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

Generation of multipartite continuous-variable entanglement via atomic spin wave: Heisenberg-Langevin approach

Open Access Open Access

Abstract

We conduct theoretical studies on the effects of various parameters on generation of multipartite continuous-variable entanglement via atomic spin wave induced by the strong coupling and probe fields in the Λ-type electromagnetically induced transparency configuration in a realistic atomic ensemble by using the Heisenberg-Langevin formalism. It is shown that the increase of the atomic density and/or Rabi frequencies of the scattering fields, as well as the decrease of the coherence decay rate of the lower doublet would strengthen the degree of multipartite entanglement. This provides a clear evidence that the creation of multicolor multipartite entangled narrow-band fields to any desired number with a long correlation time can be achieved conveniently by using atomic spin wave in an atomic ensemble with large optical depth, which may find interesting applications in quantum information processing and quantum networks.

© 2014 Optical Society of America

1. Introduction

Much attention has been paid to the generation of multipartite continuous-variable (CV) entanglement in recent years due to its wide applications, such as quantum computation, quantum communication, quantum cryptography, and quantum networks [13]. Apart from the routinely-used way of combining the entangled twin beams produced through spontaneous parametric down-conversion in nonlinear crystals to create entanglement among multiple beams by employing linear optical elements, i.e., polarizing beam splitters [4,5], an alternative promising avenue is the use of nonlinear processes [e.g., nondegenerate four-wave mixing (FWM) or Raman scattering] in an atomic ensemble to produce multi-entangled beams. These generated beams have the virtues of narrow bandwidth, long correlation time, and nondegenerate frequencies, which are quite necessary for connecting different physical systems at the nodes of quantum networks and quite suitable for quantum memory required in quantum communication [3]. Based on the seminal paper by Duan et al. [6], nondegenerate quantum correlated and entangled narrow-band photon pairs have been experimentally achieved with both hot atoms [79] and magneto-optical trap (MOT) [10,11] by using the electromagnetically induced transparency [1215] (EIT)-based double-Λ-type system. Recently, Kolchin [16] and Glorieux et al. [17] have theoretically studied EIT-based paired photon generation with the consideration of Langevin noise terms by using the Heisenberg-Langevin method. Through multiorder coherent Raman scattering processes, the generation of multipartite entanglement in a far-off-resonance medium with a prepared coherence has been investigated [18]. Very recently, we have proposed a convenient and flexible way to create multicolor multipartite CV entanglement to any desired number and quantum entangler via an atomic spin wave pre-established by strong coupling and probe fields in the Λ-type EIT configuration [19, 20]. However, in a realistic atomic ensemble, there will be various dephasing processes, such as finite interaction time between atoms and light, and collision-induced atomic coherence decay, etc, which would degrade the degree of entanglement and have not been taken into account in [19, 20].

In this paper, based on the experimental observation of generating multi-field correlations and anti-correlations via atomic spin coherence in 85Rb atomic system [21] and theoretical examination of generating multipartite CV entanglement and quantum entangler via atomic spin wave [19, 20], we discuss in detail the effects of the atomic density, Rabi frequencies of the scattering fields, and coherence decay rate of the lower doublet on the generation of nondegenerate multipartite CV entanglement via atomic spin wave in the Λ-type EIT configuration in an atomic ensemble. By using the Heisenberg-Langevin formalism, we demonstrate that increasing the atomic density and Rabi frequencies of the scattering fields, and decreasing the coherence decay rate of the lower doublet, would strengthen the degree of multipartite entanglement. This method provides a proof-of-principle demonstration of conveniently generating multicolor multipartite entangled narrow-band fields to any desired number via atomic spin wave in an atomic ensemble with high optical depth, which may have promising applications in quantum information processing and quantum networks.

