1, â2 and â3) existed in the Fig. 1(a). A similar analysis can be made for the tripartite entanglement in Fig. 1(c). The input-output relation of the cascaded FWM processes in Fig. 1(a) can be written as

a^1=G1a^ν1+g1a^0,a^2=G1G2a^0+G2g1a^ν1+g2a^ν2,a^3=G2a^ν2+G1g2a^0+g1g2a^ν1,
where â0 is coherent input signal, âν1 and âν2 are vacuum inputs. Gj (j=1, 2) is the power gain in the FWM process and Gjgj=1. Eq. (2) can be rewritten in terms of the quadrature operators using Eq. (1) as follows
(X^1X^2X^3)=(G1g10G2g1G1G2g2g1g2G1g2G2)(X^ν1X^0X^ν2),
and
(Y^1Y^2Y^3)=(G1g10G2g1G1G2g2g1g2G1g2G2)(Y^ν1Y^0Y^ν2).
From the above expressions for the quadrature operators, it is possible to get the variances and covariances of the amplitude and phase quadratures necessary to calculate the entanglement criteria. V(X^m)=X^m2X^m2 denotes the amplitude quadrature variance. Whereas for the covariance, we use the notation Vmn = (〈mn〉 + 〈nm〉)/2 − 〈m〉〈n〉 and for the case where m=n, the covariance, denoted Vmn, reduces to the usual variance, V(m). In fact, the variances are given by the following gain-dependent moments (we assume that â0 is vacuum input in the following calculations because homodyne detection technique needs three output beams â1, â2 and â3 are all at vacuum level.)
X^12=Y^12=2G1,X^22=Y^22=2G1G21,X^32=Y^32=2G1G22G1+1,
since the expectation values of quadratures are all zero. Here we have used the fact that 〈mn〉 = 〈ŶmŶn〉 = δmn (m, n = 0, ν1 and ν2). A similar approach can be used to calculate the covariances which are equivalent to the gain-dependent moments 〈mn〉 and 〈ŶmŶn〉. The covariances are given by the following gain-dependent moments
X^1X^2=Y^1Y^2=2G1G2(G11),X^1X^3=Y^1Y^3=2G1(G11)(G21),X^2X^3=Y^2Y^3=2G1G2(G21),
we will use the above expressions to analyze the following entanglement criteria, because the three beams output from the cascaded FWM system are Gaussian states, and the entanglement properties of Gaussian states can be completely characterized by their covariance matrix.

 figure: Fig. 1

Fig. 1 Proposed schemes for generating tripartite entanglement based on the cascaded FWM processes. (a) Cascaded FWM processes, â0 is coherent input signal, âν1 and âν2 are vacuum inputs, G1 and G2 are the power gains of cell1 and cell2, respectively. â1, â2 and â3 are three output beams. (b) Energy level diagram of 85Rb D1 line for the single FWM process. 0.8 GHz is one-photon detuning, 4 MHz is two-photon detuning. (c) Another cascaded FWM processes.

Download Full Size | PDF

3. Bipartite entanglement

3.1. DGCZ criterion

In this section, we will focus on the bipartite entanglement potentially existed in the system. Firstly, we use the DGCZ criterion [12] to analyze the bipartite entanglement potentially existed in the cascaded FWM processes. Using our quadrature definitions in Eq. (1), the inequalities of the criteria are written as below

D12=V(X^1X^2)+V(Y^1+Y^2)4,D13=V(X^1X^3)+V(Y^1+Y^3)4,D23=V(X^2X^3)+V(Y^2+Y^3)4.
We calculate the dependence of D12, D13 and D23 on the gains G1 and G2 as follows
D12=4[G1(G2+1)2G1G2(G11)1],D13=4G1G2,D23=4[2G1G22G1G2(G21)G1].

The violation of the first inequality in Eq. (7) can be used to demonstrate the bipartite entanglement between â1 and â2, so also can the second (third) inequality in Eq. (7) be used to demonstrate the bipartite entanglement between â1 (â2) and â3 (â3). The region plot of Eq. (7) is shown in Fig. 2(a). The entanglement region of â1 and â2 (D12<4) is the region B in Fig. 2(a). As shown in Fig. 2(b), if we set G2 ≈1, the cascaded FWM processes reduce to the single FWM process, i. e., the first FWM process (cell1). Under this condition, â1 and â2 is a simple bipartite entangled beams case. As G1 is getting larger, D12 is limited by smaller G2. This is because only beam â2 experiences the quantum amplification from the second FWM process (cell2), which leads to their noise unbalance, thus the performance of the entanglement between beams â1 and â2 is very sensitive to the G2. Similarly, The entanglement region of â2 and â3 (D23<4) is the region A in Fig. 2(a). As shown in Fig. 2(c), if we set G1 ≈1, the cascaded FWM processes reduce to the single FWM process, i. e., the second FWM process (cell2). In this case, â2 and â3 is also a simple bipartite entangled beams case. It should be noted that the value of D13 is always more than or equal to 4 for any G1, G2>1.

