Tunable two-axis spin model and spin squeezing in two cavities
Yu Lixian1, Li Caifeng2, 3, Fan Jingtao2, Chen Gang2, 5, Zhang Tian-Cai4, 5, Jia Suotang2, 5
Department of Physics, Shaoxing University, Shaoxing 312000, China
State Key Laboratory of Quantum Optics and Quantum Optics Devices, Institute of Laser spectroscopy, Shanxi University, Taiyuan 030006, China
Institute of Theoretical Physics, Shanxi University, Taiyuan 030006, China
State Key Laboratory of Quantum Optics and Quantum Optics Devices, Institute of Opto-Electronics, Shanxi University, Taiyuan 030006, China
Collaborative Innovation Center of Extreme Optics, Shanxi University, Taiyuan 030006, China

 

† Corresponding author. E-mail: chengang971@163.com

Project supported by the National Natural Science Foundation of China (Grant Nos. 11422433, 11447028, 61227902, 11434007, and 61275211), the Natural Science Foundation of Zhejiang Province, China (Grant No. LY13A040001), and the Scientific Research Foundation of the Education Department of Zhejiang Province, China (Grant No. Y201122352).

Abstract
Abstract

Multi-mode cavities have now attracted much attention both experimentally and theoretically. In this paper, inspired by recent experiments of cavity-assisted Raman transitions, we realize a two-axis spin Hamiltonian in two cavities. This realized Hamiltonian has a distinct property that all parameters can be tuned independently. For proper parameters, the well-studied one- and two-axis twisting Hamiltonians are recovered, and the scaling of N−1 of the maximal squeezing factor can occur naturally. On the other hand, in the two-axis twisting Hamiltonian, spin squeezing is usually reduced when increasing the atomic resonant frequency ω0. Surprisingly, we find that by combining with the dimensionless parameter χ (> −1), this atomic resonant frequency ω0 can enhance spin squeezing greatly. These results are beneficial for achieving the required spin squeezing in experiments.

1. Introduction

Squeezed spin states, which were first introduced by Kitagawa and Ueda,[1] are quantum correlated states with reduced fluctuations in one of the collective spin components.[2,3] Such states have attracted considerable interest because they not only play significant roles in investigating many-body entanglement,[4,5] but also have important applications in atom interferometers and high-precision atom clocks.[6,7] In general, there are two methods to produce spin squeezing. One is based on a one-axis twisting Hamiltonian , where q is the nonlinear spin–spin interaction strength and Jx is the collective spin operator in the x direction. When the initial state is prepared as ∣Jz = −j⟩ for q > 0 (∣Jz = j⟩ for q < 0), the maximal squeezing factor for this one-axis twisting Hamiltonian scales as N−2/3,[1] where N is the total atomic number and J = N/2. In contrast, for the other two-axis twisting Hamiltonian , the maximal squeezing factor scales as N−1 with the same initial state.[1] Since the scaling of N−1 approaches the Heisenberg limit, implementing the two-axis twisting Hamiltonian in current experimental setups is very important and necessary.[815] A possible scheme is to transform the one-axis twisting Hamiltonian into the two-axis twisting Hamiltonian by applying pulse sequences or continuous driving in the two-component Bose–Einstein condensates.[1619] However, the experimental achievement of such two-axis twisting Hamiltonian is still challenging.

The ultrahigh-finesse cavities interacting with two-level particles (such as atoms, nitrogen vacancy centers, superconducting qubits, etc) are also powerful platforms to produce effective spin squeezing, since the photon of cavity mode can induce nonlinear spin–spin interaction and thus generate the required one-axis twisting Hamiltonian.[2030] Recently, multi-mode cavities[31] have attracted much attention both experimentally and theoretically.[3237] On one hand, these setups can be used to explore novel physics, such as the spin-orbit-induced anomalous Hall effect,[38,39] the crystallization and frustration,[40,41] the spin glass,[4245] and the gapless Nambu–Goldstone-type mode without rotating-wave approximation.[46] Moreover, two-mode field squeezing[47] and unconditional preparation of a two-mode squeezed state of effective bosonic modes[48] have also been achieved by introducing two cavities.