2. Heisenberg-langevin equations

We consider a collection of atoms having the Λ-type configuration shown in Fig. 1(a), similar to the one studied in [1921], where the relevant energy levels and the applied and generated laser fields form a quintuple-Λ-type system. Levels |1, |2, and |3 correspond, respectively, to the ground-state hyperfine levels 5S1/2 (F = 3), 5S1/2 (F = 2), and the excited state 5P1/2 in D1 line of 85Rb atom with the ground-state hyperfine splitting of 3.036 GHz. We assume the strong probe field Ep (with frequency ωp and Rabi frequency Ωp) is tuned to resonance with the transition |2-|3, whereas the strong coupling field Ec (with frequency ωc and Rabi frequency Ωc) is tuned close to the resonant transition |1-|3 with a small finite detuning Δ = ωc-ω31 with ωij (i ≠ j) being the atomic resonant frequency between levels i and j; thus, the two-photon detuning of the coupling and probe fields from the Raman transition |2-|3-|1 is also equal to Δ. A third mixing field Em with frequency ωm, which can be generated from the coupling or probe field with an acousto-optic modulator, off-resonantly couples levels |2 (or |1) and |3 with the detuning Δ3(4) = ωm-ω32(31). As discussed in [1924], at high atomic density and high laser powers, two Stokes fields E1 & E3 (with frequencies ω1 & ω3) and two anti-Stokes fields E2 & E4 (with frequencies ω2 & ω4) can be simultaneously created through nondegenerate FWM processes with the coupling, probe, and mixing fields acting on both |1-|3 and |2-|3 transitions. Equivalently, the above simultaneously generated four FWM fields can be regarded as scattering the coupling, probe, and mixing fields off the atomic spin wave (S) pre-built by the strong coupling and probe fields in the Λ-type EIT configuration formed by levels |1, |2, and |3, where the induced spin wave acts as a frequency converter with frequency equal to the separation between the lower doublet [19, 25]. The equivalent configuration is shown in Fig. 1(b), which can be readily generalized to multi-Λ-type system by applying more laser fields tuned to the vicinity of the transitions |1-|3 and/or |2-|3 to mix with the induced atomic spin wave.

 figure: Fig. 1

Fig. 1 (a) The quintuple-Λ-type system of the D1 transitions in 85Rb atom coupled by the coupling (Ec), probe (Ep), and mixing (Em) fields based on the configuration in Refs [19, 21], where Ep, Ec, and Em fields all drive both |1|3 and |2|3 transitions, and the corresponding Stokes fields (E1 and E3), and anti-Stokes fields (E2 and E4) are generated through four FWM processes. (b) The equivalent configuration of (a) with the two lower states driven by the atomic spin wave S induced by the strong Ec and Ep fields in the Λ-type EIT configuration.

Download Full Size | PDF

With the Heisenberg-Langevin method, we first investigate the generation of quantum anti-correlated and entangled Stokes field E1 and anti-Stokes field E2 via atomic spin wave in a triple-Λ-type system by blocking the mixing field Em. We use the equivalent configuration shown in Fig. 1(b) to treat the generated bipartite entanglement. The detunings of the probe field Ep (with the Rabi frequency Ω2) and coupling field Ec (with the Rabi frequency Ω1) from the resonant transitions |1-|3 and |2-|3 are denoted as Δ2 = ωp-ω31 and Δ1 = ωc-ω32, respectively. In the present scheme, we assume that the generated Stokes and anti-Stokes fields are very weak as compared to the scattering fields, thus, the scattering fields can be treated classically, whereas the Stokes field E1 and anti-Stokes field E2 can be treated quantum mechanically, and we assume that the Rabi frequencies of the scattering fields Ω1 and Ω2 are far smaller than their frequency detunings Δ1 and Δ2, so that the coupling between different scattering fields can be neglected. Also, we assume that the coupling and probe fields for creating atomic spin wave are substantially strong, so that the atomic spin wave is strong enough to ensure that different scattering fields have negligible influence on it. Under this condition, the Heisenberg-Langevin equations for describing the evolution of the collective atomic operators σ12(z,t), σ13(z,t), and σ23(z,t) can be written as [16,17,26]