 figure: Fig. 2

Fig. 2 (a) Region plot of Eq. (7). Region A is a region where D23 <4, region B is a region where D12 <4, region C is a region where D12, D13 and D23 in Eq. (7) are all falling above 4; (b) The contour plot of D12; and (c) The contour plot of D23.

Download Full Size | PDF

3.2. PPT criterion

Secondly, compared with the above sufficient criterion (DGCZ criterion), here a necessary and sufficient criterion, i. e., PPT criterion, can also be used to characterize the bipartite entanglement potentially existed in the system. For example, the entanglement between â1 and â2 can be completely characterized by their covariance matrix (CM12) which can be written as

CM12=[X^120X^1X^200Y^120Y^1Y^2X^1X^20X^2200Y^1Y^20Y^22].
The bipartite entanglement between â1 and â2 is absent if and only if both of the symplectic eigenvalues of the partially transposed (PT) CM12 are greater than or equal to 1 [13,16,29,30]. Following this idea, the entanglement between â1 and â2 can be characterized by the smaller symplectic eigenvalue B1. If the smaller symplectic eigenvalue B1 is smaller than 1, the bipartite entanglement exists between â1 and â2. Substituting Eq. (5) and Eq. (6) in Eq. (9), we can get the detailed expression for B1 as shown below
B1=1+G1+G1G2G124G1G2+2G12G2+G12G22,
the contour plot of B1 is shown in Fig. 3(a). It can be proved that the value of the whole region for any G1, G2 >1 is smaller than 1 meaning that â1 is entangled with â2. Similarly, the smaller symplectic eigenvalue B2 of PT CM23 of â2 and â3 is given by
B2=G1+2G1G2(12G1)+(G1+2G1G2)2,
its contour plot is shown in Fig. 3(b), its value is smaller than 1 for any G1, G2 >1 meaning that â2 is entangled with â3.

 figure: Fig. 3

Fig. 3 The contour plot of (a) B1; and (b) B2; and (c) B3.

Download Full Size | PDF

In addition, the smaller symplectic eigenvalue B3 of PT CM13 of â1 and â3 is given by

B3=1+2G1G1G2+(12G1+G1G2)2(12G1G2),
its value is always larger than 1 for any G1, G2 >1 as shown in Fig. 3(c) meaning that â1 is not entangled with â3. This is due to the fact that â1 and â3 never interact with each other directly.

4. Tripartite entanglement

4.1. Two-condition criterion

In this section, the tripartite entanglement will be investigated using different criteria, i. e., two-condition criterion, single-condition criterion, the optimal single-condition criterion and the PPT criterion. Firstly, we will consider the genuine tripartite entanglement between the three output modes from the cascaded FWM system using the two-condition criterion [15]. Using our quadrature definitions in Eq. (1), the two-condition criterion gives a set of inequalities

V12=V(X^1X^2)+V(Y^1+Y^2+O3Y^3)4,V23=V(X^2X^3)+V(O1Y^1+Y^2+Y^3)4,
where Oi (i=1, 3) are arbitrary real numbers. The violation of the two inequalities in Eq. (13) is sufficient to demonstrate genuine tripartite entanglement. We will investigate the optimization of the two-condition criterion using the freedom allowed in the choice of the Oi. The optimal expressions of Oi ( Oiopt) can be obtained by the direct differentiation for the left-hand side of Eq. (13) with respect to the Oi, then we get
O1opt=(Y^1Y^2+Y^1Y^3)Y^12=2G1G2(G11)2G1(G11)(G21)2G11,O3opt=(Y^1Y^3+Y^2Y^3)Y^32=2G1G2(G21)2G1(G11)(G21)2G1G22G1+1,