In this paper, inspired by recent experiments of cavity-assisted Raman transitions,[49,50] we mainly realize a generalized two-axis spin Hamiltonian , where χ is an anisotropic parameter and ω0 is an effective atomic resonant frequency, when an ensemble of ultracold six-level atoms interacts with two quantized cavity fields and two pairs of Raman lasers. The Hamiltonian achieved has a distinct property that the interaction strength q, the anisotropic parameter χ, and the effective atomic resonant frequency ω0 can be tuned independently. For reasonable parameters, the well-studied one- and two-axis twisting Hamiltonians are recovered. Numerical results reveal that for the standard two-axis twisting Hamiltonian (χ = −1 and ω0 = 0), the corresponding maximal squeezing factor scales as N−1, as expected. On the other hand, in the two-axis twisting Hamiltonian HTAT, spin squeezing is usually reduced when increasing the effective atomic resonant frequency ω0. Surprisingly, we find that by combining with the anisotropic parameter χ (> −1), this effective atomic resonant frequency ω0 can enhance spin squeezing greatly. These results are of benefit for achieving the required spin squeezing in experiments.

2. Two-axis spin Hamiltonian with independently-tunable parameters

Figure 1(a) shows our proposal, in which an ensemble of ultracold six-level atoms interacts with two quantized cavity fields and two pairs of Raman lasers. These six levels consist of two stable ground states (|0⟩ and |1⟩) and four excited states (|r1⟩, |r1⟩, |s1⟩, and |s2⟩) (see Fig. 1(b)). For 87Rb atom, the detailed levels can be chosen as |0⟩ → 5S1/2 |F = 1,mF = 0⟩, |1⟩ → 5S1/2 |F = 2,mF = 0⟩, |r1⟩ → 5P1/2 |F = 1,mF = 0⟩, |s1⟩ → 5P3/2 |F = 1,mF = 0⟩, |r2⟩ → 5P1/2 |F = 1,mF = − 1⟩, and |s2⟩ → 5P3/2 |F = 1,mF = −1⟩. Two independent photon modes, whose creation and annihilation operators are a (b) and a (b), mediate the |0⟩ ↔ |s1⟩ and |1⟩ ↔ |r1⟩ (|0⟩ ↔ |s2⟩ and |1⟩ ↔ |r2⟩) transitions (red and blue solid lines) with atom–photon coupling strengths gs1 (gs2) and gr1 (gr2), respectively. Whereas two pairs of Raman lasers govern the other transitions {|0⟩ ↔ |r1⟩, |1⟩ ↔ | s1⟩} and {|0⟩ ↔ | r2⟩, |1⟩ ↔ | s2⟩} (red and blue dashed lines) with Rabi frequencies {Ωr1, Ωs1} and {Ωr2, Ωs2}, respectively. Δr1,2 and Δs1,2 are the detunings from the excited states.

Fig. 1. (a) Proposed experimentally-feasible setup that an ensemble of ultracold six-level atoms interacts with two quantized cavity fields and two pairs of Raman lasers. A couple of orthogonally placed cavities Ca and Cb are arranged in a specific plane, and a guided magnetic field B is applied along one of the two cavities (see, for example, Ca) to fix a quantized axis and split the Zeeman sublevels of the atomic ensemble. Both the cavity modes and Raman lasers propagated in Ca (Cb) are assumed to be right-handed circularly polarized (linearly polarized along the guided magnetic field B). (b) Energy-level structure and their transitions induced by two photon modes and two pairs of Raman lasers. For 87Rb atom, the detailed levels can be chosen as |0⟩ → 5S1/2 |F = 1,mF = 0⟩, |1⟩ → 5S1/2 |F = 2,mF = 0⟩, |r1⟩ → 5P1/2 |F = 1,mF = 0⟩, |s1⟩ → 5P3/2 |F = 1,mF = 0⟩, |r2⟩ → 5P1/2 |F = 1,mF = −1⟩, and |s2⟩ → 5P3/2 |F = 1,mF = −1⟩. In such an arrangement, the transitions labeled by the red lines (blue lines) are thus π transitions (σ transitions).