σ˙12(z,t)=(γ0+iΔ)σ12ig1a1(z,t)σ23++ig2a2+(z,t)σ13+iΩ1σ13iΩ2σ23++F12(z,t),
σ˙13(z,t)=[(γ1+γ2)/2i(Δ1Δ)]σ13+ig1a1(z,t)(σ11σ33)+iΩ1σ12+F13(z,t),
σ˙23(z,t)=[(γ1+γ2)/2i(Δ2+Δ)]σ23+ig2a2(z,t)(σ22σ33)+iΩ2σ12++F23(z,t),
where γ1 and γ2 are the population decay rates from level 3 to levels 1 and 2 and γ0 the coherence decay rate between levels 1 and 2, a1 and a2 are the quantum operators of the generated Stokes E1 and anti-Stokes E2 fields, g1(2)=μ13(23)ε1(2)/ is the atom-field coupling constant with μ13(23) as the dipole moment for the 1-3 (2-3) transition and ε1(2)=ω1(2)/2ε0V as the electric field of a single Stokes (anti-Stokes) photon with V as the interaction volume with length L and beam radius r, and Fij(z,t) are the collective atomic δ-correlated Langevin noise operators. In the similar analysis described in [16, 17, 26], under the assumption that the uniformly-distributed pencil-shaped atomic sample is optically thin in the transverse direction and the scattering fields propagate without depletion, the evolution of the annihilation a1 and creation a2+ operators for the Stokes and anti-Stokes fields can be described by the coupled propagation equations
(t+cz)a1(z,t)=ig1Nσ13,
(t+cz)a2+(z,t)=ig2Nσ23+,
where N is the total number of atoms in the quantum interaction volume. As done in [16, 17], we use the perturbation analysis to treat the interaction of the atoms with the fields. In the zeroth-order perturbation expansion, by semiclassically treating the interaction of the atoms with the strong coupling and probe fields in the Λ-type EIT configuration, we get the steady-state mean values of the atomic operators σ11(0), σ22(0), σ33(0), σ12(0), σ13(0) and σ23(0). By Fourier transforming Heisenberg-Langevin equations for the atomic operators σ12(z,t), σ13(z,t), and σ23(z,t) and substituting the zeroth-order solution into equation for the operator σ12(z,t), we can get the first-order solution σ12(1)(z,ω). Then substituting σ12(1)(z,ω) and the zeroth-order solution for the other operators into the equations for σ13(z,t) and σ23(z,t), the first-order solution σ13(1)(z,ω), and σ23(1)(z,ω) can be obtained, which are expressed as
σ13=1γ13+i(ωΔ1+Δ){[ig1(σ11(0)σ33(0))+g1Ω1σ23+(0)i(ω+Δ)+γ0]a1g2Ω1σ13(0)i(ω+Δ)+γ0a2++iΩ1(iΩ1σ13(0)iΩ2σ23+(0))i(ω+Δ)+γ0+iΩ1F12i(ω+Δ)+γ0+F13},
σ23+=1γ23+i(ω+Δ2+Δ){g1Ω2σ23+(0)i(ω+Δ)+γ0a1[ig2(σ22(0)σ33(0))g2Ω2*σ13(0)i(ω+Δ)+γ0]a2+iΩ2(iΩ1σ13(0)iΩ2σ23+(0))i(ω+Δ)+γ0iΩ2F12i(ω+Δ)+γ0+F23+}.
Note that in the above first-order analysis, as done in [16, 17], we only consider the steady-state mean values of the zeroth-order atomic operators (neglecting the noise operators) for the first-order solution. Substituting σ13(1) and σ23(1) into Fourier- transformed coupled propagation equations, the output of the annihilation operator a1(L,ω) and creation operator a2+(L,ω) for the Stokes and anti-Stokes fields with respect to the Fourier frequency ω can be obtained, which is a linear combination of the input of the operators a1(0,ω) and a2+(0,ω) and Langevin noise terms. In a similar way as that in [19], we use the criterion V=(Δu)2+(Δv)2<4 proposed in [27] to verify the entanglement of the generated Stokes and anti-Stokes fields, where u=x1+x2 and v=p1p2 with xj=(aj+aj+) and pj=i(ajaj+). In what follows, the relevant parameters are scaled with m and MHz, or m−1 and MHz−1, and set according to the experimental conditions in [20,21,28,29] as follows: atomic density n0=1016, r=1×10-4, L=0.02, γ1=γ2=3, γ0=0.1, ω12=3036, Ωp=Ωc=600, Δ2=Δ1Δ=3036, Δ=150, and Δ3=2900.

3. Results and discussions

Figure 2 shows the evolution of V as a function of the Fourier frequency ω with Δ = -ω (i.e., energy conservation is fulfilled). It can be seen that, in a moderate range (i.e., 0~840MHz) of the Fourier frequency ω (that is, with the two-photon detuning Δ in the range of −840~0 MHz), V becomes less than 4, which sufficiently demonstrates the generation of genuine bipartite entanglement. The similar range of the negative two-photon detuning for optimizing the generation of two-field entanglement with FWM has been examined in [17,28,29]. In the following, we set Δ = -ω = −150 MHz, and investigate the effects of the atomic density, Rabi frequencies of the scattering fields, and coherence decay rate of the lower doublet on the generation of multipartite CV entanglement.

 figure: Fig. 2

Fig. 2 The evolution of V as a function of the Fourier frequency ω with L=0.02, r=1×10-4, n0=1016, γ1=γ2=3, γ0=0.1, ω12=3036, Ωp=Ωc=600, Ω1=Ω2=120, Δ2=Δ1Δ=3036, and Δ=ω in corresponding units of m and MHz, or m−1 and MHz−1.