The contour plot of Eq. (13) is shown in Fig. 4. The dependence of V12 on G1 and G2 is shown in Fig. 4(a). The region of V12<4 is enlarged compared to the one of D12 <4 in Fig. 2(b) when we consider the phase quadrature of â3 (Ŷ3). The variance of Ŷ1+Ŷ2+Ŷ3 becomes smaller than the one of Ŷ1+Ŷ2 which claims that â3 has correlation with â1+â2. The dependence of V23 on G1 and G2 is shown in Fig. 4(b), the region of V23<4 is enlarged compared to the one of D23<4 in Fig. 2(c) when we consider the phase quadrature of â1 (Ŷ1). The variance of Ŷ1+Ŷ2+Ŷ3 becomes smaller than the one of Ŷ2+Ŷ3 which claims that â1 has correlation with â2+â3.

 figure: Fig. 4

Fig. 4 (a) The contour plot of V12; (b) The contour plot of V23; and (c) The light blue region is the region of V23 <4, the light orange region is the region of V12<4.

Download Full Size | PDF

The overlapped region of V12 <4 and V23 <4 in Fig. 4(c) means that the genuine tripartite entanglement is present in this system.

4.2. Single-condition criterion

Secondly, for simplicity, it is also possible to develop a single-condition criterion to demonstrate genuine tripartite entanglement using the combined quadrature variances [15]. If one of the following formulas in Eq. (15) is less than 2, we could claim that there exists genuine tripartite entanglement in this system

V213=V[X^212(X^1+X^3)]+V[Y^2+12(Y^1+Y^3)],V312=V[X^312(X^1+X^2)]+V[Y^3+12(Y^1+Y^2)],V123=V[X^112(X^2+X^3)]+V[Y^1+12(Y^2+Y^3)].

The dependence of V213 on the gains G1 and G2 can be written as follows

V213=2[G21G22]2+2[(G11)G2G12(G11)(G21)2]2+2[G1G2G112G1(G21)2]2.

The contour plot of V213 is shown in Fig. 5, the value of most of the region in Fig. 5 is less than 2. We can use this entanglement criterion to directly detect the three-mode entanglement, and it also clearly shows the existence of tripartite entanglement in this cascaded system. But the values of V312 and V123 are always bigger than 2 due to the fact that â2 is correlated with both â1 and â3, whereas â1 and â3 are not correlated with each other directly. In fact, they are from two independent FWM processes in serial connection.

 figure: Fig. 5

Fig. 5 The contour plot for V213 in Eq. (16).

Download Full Size | PDF

4.3. Optimal single-condition criterion

Inspired by above results in Sec. 4.2, we should pay more attention to the optimization of V213 ( V213opt) through introducing different factors [15] to cancel the impact of noise unbalance between â1 and â3. In this way, we can express the V213opt as follows

V213opt=V[X^2F1X^1F3X^3]+V[Y^2+F1Y^1+F3Y^3],
where F1 and F3 are arbitrary real numbers whose optimal expressions ( F1opt and F3opt) can be obtained by differentiating either 2F11F33 or Ŷ2 + F1Ŷ1 + F3Ŷ3 with respect to F1 and F3, then we get
F1opt=Y^1Y^2Y^32Y^1Y^3Y^2Y^3Y^1Y^32Y^12Y^32,F3opt=Y^2Y^3Y^12Y^1Y^3Y^1Y^2Y^1Y^32Y^12Y^32.

The contour plot of V213opt is shown in Fig. 6(a), it depends on G1 and G2 symmetrically which is similar to the squeezing degree of the triple beams (1/[2G1G2 − 1]) [23]. Thus in the following discussions we will analyze the tripartite entanglement using the optimal single-condition criterion in the condition G=G1=G2. In this case, Eq. (17) can be reduced to

V213opt=22G21,
and Eq. (18) can be reduced to
F1opt=2GG12G21,F3opt=2GG(G1)2G21,

 figure: Fig. 6

Fig. 6 (a) The contour plot of V213opt in Eq. (17); (b) The dependence of V213opt (trace A); 2(|F12|+|1F32|) (trace B); 2(|F32|+|1F12|) (trace C); and 2(1+|F12+F32|) (trace D) on gain G; and (c) The dependence of R on gain G. 1 for trace A, R=V213opt/2(|F12|+|1F32|) for trace B.

Download Full Size | PDF

F1opt and F3opt in Eq. (20) tend to 1/G and 1 respectively when G becomes infinite, resulting that the coefficients of the quadratures from the modes â2 and â3 become equal. This makes sense because when G becomes infinite, the powers of the modes â2 and â3 are equal and it is unnecessary to introduce different factors to cancel the noise unbalance between the modes â2 and â3.

