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

Noise squeezing of fields that bichromatically excite atoms in a cavity

Open Access Open Access

Abstract

It is well known that bichromatic excitation on one common transition can tune the emission or absorption spectra of atoms due to the modulation frequency dependent non-linearities. However little attention has been focused on the quantum dynamics of fields under bichromatic excitation. Here we present dissipative effects on noise correlations of fields in bichromatic interactions with atoms in cavities. We first consider an ensemble of two-level atoms that interacts with the two cavity fields of different frequencies and considerable amplitudes. By transferring the atom-field nonlinearities to the dressed atoms we separate out the dissipative interactions of Bogoliubov modes with the dressed atoms. The Bogoliubov mode dissipation establishes stable two-photon processes of two involved fields and therefore leads to two-mode squeezing. As a generalization, we then consider an ensemble of three-level Λ atoms for cascade bichromatic interactions. We extract the Bogoliubov-like four-mode interactions, which establish a quadrilateral of the two-photon processes of four involved fields and thus result in four-mode squeezing.

© 2016 Optical Society of America

1. Introduction

The reduction of quantum noise is one of the important issues in quantum optics, laser physics, and nonlinear optics. Quantum noise is harmful for the high-precision experiments and thus ones seek many ways to suppress the quantum noise. Earlier, Caves et al. [1] first noted the possibilities of manipulating quantum fluctuations with the aim of precision measurement. When the quantum noise of one quadrature of an optical field is reduced below the standard quantum limit at the expense of the increasing noise in the conjugate quadrature, we can say that the fields are prepared in the squeezed states [2,3]. The bi- or multimode squeezed states are most closely related to the entangled states, which are indispensable resources in quantum information and quantum communication networks [4–7]. Many schemes for generation of the squeezed and entangled states have been proposed based on either the coherent evolutions [8–16] or the reservoir engineering [17–23]. The essential physics is due to the two-photon processes, which occur typically in a correlated emission laser or in four-wave mixing [16, 24, 25]. It is meaningful to seek other systems for realizing the noise squeezing of fields.

It has become aware that the bichromatic or ploychromatic interactions constitute an important field in quantum optics, laser physics, and nonlinear optics. Bichromatic excitation on one common atomic transition can lead to the modulation frequency dependent nonlinearities. Thus many novel effects occur while they are absent for monochromatic interactions with atoms. The spontaneous emission, absorption and dispersion spectra become comblike and dependent on the modulation frequency [26–44]. Spontaneous-emission-mediated emissive multiphoton transitions (the usual processes of stimulated emission and absorption are reversible) as irreversible processes can give rise to lasing without population inversion [45, 46]. Two-photon absorption is dramatically suppressed when a pair of bichromatic fields drive a three-level atomic system [47]. Experimental and theoretical investigation of cavity-modified dynamics of two-level atoms driven by a strong bichromatic field also shows that the width of atomic resonances may become smaller than the cavity linewidth [48]. Additionally, selective elimination of the spectral lines is obtained by varying the amplitudes and phases of the trichromatic components [49]. For a bichromatically excited two-level atom, there exist fluorescent two-photon processes, which are responsible for the possible noise squeezing in the central fluorescence spectrum [50, 51]. However, the above work has focused on the dynamics of atoms and there has been little work on noise squeezing of fields under bichromatic excitation. Once two fields of different frequencies and considerable amplitudes are simultaneously coupled to one common atomic transition, there appear the cross nonlinearities and both of the self and cross nonlinearities depend strongly on the modulation frequency. In order to study the effects of the modulation frequency dependent nonlinearities, one usually has to employ the harmonic expansions. In particular, the expanded terms of so many high orders prevent one from identifying the direct mechanisms for the quantum correlations.