Download Full Size | PDF

The dependence of V on the atomic density is depicted in Fig. 3. It is obvious that, V, with the initial value of 4, decrease gradually with the increase of the atomic density, which indicates two entangled fields are produced. When n0 is about larger than 4×1015/m3, V nearly becomes equal to zero; this means two perfectly squeezed fields can be achieved. Similar evolution of V with respect to the Rabi frequency of the scattering fields is observed, as displayed in Fig. 4, where when Ω1 and Ω2 is about larger than 40 MHz, V nearly reaches zero. Thus, in order to generate high degree of bipartite entanglement, the atomic density and Rabi frequency of the scattering fields should be high enough.

 figure: Fig. 3

Fig. 3 The evolution of V as a function of the atomic density n0 with Δ = -ω = −150 MHz, and the other parameters are the same as those in Fig. 2.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 The evolution of V as a function of the Rabi frequency Ω = Ω1 = Ω2of the scattering field with Δ = -ω = −150 MHz, and the other parameters are the same as those in Fig. 2.

Download Full Size | PDF

As is well known, the coherence decay rate of the lower doublet plays a critical role in the generation of the bipartite entanglement in the Λ-type three-level system. This is clearly demonstrated in Fig. 5 by examining the dependence of V on the coherence decay rate γ0 of the lower doublet. When the coherence decay rate is very small, V is almost equal to zero, and two nearly perfectly squeezed fields can be created. The increase of the coherence decay rate would increase the value of V, that is, the degree of bipartite entanglement would be weakened. When the coherence decay rate grows high enough, V becomes equal to 4 and no entanglement can be produced. In addition, in this Λ-type atomic system, the correlation time of the generated bipartite entanglement is determined by the coherence decay time between the two lower states, which, in practice, can be long (ms or even s [30,31]), thereby having the virtue suitable for quantum memory required in quantum communication [69].

 figure: Fig. 5

Fig. 5 The evolution of V as a function of the coherence decay rate γ0 of the lower doublet with Δ = -ω = −150 MHz, and the other parameters are the same as those in Fig. 2.

Download Full Size | PDF

In [19], we did not take into account the case of the generated Stokes and anti-Stokes fields with equal coupling strength. In fact, the highest degree of bipartite entanglement is achieved with the two generated fields nearly having equal coupling strength. This is confirmed in Fig. 6 by varying the ratio the coupling coefficient g23 to g13. It is obvious that the larger or smaller g23 with respect to g13, the weaker the degree of bipartite entanglement, and when the two coupling coefficients g23 and g13 are nearly equal to each other, the minimum value of V (almost equal to zero) is obtained, i.e., the maximum bipartite entanglement is produced.

 figure: Fig. 6

Fig. 6 The evolution of V as a function of the ratio of the coupling coefficients g23/g13 with Δ = -ω = −150 MHz, and the other parameters are the same as those in Fig. 2.

Download Full Size | PDF

As discussed in [19], the above idea for generating bipartite entanglement via atomic spin wave can be easily extended to create multipartite entanglement to any desired order when more scattering fields are applied. With the Heisenberg-Langevin formalism, we demonstrate this concept by realizing tripartite entanglement through scattering an additional mixing field E3 off the atomic spin wave, as shown in Fig. 1(b). We consider the case of generating three fields E1, E2, and E3. Also, we assume that the Rabi frequency Ω3 of the mixing field is small as compared to its frequency detuning Δ3 = ωm-ω31, so that the coupling between different scattering fields can be neglected. Under this condition, the evolution of the operator σ12 should include the terms from the interaction of the mixing field Em and the correspondingly generated Stokes field E3 with the atomic medium, whereas the evolution of the annihilation operator a3 for the generated Stokes field E3 can be described by the same coupled propagation equation as the annihilation operator a1 for the Stokes field E1. In a similar way, by Fourier transforming the Heisenberg-Langevin equations and coupled propagation equations, we can get the output of the operators a1(L,ω), a3(L,ω), and a2+(L,ω) for the Stokes fields E1 and E3 and anti-Stokes field E2 with respect to the Fourier frequency ω after the interaction of the atoms and the fields. As done in [19], the tripartite entanglement of the generated fields E1, E2, and E3 can be demonstrated according to the criterion proposed by van Lock-Furusawa (VLF) [32,33] with inequalities:

V12=V(x1+x2)+V(p1-p2+g3p3)<4V13=V(x1-x3)+V(p1+g2p2+p3)<4V23=V(x2+x3)+V(g1p1+p2-p3)<4
where V(A)=<A2><A>2 and gi is an arbitrary real number. Following Ref [34], we set g1=(<p1p2><p1p3>)<p12>, g2=(<p1p2>+<p2p3>)<p22>, and g3=(<p1p3><p2p3>)<p32>. Satisfying any pair of these three inequalities is sufficient to demonstrate the creation of tripartite entanglement.

Figures 7(a)7(c) display the evolutions of the VLF correlations as a function of the Fourier frequency ω with Δ = -ω, Δ3 = 2900 MHz, and Ω1 = Ω3 = 80 MHz. It can be seen that, in a moderate range of the Fourier frequency ω (i.e., about 100~700 MHz), V12, V13, and V23 are all smaller than 4, and the minimal values of V12, V13, and V23 are, respectively, about 0.75, 2, and 0.75, which sufficiently demonstrates that the generated fields E1, E2, and E3 are CV entangled with each other. Beyond the ranges, the entanglement among the generated three fields E1, E2, and E3 would disappear. Further calculations show that the dependence of the tripartite entanglement on the atomic density, Rabi frequencies of the scattering fields, and coherence decay rate of the lower doublet exhibits the similar behaviors as those of the bipartite entanglement; that is, the increase of the atomic density and Rabi frequencies of the scattering fields and the decrease of the coherence decay rate of the lower doublet would strengthen the degree of tripartite entanglement. In the same way, we can demonstrate the generation of tripartite entanglement among E1, E2, and E4 fields as well. This clearly indicates that the above concept for producing bipartite and tripartite entanglement via atomic spin wave can be easily extended to generate multipartite entanglement to any desired number with more scattering fields, tuned to the vicinity of the transitions |1-|3 and/or |2-|3, mixing with the prebuilt atomic spin wave.

 figure: Fig. 7

Fig. 7 The evolutions of the VLF correlations V12 (a), V13 (b), and V23 (c) as a function of the Fourier frequency ω with Δ = -ω, Δ3 = 2900 MHz, and Ω1 = Ω3 = 80 MHz, and the other parameters are the same as those in Fig. 2.

Download Full Size | PDF

The generated multipartite quantum correlations, quantum anti-correlations, and entanglement can be intuitively understood in terms of the interaction between the laser fields and atomic medium. As seen in Fig. 1(a), all the Stokes fields and anti-Stokes fields are produced through FWM processes, where every Stokes (anti-Stokes) photon generation is obtained by absorbing one scattering and one coupling (probe) photons and emitting one probe (coupling) photon; equivalently, the generated Stokes (or anti-Stokes) field can be regarded as the result of frequency down- (up-) conversion process through mixing the scattering field with the atomic spin wave S prebuilt by the strong coupling and probe fields (as shown in Fig. 1(b)). Since the generation of a Stokes (or an anti-Stokes) photon is accompanied with the creation (or annihilation) of an atomic spin wave excitation, the up-converted frequency component (i.e., anti-Stokes field) is quantum anti-correlated with the down-converted frequency component (i.e., Stokes field), therefore strong multipartite entanglement can be achieved. In this sense, the prebuilt atomic spin wave acts as a quantum beam splitter, and in principle, by using the atomic spin wave excitation, multipartite entangled CV fields to any desired number can be created through applying more scattering fields, as long as the scattering fields are weak and the atomic spin wave is strong enough to ensure that different scattering fields have negligible influence on it, which can be realized by employing substantially strong probe and coupling fields.