To confirm the existence of the tripartite entanglement in this system, we should consider whether Eq. (17) is smaller than its boundaries according to Ref. [15]. The boundaries of V213opt are 2(|F12|+|1F32|) (â2 and â3 may be entangled or not, but â1 is not entangled with the rest), 2(|F32|+|1F12|) (â1 and â2 may be entangled or not, but â3 is not entangled with the rest), and 2(1+|F12+F32|) (â1 and â3 may be entangled or not, but â2 is not entangled with the rest). To answer this question, we plot the dependence of V213opt (Trace A), 2(1+|F12+F32|) (Trace B), 2(|F32|+|1F12|) (Trace C), and 2(1+|F12+F32|) (Trace D) on G in Fig. 6(b). As shown in Fig. 6(b), 2(|F12|+|1F32|) (Trace B) decreases with the increasing of G, 2(|F32|+|1F12|) (Trace C) and 2(1+|F12+F32|) (trace D) increase with the increasing of G.

As shown in Fig. 6(b), 2(|F12|+|1F32|) (Trace B) is the smallest one of all the boundaries which means that tripartite entanglement is present if V213opt (Trace A) is smaller than 2(|F12|+|1F32|) (Trace B). Thus we define the quantity R=V213opt/2(|F12|+|1F32|) to characterize the dependence of tripartite entanglement on G and R < 1 means that tripartite entanglement is present. As shown in Fig. 6(c), the tripartite entanglement is present in the whole gain region. Therefore, the optimal single-condition criterion can be used to characterize the tripartite entanglement in this cascaded system.

 figure: Fig. 7

Fig. 7 The contour plot of (a) T1; (b) T2; and (c) T3.

Download Full Size | PDF

4.4. PPT criterion

Finally, the PPT criterion used to characterize the bipartite entanglement can also be applied to the tripartite entanglement because the PPT criterion is also a necessary and sufficient criterion for all 1× N decompositions of Gaussian states, where 1+N is the total number of the entangled modes [13, 16, 31]. For the tripartite entanglement condition, the three possible 1× 2 partitions have to be tested. All the partitions of the three-mode state are inseparable when the smallest symplectic eigenvalue for each of the three PT covariance matrices is smaller than 1, i. e., genuine tripartite entanglement is present in this system. Generally, the covariance martix CM123 for the three modes can be written as

CM123=[X^120X^1X^20X^1X^300Y^120Y^1Y^20Y^1Y^3X^1X^20X^220X^2X^300Y^1Y^20Y^220Y^2Y^3X^1X^30X^2X^30X^3200Y^1Y^30Y^2Y^30Y^32].
When the PT operation is applied to beam â1, the entanglement between â1 and the rest beams (â2 and â3) can be characterized by the smallest symplectic eigenvalue T1. Substituting Eq. (5) and Eq. (6) in Eq. (21), we can get the detailed expression of T1 which is not listed here due to its complexity. As shown in Fig. 7(a), the value of T1 in the whole gain region is smaller than 1 for any G1, G2 >1 meaning that beam â1 is entangled with the rest beams (â2 and â3). It should be noted that the behavior of T1 is independent on the gain G2. This is due to the fact that the entanglement between beam â1 and the rest beams (â2 and â3) is only decided by the first FWM process (cell1). Similarly, when the PT operation is applied to the beam â2 (â3), the smallest symplectic eigenvalue is T2 (T3). The contour plots of T2 and T3 are shown in Fig. 7(b) and Fig. 7(c) respectively. It can be clearly seen that the values of T2 and T3 are both smaller than 1 for any G1, G2 >1 in the whole gain region, showing the presence of the entanglement for these two partitions. Therefore, the three partitions â1−(â2, â3), â2−(â1, â3), and â3−(â1, â2) are all inseparable and the genuine tripartite entanglement between the three beams â1, â2 and â3 is present for the whole gain region with any G1, G2 >1.

5. Conclusion

In summary, we have theoretically predicted that the cascaded FWM processes is a simple system for tripartite entanglement generation. We have used the DGCZ criterion, and the PPT criterion to characterize the bipartite entanglement potentially existed in the cascaded FWM processes. Then we have analyzed the tripartite entanglement using two-condition criterion and gave the region in which the tripartite entanglement is present. Single-condition criterion can also be used to search for a more simpler measurement scheme. We have proposed an optimal single-condition criterion to characterize the tripartite entanglement in this system. Under this optimal single condition criterion, the tripartite entanglement is present in the whole gain gain region. More importantly, the PPT criterion as a necessary and sufficient criterion for Gaussian state under certain conditions has also been used to demonstrate the genuine tripartite entanglement. Under the PPT criterion, the genuine tripartite entanglement is also present in the whole gain region. Our cascaded system for generating triple quantum entangled beams is scalable, simple and phase insensitive, and it can be extended to generate multiple quantum entangled beams via cascading more FWM processes.