The total Hamiltonian illustrated in Fig. 1 can be written, under the rotating wave approximation, as

where

In the Hamiltonians (2)–(4), ωr1,2, ωs1,2, and ω1 are the atomic frequencies, ω1s1,2 (φs1,2) and ω0r1,2 (φr1,2) are the frequencies (phases) of Raman lasers, respectively, and H.c. denotes the Hermitian conjugate.

By means of the Hamiltonian (1), we first realize a two-mode Dicke model with independently-tunable parameters. In the interaction picture with respect to the free Hamiltonian

where , , and , the Hamiltonian (1) can be rewritten as (see Appendix A)

Here Jx = (J+ + J)/2 and Jy = −i(J+J)/2, with , , and , are the collective spin operators, and , are the effective cavity frequencies, is the effective resonant atomic frequency, λ1 = Ωs1gs1/Δs1 = Ωr1gr1/Δr1 and λ2 = Ωs2gs2/Δs2 = Ωr2gr2/Δr2 are the effective atom photon coupling strengths, with , , and . We show that the Hamiltonian (5) is our realized two-mode Dicke model, based on recent experiments of cavity-assisted Raman transitions.[49,50] In contrast to the convectional two-mode Dicke model achieved in the two-level atoms, the Hamiltonian (5) has a distinct property that all parameters can be tuned independently. For example, ωA and ωB as we mentioned depend on the detunings Δa and Δb, respectively. Thus, they can range from the positive to the negative. The choice of the different cavity frequencies ωA and ωB help us to create a tunable two-axis spin Hamiltonian, see below. The effective atomic resonant frequency ω0 can also be controlled by the detuning Δ1. In addition, the effective atom–photon coupling strengths λ1 and λ2 can be driven by the Rabi frequencies of Raman lasers, respectively.

We now introduce two new unitary transformations, U1 = exp [G1Jx(aa)] and U2 = exp [G2Jy(bb)] with G1 = λ1/ωA and G2 = λ2/ωB, to rewrite the Hamiltonian (5) as

In the dispersive regime ({|ωA|,|ωB|} ≫ {λ1,λ2}), we can expand the Hamiltonian H with respect to G1 and G2 (up to second order). In addition, in such limit, the cavity modes only support the excitation of virtual photons (⟨aa⟩ → 0 and ⟨bb⟩ → 0),[2027] i.e., all terms with respect to the photons are eliminated. As a consequence, the Hamiltonian H turns into

where

In the Hamiltonian (6), the parameter q determines the nonlinear spin–spin interaction induced by the virtual photon, and the anisotropic parameter χ reflects the ratio between the different nonlinear spin–spin interactions and . Due to existence of these nonlinear spin–spin interactions with the anisotropic parameter χ and the effective atomic resonant frequency ω0, here we call the Hamiltonian (6) a generalized two-axis spin Hamiltonian. In addition, It shows clearly that all the parameters, including the interaction strength q, the anisotropic parameter χ, and the effective atomic resonant frequency ω0, can also be tuned independently in experiments.

For different parameters, the generalized two-axis spin Hamiltonian (6) can reduce to some well-studied Hamiltonians. For example, when χ > 0, the Hamiltonian (6) becomes a standard Lipkin–Meshkov–Glick model.[5258] When ω0 = 0, the Hamiltonian (6) reduces to a generalized two-axis twisting Hamiltonian . If further setting χ = −1, a standard two-axis twisting Hamiltonian is derived. Finally, when χ = 0, the Hamiltonian (6) turns into a generalized one-axis twisting Hamiltonian ,[5963] which reduces to the standard one-axis twisting Hamiltonian for ω0 = 0. These results imply that the Hamiltonian (6) has an important application in achieving the required spin squeezing.