4.Conclusion

In conclusion, by using the Heisenberg-Langevin method, we have theoretically studied the effects of the atomic density, Rabi frequencies of the scattering fields, and coherence decay rate of the lower doublet on the generation of nondegenerate multipartite CV entanglement. The theoretical results provide a proof-of-principle demonstration of generating nondegenerate entangled narrow-band multiple fields to any desired number with long correlation time in an atomic ensemble with high optical depth, and give valuable evidences for further studies on multipartite CV entanglement via atomic spin wave and subsequent applications in quantum information processing and quantum networks.

Acknowledgments

This work is supported by NBRPC (Nos. 2012CB921804 and 2011CBA00205), the National Natural Science Foundation of China (Nos. 11274225, 10974132, 50932003, and 11021403), and Innovation Program of Shanghai Municipal Education Commission (No. 10YZ10).

References and Links

1. D. Bouweester, A. Ekert, and A. Zeilinger, The Physics of Quantum Information (Springer, 2000).

2. M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information (Cambridge University, 2000).

3. H. J. Kimble, “The quantum internet,” Nature 453(7198), 1023–1030 (2008). [CrossRef]   [PubMed]  

4. E. Knill, R. Laflamme, and G. J. Milburn, “A scheme for efficient quantum computation with linear optics,” Nature 409(6816), 46–52 (2001). [CrossRef]   [PubMed]  

5. T. Aoki, N. Takei, H. Yonezawa, K. Wakui, T. Hiraoka, A. Furusawa, and P. van Loock, “Experimental Creation of a Fully Inseparable Tripartite Continuous-Variable State,” Phys. Rev. Lett. 91(8), 080404 (2003). [CrossRef]   [PubMed]  

6. L. M. Duan, M. D. Lukin, J. I. Cirac, and P. Zoller, “Long-distance quantum communication with atomic ensembles and linear optics,” Nature 414(6862), 413–418 (2001). [CrossRef]   [PubMed]  

7. C. H. van der Wal, M. D. Eisaman, A. André, R. L. Walsworth, D. F. Phillips, A. S. Zibrov, and M. D. Lukin, “Atomic Memory for Correlated Photon States,” Science 301(5630), 196–200 (2003). [CrossRef]   [PubMed]  

8. V. Boyer, A. M. Marino, R. C. Pooser, and P. D. Lett, “Entangled Images from Four-Wave Mixing,” Science 321(5888), 544–547 (2008). [CrossRef]   [PubMed]  

9. H. Wu and M. Xiao, “Bright correlated twin beams from an atomic ensemble in the optical cavity,” Phys. Rev. A 80(6), 063415 (2009). [CrossRef]  

10. A. Kuzmich, W. P. Bowen, A. D. Boozer, A. Boca, C. W. Chou, L. M. Duan, and H. J. Kimble, “Generation of nonclassical photon pairs for scalable quantum communication with atomic ensembles,” Nature 423(6941), 731–734 (2003). [CrossRef]   [PubMed]  

11. V. Balić, D. A. Braje, P. Kolchin, G. Y. Yin, and S. E. Harris, “Generation of Paired Photons with Controllable Waveforms,” Phys. Rev. Lett. 94(18), 183601 (2005). [CrossRef]   [PubMed]  

12. S. E. Harris, “Electromagnetically Induced Transparency,” Phys. Today 50(7), 36–42 (1997). [CrossRef]  

13. M. Xiao, Y. Li, S. Jin, and J. Gea-Banacloche, “Measurement of Dispersive Properties of Electromagnetically Induced Transparency in Rubidium Atoms,” Phys. Rev. Lett. 74(5), 666–669 (1995). [CrossRef]   [PubMed]  

14. M. Fleischhauer, A. Imamoglu, and J. P. Marangos, “Electromagnetically induced transparency: Optics in coherent media,” Rev. Mod. Phys. 77(2), 633–673 (2005). [CrossRef]  

15. E. Arimondo, “V Coherent Population Trapping in Laser Spectroscopy,” Prog. Opt. 35, 257–354 (1996). [CrossRef]  

16. P. Kolchin, “Electromagnetically-induced-transparency-based paired photon generation,” Phys. Rev. A 75(3), 033814 (2007). [CrossRef]  

17. Q. Glorieux, R. Dubessy, S. Guibal, L. Guidoni, J.-P. Likforman, T. Coudreau, and E. Arimondo, “Double-Λ microscopic model for entangled light generation by four-wave mixing,” Phys. Rev. A 82(3), 033819 (2010). [CrossRef]  

18. F. Le Kien, A. Patnaik, and K. Hakuta, “Multiorder coherent Raman scattering of a quantum probe field,” Phys. Rev. A 68(6), 063803 (2003). [CrossRef]  