The major experimental limitation concerning the maximum number of entangled beams comes from the increasing of the number of the atomic vapor cells. This can be solved by integrating all the pumps into a single cell and crossing them with the probe beam one by one.

Funding

This work was supported by the National Natural Science Foundation of China (91436211, 11374104, 10974057,11547141,11604297); the SRFDP (20130076110011); the Program for Professor of Special Appointment (Eastern Scholar) at Shanghai Institutions of Higher Learning; the Program for New Century Excellent Talents in University (NCET-10-0383); the Shu Guang project supported by Shanghai Municipal Education Commission and Shanghai Education Development Foundation (11SG26); the Shanghai Pujiang Program (09PJ1404400); the Scientific Research Foundation of the Returned Overseas Chinese Scholars, State Education Ministry; and Program of State Key Laboratory of Advanced Optical Communication Systems and Networks (2016GZKF0JT003).

Acknowledgments

J. J. would like to thank A. M. Marino for useful discussions.

References and links

1. A. Einstein, B. Podolsky, and N. Rosen, “Can Quantum-Mechanical Description of Physical Reality Be Considered Complete,” Phys. Rev. 47, 777–780 (1935). [CrossRef]  

2. D. M. Greenberger, M. A. Horne, and A. Zeilinger, “Multiparticle Interferometry and the Superposition Principle,” Phys. Today 46, 22–29 (1993). [CrossRef]  

3. C. Weedbrook, S. Pirandola, R. G. Patrón, N. J. Cerf, T. C. Ralph, J. H. Shapiro, and S. Lloyd, “Gaussian quantum information,” Rev. Mod. Phys. 84, 621–669 (2012). [CrossRef]  

4. J. Jing, J. Zhang, Y. Yan, F. Zhao, C. Xie, and K. Peng, “Experimental Demonstration of Tripartite Entanglement and Controlled Dense Coding for Continuous Variables,” Phys. Rev. Lett. 90, 167903 (2003). [CrossRef]   [PubMed]  

5. T. Aoki, N. Takei, H. Yonezawa, K. Wakui, T. Hiraoka, A. Furusawa, and Peter v. Loock, “Experimental Creation of a Fully Inseparable Tripartite Continuous-Variable State,” Phys. Rev. Lett. 91, 080404 (2003). [CrossRef]   [PubMed]  

6. H. Yonezawa, T. Aoki, and A. Furusawa, “Demonstration of a quantum teleportation network for continuous variables,” Nature 431, 430–433 (2004). [CrossRef]   [PubMed]  

7. Peter v. Loock and S. L. Braunstein, “Multipartite Entanglement for Continuous Variables: A Quantum Teleportation Network,” Phys. Rev. Lett. 84, 3482–3485 (2000). [CrossRef]   [PubMed]  

8. S. Yokoyama, R. Ukai, S. C. Armstrong, C. Sornphiphatphong, T. Kaji, S. Suzuki, J. I. Yoshikawa, H. Yonezawa, N. C. Menicucci, and A. Furusawa, “Ultra-large-scale continuous-variable cluster states multiplexed in the time domain,” Nature Photon. 7, 982–986 (2013). [CrossRef]  

9. J. Roslund, R. M. de Araújo, S. F. Jiang, C. Fabre, and N. Treps, “Wavelength-multiplexed quantum networks with ultrafast frequency combs,” Nature Photon. 8, 109–112 (2014). [CrossRef]  

10. S. Gerke, J. Sperling, W. Vogel, Y. Cai, J. Roslund, N. Treps, and C. Fabre, “Full Multipartite Entanglement of Frequency-Comb Gaussian States,” Phys. Rev. Lett. 114, 050501 (2015). [CrossRef]   [PubMed]  

11. M. Chen, N. C. Menicucci, and O. Pfister, “Experimental Realization of Multipartite Entanglement of 60 Modes of a Quantum Optical Frequency Comb,” Phys. Rev. Lett. 112, 120505 (2014). [CrossRef]   [PubMed]  

12. 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]   [PubMed]  