Notice that when ωA < 0, the effective spin–spin interaction strength q > 0. Thus, we can use the initial state |Jz = −j⟩ to discuss spin squeezing of the generalized two-axis spin Hamiltonian (6). This initial state |Jz = −j⟩ can be easily prepared in experiments.

3. Spin squeezing factor

In order to investigate spin squeezing, it is very necessary to consider the following time-dependent squeezing factor:[1]

where n refers to an axis, which is perpendicular to the mean-spin direction n0 = J/|J| with , and ΔA2 = ⟨A2⟩ − ⟨A2 is the standard deviation. If , the spin state is squeezed, and vice versa. It should be noticed that here we do not use another spin squeezing factor proposed by Wineland et al.[64] in the study of Ramsey spectroscopy, which is substantially connected to the phase sensitivity in experiments. In fact, is related to via .

In the spherical coordinates, n0 = (sin θ cos φ, sinθ sin φ, cos θ), where θ = arccos (⟨Jz⟩ /| J|), and φ = arccos (⟨Jx⟩/|J| sin θ) for ⟨Jy⟩ > 0 or φ = 2π − arccos (⟨ Jx⟩/| J| sin θ) for ⟨Jy⟩ ≤ 0. Two orthogonal bases are given by n1 = (−sin φ,cos φ,0) and n2 = (− cos θ cos φ,cos θ sin φ,−sin θ). Thus, Jn1,2 = J· n1,2, Jn = Jn1 cos ϕ + Jn2 sin ϕ, and could be achieved, when ϕ varies from 0 to 2π in the plane that is perpendicular to the mean-spin direction n0. It should be noticed that in experiments, the maximal squeezing factor

is usually measured.[2,3] Thus, in the following discussion, we mainly focus on this physical quantity.

Before proceeding, we check the validity of the Hamiltonian (6), when the initial state is chosen as |Jz = −j⟩. For the generalized two-axis spin Hamiltonian, it is very hard to obtain the analytical result of the spin squeezing factor .[2] In Fig. 2, we plot the corresponding spin squeezing factors of the Hamiltonians (5) and (6). It can be seen clearly that the results of the Hamiltonian (6) are almost identical to those of the Hamiltonian (5). Therefore, we will apply the Hamiltonian (6) to discuss the experimentally measurable maximal squeezing factor in the rest of this paper.

Fig. 2. Numerical plot of the time-dependent spin squeezing factors of both the Hamiltonians (5) and (6), with the same initial states |Jz = −j⟩. In panels (a) and (b), ωA/ω0 = 200, ωB/ω0 = ± 200 [“+” for (a) and “−” for (b)], λ1/ω0 = 2, λ2/ω0 = 1, and N = 15. In panels (c) and (d), ωA/ω0 = 200, ωB/ω0 = ± 200 [“+” for (c) and “−” for (d)], λ1/ω0 = 1, λ2/ω0 = 2, and N = 20.
4. Maximal squeezing factor

We first address a simple case without the effective atomic resonant frequency (ω0 = 0), in which the generalized two-axis spin Hamiltonian (6) reduces to the generalized two-axis twisting Hamiltonian . In Fig. 3, we numerically plot the maximal squeezing factor of the Hamiltonian HGTAT as a function of the atomic number N for the different anisotropic parameters χ, when the initial state is chosen as |Jz = −j⟩. This figure shows clearly that when χ = 0, the generalized two-axis twisting Hamiltonian HGTAT becomes the standard one-axis twisting Hamiltonian , whose maximal squeezing factor scales as N−2/3.[1] When increasing the anisotropic parameter χ, the maximal squeezing factor decreases, i.e., spin squeezing is enhanced. In particular, when χ = −1, the Hamiltonian HGTAT turns into the standard two-axis twisting Hamiltonian , whose maximal squeezing factor scales as N−1,[1] as expected. In addition, the maximal squeezing factor as a function of the anisotropic parameter χ is also plotted in the insert part of Fig. 3. This figure shows that χ = −1 is an optimal point to achieve the maximal squeezing factor of the generalized two-axis twisting Hamiltonian HGTAT.