19. X. H. Yang, Y. Y. Zhou, and M. Xiao, “Generation of multipartite continuous- variable entanglement via atomic spin wave,” Phys. Rev. A 85(5), 052307 (2012). [CrossRef]  

20. X. H. Yang, Y. Y. Zhou, and M. Xiao, “Entangler via electromagnetically induced transparency with an atomic ensemble,” Sci. Rep. 3, 3479–3483 (2013). [PubMed]  

21. X. H. Yang, J. T. Sheng, U. Khadka, and M. Xiao, “Generation of correlated and anticorrelated multiple fields via atomic spin coherence,” Phys. Rev. A 85(1), 013824 (2012). [CrossRef]  

22. V. Wong, R. S. Bennink, A. M. Marino, R. W. Boyd, C. R. Stroud Jr, and F. Narducci, “Influence of coherent Raman scattering on coherent population trapping in atomic sodium vapor,” Phys. Rev. A 70(5), 053811 (2004). [CrossRef]  

23. K. I. Harada, T. Kanbashi, M. Mitsunaga, and K. Motomura, “Competition between electromagnetically induced transparency and stimulated Raman scattering,” Phys. Rev. A 73(1), 013807 (2006). [CrossRef]  

24. G. S. Agarwal, T. N. Dey, and D. J. Gauthier, “Competition between electromagnetically induced transparency and Raman processes,” Phys. Rev. A 74(4), 043805 (2006). [CrossRef]  

25. A. J. Merriam, S. J. Sharpe, M. Shverdin, D. Manuszak, G. Y. Yin, and S. E. Harris, “Efficient Nonlinear Frequency Conversion in an All-Resonant Double- Lambda System,” Phys. Rev. Lett. 84(23), 5308–5311 (2000). [CrossRef]   [PubMed]  

26. M. Fleischhauer and T. Richter, “Pulse matching and correlation of phase fluctuations in Λ systems,” Phys. Rev. A 51(3), 2430–2442 (1995). [CrossRef]   [PubMed]  

27. L. M. Duan, G. Giedke, J. I. Cirac, and P. Zoller, “Inseparability Criterion for Continuous Variable Systems,” Phys. Rev. Lett. 84(12), 2722–2725 (2000). [CrossRef]   [PubMed]  

28. C. F. McCormick, V. Boyer, E. Arimondo, and P. D. Lett, “Strong relative intensity squeezing by four-wave mixing in rubidium vapor,” Opt. Lett. 32(2), 178–180 (2007). [CrossRef]   [PubMed]  

29. C. F. McCormick, A. M. Marino, V. Boyer, and P. D. Lett, “Strong low-frequency quantum correlations from a four-wave-mixing amplifier,” Phys. Rev. A 78(4), 043816 (2008). [CrossRef]  

30. R. Zhang, S. R. Garner, and L. V. Hau, “Creation of Long-Term Coherent Optical Memory via Controlled Nonlinear Interactions in Bose-Einstein Condensates,” Phys. Rev. Lett. 103(23), 233602 (2009). [CrossRef]   [PubMed]  

31. U. Schnorrberger, J. D. Thompson, S. Trotzky, R. Pugatch, N. Davidson, S. Kuhr, and I. Bloch, “Electromagnetically Induced Transparency and Light Storage in an Atomic Mott Insulator,” Phys. Rev. Lett. 103(3), 033003 (2009). [CrossRef]   [PubMed]  

32. P. van Loock and S. L. Braunstein, “Multipartite Entanglement for Continuous Variables: A Quantum Teleportation Network,” Phys. Rev. Lett. 84(15), 3482–3485 (2000).

33. P. van Loock and A. Furusawa, “Detecting genuine multipartite continuous- variable entanglement,” Phys. Rev. A 67(5), 052315 (2003). [CrossRef]  

34. M. K. Olsen and A. S. Bradley, “Asymmetric polychromatic tripartite entanglement from interlinked χ(2) parametric interactions,” Phys. Rev. A 74(6), 063809 (2006). [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 (7)