13. R. Simon, “Peres-Horodecki Separability Criterion for Continuous Variable Systems,” Phys. Rev. Lett. 84, 2726–2729 (2000). [CrossRef]   [PubMed]  

14. V. Giovannetti, S. Mancini, D. Vitali, and P. Tombesi, “Characterizing the entanglement of bipartite quantum systems,” Phys. Rev. A 67, 022320 (2003). [CrossRef]  

15. Peter v. Loock and A. Furusawa, “Detecting genuine multipartite continuous-variable entanglement,” Phys. Rev. A 67, 052315 (2003). [CrossRef]  

16. R. F. Werner and M. M. Wolf, “Bound Entangled Gaussian States,” Phys. Rev. Lett. 86, 3658–3661 (2001). [CrossRef]   [PubMed]  

17. V. Boyer, A. M. Marino, R. C. Pooser, and P. D. Lett, “Entangled Images from Four-Wave Mixing,” Science 321, 544–547 (2008). [CrossRef]   [PubMed]  

18. A. M. Marino, R. C. Pooser, V. Boyer, and P. D. Lett, “Tunable delay of Einstein–Podolsky–Rosen entanglement,” Nature 457, 859–862 (2009). [CrossRef]   [PubMed]  

19. R. C. Pooser, A. M. Marino, V. Boyer, K. M. Jones, and P. D. Lett, “Low-Noise Amplification of a Continuous-Variable Quantum State,” Phys. Rev. Lett. 103, 010501 (2009). [CrossRef]   [PubMed]  

20. J. Jing, C. Liu, Z. Zhou, Z. Y. Ou, and W. Zhang, “Realization of a nonlinear interferometer with parametric amplifiers,” Appl. Phys. Lett. 99, 011110 (2011). [CrossRef]  

21. F. Hudelist, J. Kong, C. Liu, J. Jing, Z. Y. Ou, and W. Zhang, “Quantum metrology with parametric amplifier-based photon correlation interferometers,” Nature Commun. 5, 3049 (2014).

22. J. B. Clark, R. T. Glasser, Q. Glorieux, U. Vogl, T. Li, K. M. Jones, and P. D. Lett, “Quantum mutual information of an entangled state propagating through a fast-light medium,” Nature Photon. 8, 515–519 (2014). [CrossRef]  

23. Z. Qin, L. Cao, H. Wang, A. M. Marino, W. Zhang, and J. Jing, “Experimental Generation of Multiple Quantum Correlated Beams from Hot Rubidium Vapor,” Phys. Rev. Lett. 113, 023602 (2014). [CrossRef]   [PubMed]  

24. K. N. Cassemiro and A. S. Villar, “Scalable continuous-variable entanglement of light beams produced by optical parametric oscillators,” Phys. Rev. A 77, 022311 (2008). [CrossRef]  

25. A. Tan, C. Xie, and K. Peng, “Bright three-color entangled state produced by cascaded optical parametric oscillators,” Phys. Rev. A 85, 013819 (2012). [CrossRef]  

26. X. Jia, Z. Yan, Z Duan, X. Su, H. Wang, C. Xie, and K. Peng, “Experimental Realization of Three-Color Entanglement at Optical Fiber Communication and Atomic Storage Wavelengths,” Phys. Rev. Lett. 109, 253604 (2012). [CrossRef]  

27. H. Hübel, D. R. Hamel, A. Fedrizzi, S. Ramelow, K. J. Resch, and T. Jennewein, “Direct generation of photon triplets using cascaded photon-pair sources,” Nature 466, 601–603 (2010). [CrossRef]   [PubMed]  

28. L. K. Shalm, D. R. Hamel, Z. Yan, C. Simon, K. J. Resch, and T. Jennewein, “Three-photon energy-time entanglement,” Nature Phys. 9, 19–22 (2013). [CrossRef]  

29. F. A. S. Barbosa, A. J. de Faria, A. S. Coelho, K. N. Cassemiro, A. S. Villar, P. Nussenzveig, and M. Martinelli, “Disentanglement in bipartite continuous-variable systems,” Phys. Rev. A 84, 052330 (2011). [CrossRef]  

30. F. A. S. Barbosa, A. S. Coelho, A. J. de Faria, K. N. Cassemiro, A. S. Villar, P. Nussenzveig, and M. Martinelli, “Robustness of bipartite Gaussian entangled beams propagating in lossy channels,” Nature Photon. 4, 858–861 (2010). [CrossRef]  