Fig. 3. Numerical plot of the maximal squeezing factors as a function of the atomic number N for the different anisotropic parameters χ. Insert: numerical plot of the maximal squeezing factor as a function of the anisotropic parameter χ, when the atomic number is chosen as N = 100. In both subfigures, q = 1 and the initial states are chosen as |Jz = −j⟩.

In the real experiments, the effective atomic resonant frequency ω0 always exists. In Fig. 4, we numerically plot the maximal squeezing factor of the Hamiltonian as a function of the effective atomic resonant frequency ω0, when the the initial state is chosen as |Jz = −j⟩. It can be seen from this figure that with the increasing of the effective atomic resonant frequency ω0 in the standard two-axis twisting Hamiltonian HTAT, the maximal squeezing factor increases, i.e., spin squeezing is reduced.

Fig. 4. Numerical plot of the maximal squeezing factor as a function of the effective atomic resonant frequency ω0. Insert: numerical plot of the time-dependent squeezing factor , when ω0/q = 0.3. In both subfigures, the initial states, the anisotropic parameter, and the atomic number are chosen as |Jz = −j⟩, χ = −1, and N = 100, respectively.

According to the above discussion, we argue that when increasing the anisotropic parameter χ (from χ = −1) or introducing the effective atomic resonant frequency ω0 in the standard two-axis twisting Hamiltonian HTAT, the maximal squeezing factor increases, i.e., spin squeezing is reduced. Surprisingly, when we control these two parameters simultaneously, spin squeezing can be enhanced greatly. To see this clearly, in Fig. 5 we numerically plot the maximal squeezing factors of the generalized two-axis spin Hamiltonian (6), i.e., , as a function of the effective atomic resonant frequency ω0 for different anisotropic parameters χ, when the initial state is chosen as |Jz = −j⟩. This figure shows that in the case of χ = −0.05 or χ = −0.5, when increasing the effective atomic resonant frequency ω0, the maximal squeezing factor first decreases greatly and then increases, i.e., spin squeezing is first enhanced greatly and then reduced. For the generalized one-axis twisting model , the maximal squeezing factor has a similar behavior (see green dashed-dotted line in Fig. 5),[59] but its magnitude cannot arrive at the order of our considered two-axis spin Hamiltonian (6), because these two Hamiltonians have different scalings with respect to the atomic number N. These results are of benefit for achieving the required spin squeezing in experiments.

Fig. 5. Numerical plot of the maximal squeezing factors as a function of the effective atomic resonant frequency ω0 for the different anisotropic parameters χ. The initial state and the atomic number are chosen as |Jz = −j⟩ and N = 100, respectively.
5. The experimental feasibility

Finally, we address the experimental feasibility by considering 87Rb atom, with the six levels depicted in Fig. 1. In this case, the decay rates of the atomic excited states and photons are about γ/2π = 6.0 MHz and κ/2π = 1.3 MHz (the ground states are stable and we do not need to consider the corresponding decays).[49,50] On the other hand, based on the transitions between the different levels, the atom–photon coupling strengths can reach gs1/2π = 20 MHz and gs2/2π = 15 MHz, respectively, both of which are much larger than the photon decay rate κ. As a consequence, the system is dominated by the Hamiltonian dynamics. When Ωs1/2π = 20 MHz and Δs1/2π = 100 MHz, together with the above estimated parameter values, the adiabatic condition for deriving Eq. (A2) is reasonable, and moreover, we have λ1/2π = 4 MHz and q/2π = 0.8 MHz for ωA/2π = −20 MHz. This choice of the detuning ωA also satisfies the dispersive condition, which is a key condition to realize our generalized two-axis spin Hamiltonian (6). For the anisotropic parameter χ, it can be easily controlled by tuning both the detuning ωB and the Rabi frequency Ωs2.