Fig. 1
Fig. 1 (a) The quintuple-Λ-type system of the D1 transitions in 85Rb atom coupled by the coupling (Ec), probe (Ep), and mixing (Em) fields based on the configuration in Refs [19, 21], where Ep, Ec, and Em fields all drive both |1|3 and |2|3 transitions, and the corresponding Stokes fields (E1 and E3), and anti-Stokes fields (E2 and E4) are generated through four FWM processes. (b) The equivalent configuration of (a) with the two lower states driven by the atomic spin wave S induced by the strong Ec and Ep fields in the Λ-type EIT configuration.
Fig. 2
Fig. 2 The evolution of V as a function of the Fourier frequency ω with L=0.02 , r=1× 10 -4 , n 0 =10 16 , γ 1 = γ 2 =3 , γ 0 =0.1 , ω 12 =3036 , Ω p = Ω c =600 , Ω 1 = Ω 2 =120 , Δ 2 = Δ 1 Δ=3036 , and Δ=ω in corresponding units of m and MHz, or m−1 and MHz−1.
Fig. 3
Fig. 3 The evolution of V as a function of the atomic density n0 with Δ = -ω = −150 MHz, and the other parameters are the same as those in Fig. 2.
Fig. 4
Fig. 4 The evolution of V as a function of the Rabi frequency Ω = Ω 1 = Ω 2 of the scattering field with Δ = -ω = −150 MHz, and the other parameters are the same as those in Fig. 2.
Fig. 5
Fig. 5 The evolution of V as a function of the coherence decay rate γ 0 of the lower doublet with Δ = -ω = −150 MHz, and the other parameters are the same as those in Fig. 2.
Fig. 6
Fig. 6 The evolution of V as a function of the ratio of the coupling coefficients g23/g13 with Δ = -ω = −150 MHz, and the other parameters are the same as those in Fig. 2.
Fig. 7
Fig. 7 The evolutions of the VLF correlations V12 (a), V13 (b), and V23 (c) as a function of the Fourier frequency ω with Δ = -ω, Δ3 = 2900 MHz, and Ω1 = Ω3 = 80 MHz, and the other parameters are the same as those in Fig. 2.

Equations (8)

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

σ ˙ 12 (z,t)=( γ 0 +iΔ) σ 12 i g 1 a 1 (z,t) σ 23 + +i g 2 a 2 + ( z,t ) σ 13 +i Ω 1 σ 13 i Ω 2 σ 23 + + F 12 ( z,t ),
σ ˙ 13 (z,t)=[( γ 1 + γ 2 )/2i( Δ 1 Δ)] σ 13 +i g 1 a 1 (z,t)( σ 11 σ 33 )+i Ω 1 σ 12 + F 13 (z,t),
σ ˙ 23 (z,t)=[( γ 1 + γ 2 )/2i( Δ 2 +Δ)] σ 23 +i g 2 a 2 (z,t)( σ 22 σ 33 )+i Ω 2 σ 12 + + F 23 (z,t),
( t +c z ) a 1 (z,t)=i g 1 N σ 13 ,
( t +c z ) a 2 + (z,t)=i g 2 N σ 23 + ,
σ 13 = 1 γ 13 +i(ω Δ 1 +Δ) {[i g 1 ( σ 11 (0) σ 33 (0) )+ g 1 Ω 1 σ 23 +(0) i(ω+Δ)+ γ 0 ] a 1 g 2 Ω 1 σ 13 (0) i(ω+Δ)+ γ 0 a 2 + + i Ω 1 (i Ω 1 σ 13 (0) i Ω 2 σ 23 +(0) ) i(ω+Δ)+ γ 0 + i Ω 1 F 12 i(ω+Δ)+ γ 0 + F 13 },
σ 23 + = 1 γ 23 +i(ω+ Δ 2 +Δ) { g 1 Ω 2 σ 23 +(0) i(ω+Δ)+ γ 0 a 1 [ i g 2 ( σ 22 (0) σ 33 (0) ) g 2 Ω 2 * σ 13 (0) i(ω+Δ)+ γ 0 ] a 2 + i Ω 2 (i Ω 1 σ 13 (0) i Ω 2 σ 23 +(0) ) i(ω+Δ)+ γ 0 i Ω 2 F 12 i(ω+Δ)+ γ 0 + F 23 + }.
V 12 =V( x 1 + x 2 )+V( p 1 - p 2 + g 3 p 3 )<4 V 13 =V( x 1 - x 3 )+V( p 1 + g 2 p 2 + p 3 )<4 V 23 =V( x 2 + x 3 )+V( g 1 p 1 + p 2 - p 3 )<4
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.