31. A. S. Coelho, F. A. S. Barbosa, K. N. Cassemiro, A. S. Villar, M. Martinelli, and P. Nussenzveig, “Three-Color Entanglement,” Science 326, 823–826 (2009). [CrossRef]   [PubMed]  

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 (7)

Fig. 1
Fig. 1 Proposed schemes for generating tripartite entanglement based on the cascaded FWM processes. (a) Cascaded FWM processes, â0 is coherent input signal, âν1 and âν2 are vacuum inputs, G1 and G2 are the power gains of cell1 and cell2, respectively. â1, â2 and â3 are three output beams. (b) Energy level diagram of 85Rb D1 line for the single FWM process. 0.8 GHz is one-photon detuning, 4 MHz is two-photon detuning. (c) Another cascaded FWM processes.
Fig. 2
Fig. 2 (a) Region plot of Eq. (7). Region A is a region where D23 <4, region B is a region where D12 <4, region C is a region where D12, D13 and D23 in Eq. (7) are all falling above 4; (b) The contour plot of D12; and (c) The contour plot of D23.
Fig. 3
Fig. 3 The contour plot of (a) B1; and (b) B2; and (c) B3.
Fig. 4
Fig. 4 (a) The contour plot of V12; (b) The contour plot of V23; and (c) The light blue region is the region of V23 <4, the light orange region is the region of V12<4.
Fig. 5
Fig. 5 The contour plot for V213 in Eq. (16).
Fig. 6
Fig. 6 (a) The contour plot of V 213 opt in Eq. (17); (b) The dependence of V 213 opt (trace A); 2 ( | F 1 2 | + | 1 F 3 2 | ) (trace B); 2 ( | F 3 2 | + | 1 F 1 2 | ) (trace C); and 2 ( 1 + | F 1 2 + F 3 2 | ) (trace D) on gain G; and (c) The dependence of R on gain G. 1 for trace A, R = V 213 opt / 2 ( | F 1 2 | + | 1 F 3 2 | ) for trace B.
Fig. 7
Fig. 7 The contour plot of (a) T1; (b) T2; and (c) T3.

Equations (21)