When N = 100 with above parameters, the numerical result shows that the shortest time for generating the maximal squeezing factor is about tm ≃ 25 ns (see, for example, in the insert part of Fig. 4), which is shorter than the photon lifetime κ−1. With the increase of the atomic number N, tm becomes shorter and shorter, since tm = 1/N0.8, as shown in Fig. 6. In the real experiments with N = 104 or 105,[49,50] tm = 0.6 ns or 0.1 ns. This indicates that the maximal spin squeezing can be created very quickly in experiments. Notice that for more rigorous consideration of the effect inducing the photon decay κ, we should use the method of quantum master equation. However, here we mainly take the first step to realize the two-axis twisting Hamiltonian and achieve spin squeezing (approaching the Heisenberg limit) in the two cavities.

Fig. 6. Numerical plot of the shortest time tm for generating the maximal squeezing factor as a function of the atomic number N. The initial state, the effective atomic resonant frequency, and the anisotropic parameter are chosen as |Jz = −j⟩, ω0/q = 0.3, and χ = −1, respectively.
6. Conclusion

In summary, we have proposed an experimentally feasible system, in which an ensemble of ultracold six-level atoms interacts with two quantized cavity fields and two pairs of Raman lasers, to realize a generalized two-axis spin Hamiltonian . We have numerically calculated the experimentally measurable maximal squeezing factor and revealed that when ω0 = 0 and χ = −1, the maximal squeezing factor scales as N−1. More importantly, we have found that by combining with the anisotropic parameter χ (> −1), the effective atomic resonant frequency ω0 can enhance spin squeezing greatly. Our results are of benefit for achieving the required spin squeezing in experiments, and have a potential application in quantum information and quantum metrology.