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

X ^ k = a ^ k + a ^ k , Y ^ k = i ( a ^ k a ^ k ) ,
a ^ 1 = G 1 a ^ ν 1 + g 1 a ^ 0 , a ^ 2 = G 1 G 2 a ^ 0 + G 2 g 1 a ^ ν 1 + g 2 a ^ ν 2 , a ^ 3 = G 2 a ^ ν 2 + G 1 g 2 a ^ 0 + g 1 g 2 a ^ ν 1 ,
( X ^ 1 X ^ 2 X ^ 3 ) = ( G 1 g 1 0 G 2 g 1 G 1 G 2 g 2 g 1 g 2 G 1 g 2 G 2 ) ( X ^ ν 1 X ^ 0 X ^ ν 2 ) ,
( Y ^ 1 Y ^ 2 Y ^ 3 ) = ( G 1 g 1 0 G 2 g 1 G 1 G 2 g 2 g 1 g 2 G 1 g 2 G 2 ) ( Y ^ ν 1 Y ^ 0 Y ^ ν 2 ) .
X ^ 1 2 = Y ^ 1 2 = 2 G 1 , X ^ 2 2 = Y ^ 2 2 = 2 G 1 G 2 1 , X ^ 3 2 = Y ^ 3 2 = 2 G 1 G 2 2 G 1 + 1 ,
X ^ 1 X ^ 2 = Y ^ 1 Y ^ 2 = 2 G 1 G 2 ( G 1 1 ) , X ^ 1 X ^ 3 = Y ^ 1 Y ^ 3 = 2 G 1 ( G 1 1 ) ( G 2 1 ) , X ^ 2 X ^ 3 = Y ^ 2 Y ^ 3 = 2 G 1 G 2 ( G 2 1 ) ,
D 12 = V ( X ^ 1 X ^ 2 ) + V ( Y ^ 1 + Y ^ 2 ) 4 , D 13 = V ( X ^ 1 X ^ 3 ) + V ( Y ^ 1 + Y ^ 3 ) 4 , D 23 = V ( X ^ 2 X ^ 3 ) + V ( Y ^ 2 + Y ^ 3 ) 4 .
D 12 = 4 [ G 1 ( G 2 + 1 ) 2 G 1 G 2 ( G 1 1 ) 1 ] , D 13 = 4 G 1 G 2 , D 23 = 4 [ 2 G 1 G 2 2 G 1 G 2 ( G 2 1 ) G 1 ] .
CM 12 = [ X ^ 1 2 0 X ^ 1 X ^ 2 0 0 Y ^ 1 2 0 Y ^ 1 Y ^ 2 X ^ 1 X ^ 2 0 X ^ 2 2 0 0 Y ^ 1 Y ^ 2 0 Y ^ 2 2 ] .
B 1 = 1 + G 1 + G 1 G 2 G 1 2 4 G 1 G 2 + 2 G 1 2 G 2 + G 1 2 G 2 2 ,
B 2 = G 1 + 2 G 1 G 2 ( 1 2 G 1 ) + ( G 1 + 2 G 1 G 2 ) 2 ,
B 3 = 1 + 2 G 1 G 1 G 2 + ( 1 2 G 1 + G 1 G 2 ) 2 ( 1 2 G 1 G 2 ) ,
V 12 = V ( X ^ 1 X ^ 2 ) + V ( Y ^ 1 + Y ^ 2 + O 3 Y ^ 3 ) 4 , V 23 = V ( X ^ 2 X ^ 3 ) + V ( O 1 Y ^ 1 + Y ^ 2 + Y ^ 3 ) 4 ,
O 1 opt = ( Y ^ 1 Y ^ 2 + Y ^ 1 Y ^ 3 ) Y ^ 1 2 = 2 G 1 G 2 ( G 1 1 ) 2 G 1 ( G 1 1 ) ( G 2 1 ) 2 G 1 1 , O 3 opt = ( Y ^ 1 Y ^ 3 + Y ^ 2 Y ^ 3 ) Y ^ 3 2 = 2 G 1 G 2 ( G 2 1 ) 2 G 1 ( G 1 1 ) ( G 2 1 ) 2 G 1 G 2 2 G 1 + 1 ,
V 213 = V [ X ^ 2 1 2 ( X ^ 1 + X ^ 3 ) ] + V [ Y ^ 2 + 1 2 ( Y ^ 1 + Y ^ 3 ) ] , V 312 = V [ X ^ 3 1 2 ( X ^ 1 + X ^ 2 ) ] + V [ Y ^ 3 + 1 2 ( Y ^ 1 + Y ^ 2 ) ] , V 123 = V [ X ^ 1 1 2 ( X ^ 2 + X ^ 3 ) ] + V [ Y ^ 1 + 1 2 ( Y ^ 2 + Y ^ 3 ) ] .
V 213 = 2 [ G 2 1 G 2 2 ] 2 + 2 [ ( G 1 1 ) G 2 G 1 2 ( G 1 1 ) ( G 2 1 ) 2 ] 2 + 2 [ G 1 G 2 G 1 1 2 G 1 ( G 2 1 ) 2 ] 2 .
V 213 opt = V [ X ^ 2 F 1 X ^ 1 F 3 X ^ 3 ] + V [ Y ^ 2 + F 1 Y ^ 1 + F 3 Y ^ 3 ] ,
F 1 opt = Y ^ 1 Y ^ 2 Y ^ 3 2 Y ^ 1 Y ^ 3 Y ^ 2 Y ^ 3 Y ^ 1 Y ^ 3 2 Y ^ 1 2 Y ^ 3 2 , F 3 opt = Y ^ 2 Y ^ 3 Y ^ 1 2 Y ^ 1 Y ^ 3 Y ^ 1 Y ^ 2 Y ^ 1 Y ^ 3 2 Y ^ 1 2 Y ^ 3 2 .
V 213 opt = 2 2 G 2 1 ,
F 1 opt = 2 G G 1 2 G 2 1 , F 3 opt = 2 G G ( G 1 ) 2 G 2 1 ,
CM 123 = [ X ^ 1 2 0 X ^ 1 X ^ 2 0 X ^ 1 X ^ 3 0 0 Y ^ 1 2 0 Y ^ 1 Y ^ 2 0 Y ^ 1 Y ^ 3 X ^ 1 X ^ 2 0 X ^ 2 2 0 X ^ 2 X ^ 3 0 0 Y ^ 1 Y ^ 2 0 Y ^ 2 2 0 Y ^ 2 Y ^ 3 X ^ 1 X ^ 3 0 X ^ 2 X ^ 3 0 X ^ 3 2 0 0 Y ^ 1 Y ^ 3 0 Y ^ 2 Y ^ 3 0 Y ^ 3 2 ] .
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.