Reference
1Kitagawa MUeda M 1993 Phys. Rev. 47 5138
2Ma JWang XSun C Pet al. 2011 Phys. Rep. 509 89
3Robins N PAltin P ADebs J Eet al. 2013 Phys. Rep. 529 265
4Duan L MSørensen ACirac J Iet al. 2000 Phys. Rev. Lett. 85 5643
5Sørensen ADuan L MCirac J Iet al. 2001 Nature 409 63
6Santarelli GLaurent PLemonde Pet al. 1999 Phys. Rev. Lett. 82 4619
7Hammerer KSørensen A SPolzik E S 2010 Rev. Mod. Phys. 82 1041
8Helmerson KYou L 2001 Phys. Rev. Lett. 87 170402
9Bouchoule IMølmer K 2002 Phys. Rev. 65 041803
10Zhang MHelmerson KYou L 2003 Phys. Rev. 68 043622
11André ALukin M D 2002 Phys. Rev. 65 053819
12André ADuan L MLukin M D 2002 Phys. Rev. Lett. 88 243602
13Thomsen L KMancini SWiseman H M 2002 J. Phys. 35 4937
14Ng H TLaw C KLeung P T 2003 Phys. Rev. 68 013604
15Cappellaro PLukin M D 2009 Phys. Rev. 80 032311
16Liu Y CXu Z FJin G Ret al. 2011 Phys. Rev. Lett. 107 013601
17Shen CDuan L M 2013 Phys. Rev. 87 051801
18Zhang J YZhou X FGuo G Cet al. 2014 Phys. Rev. 90 013604
19Huang WZhang Y LZou C Let al. 2015 Phys. Rev. 91 043642
20Deb R NAbdalla M SebaweHassan S SNayak N 2006 Phys. Rev. 73 053817
21Nielsen A E BMølmer K 2008 Phys. Rev. 77 063811
22Schleier-Smith M HLeroux I DVuletić V 2010 Phys. Rev. 81 021804
23Leroux I DSchleier-Smith M HVuletić V 2010 Phys. Rev. Lett. 104 073602
24Chen ZBohnet J GSankar S Ret al. 2011 Phys. Rev. Lett. 106 133601
25Dalla TorreOtterbach JDemler EVuletić Vet al. 2013 Phys. Rev. Lett. 110 120402
26Chen ZBohnet J GWeiner J Met al. 2014 Phys. Rev. 89 043837
27Yu LFan JZhu Set al. 2014 Phys. Rev. 89 023838
28Auccaise RAraujo-Ferreira A GSarthour R Set al. 2015 Phys. Rev. Lett. 114 043604
29Huang X YXiang YSun F Xet al. 2015 Acta Phys. Sin. 64 160304 (in Chinese)
30Zhong WLiu JMa Jet al. 2014 Chin. Phys. 23 060302
31Messina AManiscalco SNapoli A2003J. Mod. Phys.501
32Mariantoni MDeppe FMarx AGross Ret al. 2008 Phys. Rev. 78 104508
33Mariantoni MWang HBialczak R Cet al. 2011 Nat. Phys. 7 287
34Egger D JWilhelm F K 2013 Phys. Rev. Lett. 111 163601
35Wickenbrock AHemmerling MRobb G R Met al. 2013 Phys. Rev. 87 043817
36Krimer D OLiertzer MRotter Set al. 2014 Phys. Rev. 89 033820
37McKay D CNaik RReinhold Pet al. 2015 Phys. Rev. Lett. 114 080501
38Larson JLevin S 2009 Phys. Rev. Lett. 103 013602
39Larson J 2010 Phys. Rev. 81 051803
40Gopalakrishnan SLev B LGoldbart P M 2009 Nat. Phys. 5 845
41Gopalakrishnan SLev B LGoldbart P M 2010 Phys. Rev. 82 043612
42Gopalakrishnan SLev B LGoldbart P M 2011 Phys. Rev. Lett. 107 277201
43Strack PSachdev S 2011 Phys. Rev. Lett. 107 277202
44Buchhold MStrack PSachdev Set al. 2013 Phys. Rev. 87 063622
45Andreanov AMüller M 2012 Phys. Rev. Lett. 109 177201
46Fan JYang ZZhang Yet al. 2014 Phys. Rev. 89 023812
47Guzmán RRetamal J CSolano Eet al. 2006 Phys. Rev. Lett. 96 010502
48Parkins A SSolano ECirac J I 2006 Phys. Rev. Lett. 96 053602
49Baumann KGuerlin CBrennecke Fet al. 2010 Nature 464 1301
50Baden M PArnold K JGrimsmo A Let al. 2014 Phys. Rev. Lett. 113 020408
51Dimer FEstienne BParkins A Set al. 2007 Phys. Rev. 75 013804
52Lipkin H JMeshkov NGlick N 1965 Nucl. Phys. 62 188
53Lipkin H JMeshkov NGlick N 1965 Nucl. Phys. 62 199
54Lipkin H JMeshkov NGlick N 1965 Nucl. Phys. 62 211
55Dusuel SVidal J 2004 Phys. Rev. Lett. 93 237204
56Ribeiro PVidal JMosseri R 2007 Phys. Rev. Lett. 99 050402
57Orús RDusuel SVidal J 2008 Phys. Rev. Lett. 101 025701
58Chen GLiang J QJia S 2009 Opt. Exp. 17 19682
59Law C KNg H TLeung P T2011Phys. Rev. A63055601
60Vidal JPalacios GMosseri R 2004 Phys. Rev. 69 022107
61Vidal JMosseri RDukelsky J 2004 Phys. Rev. 69 054101
62Morrison SParkins A S 2008 Phys. Rev. Lett. 100 040403
63Chen GWang XLiang J Qet al. 2008 Phys. Rev. 78 023634
64Wineland D JBollinger J JItano W Met al. 1992 Phys. Rev. 46 R6797