Here we present an efficient way without the harmonic expansions and show clearly that the Bogoliubov mode dissipation causes the noise squeezing of fields in bichromatic interactions with atoms. The atom-field nonlinearities are transferred to the dressed atoms for proper cavity-detunings, and the Bogoliubov mode dissipative interactions are separated from the system dynamics. We first consider a fundamental model in which an ensemble of two-level atoms interact with the two cavity fields of different frequencies and considerable amplitudes. It is shown that the Bogoliubov mode dissipative interactions lead to the two-photon processes of two involved fields and thus to the two-mode squeezing. Then an ensemble of three-level Λ atoms is considered for cascade bichromatic interactions. It is found that dissipative interactions of the Bogoliubov-like modes with the dressed atoms lead to a quadrilateral of the two-photon processes of four involved fields and thus to the four-mode squeezing. Our work can be extended to other coherent systems and facilitate the analysis of the noise squeezing of fields in frequency dependent nonlinear interactions.

Particularly, in our system the bichromatic fields are not in Λ-configuration interaction with three-level atoms as in electromagnetic induced transparency and/or Raman processes but are directly coupled to the two-level atomic system, where the nonlinearities occur and provide the ability to suppress the quantum noise. When the driving field couples to the two-level atoms, the most remarkable feature of the atomic dynamics is Rabi resonance. We use driving fields of the same frequencies and tune the cavity fields resonant with Rabi sidebands. The Rabi resonances give rise to the collective dissipation and to squeezing. Using cascade bichromatic excitations in a Raman system, we also can obtain the four-mode squeezing.

The remaining part of the present paper is organized as follows. In section 2, we describe the bichromatic interactions with two-level atoms in cavities, derive the dressed atom-field interactions, and present the Bogoliubov mode dissipation effects on the field correlations. In section 3, we extend to the cascade bichromatic interactions in three-level atoms for the four-mode squeezing. Finally, the conclusion is given in section 4.

2. Bichromatic interactions with two-level atoms

We consider an ensemble of N two-level atoms at the intersection of two optical cavities, each of which is pumped by one external field, as shown in Fig. 1(a). The two cavity modes are simultaneously coupled to the common electronic dipole-allowed transition of the two-level atoms from the ground state |1〉 to the excited state |2〉, as shown in Fig. 1(b). The system under consideration is clearly different from those that are based on the wave mixing interactions. For the latter, the cavities are driven only by the vacuum reservoir but not by the external classical fields and the atoms are excited only by one monochromaic field [17–19]. In the present system, in sharp contrast, the two fields have different frequencies and considerable amplitudes, and the atoms are bichromatically excited. The master equation for density operator ρ of the atom-field system is written in the dipole approximation and in an appropriate rotating frame as [2,3]

ρ˙=i[H,ρ]+ρ,
with the system Hamiltonian
H={Δσ11+[(g1a1+g2a2)σ21+H.c.]}+j=12[Δcjajaj+i(εjajεj*aj)],
where H.c. denotes Hermitian conjugate of the terms before it, a1,2 and a1,2 are the annihilation and creation operators for cavity modes, and σkl=μ=1Nσklμ ( σklμ=|kμlμ|; k, l = 1, 2) are the collective projection operators for k = l and the collective spin-flip operators for kl. Δ = ω21ω0 and Δcj = ωcjω0 are respectively the detunings of atomic transition frequency ω21 and cavity resonance frequencies ωcj from the external driving field frequencies (which are assumed to be the same ω1,2 = ω0). g1,2 are the coupling strengths between the atoms and the cavity fields, and ε1,2 are the amplitudes of the external driving fields. The decay term in Eq. (1) takes the form
ρ=γ2σ12ρ+j=12κj2ajρ,
where σ12 ρ and aj ρ describe the atomic decay with rate γ and the cavity losses with rates κj, respectively, and o ρ = 2oρooρoo, o = σ12, aj.

 figure: Fig. 1

Fig. 1 (a) An ensemble of two-level atoms is placed at the intersection of two cavities, into which two external coherent fields are input respectively. (b) The two cavity fields (annihilation operators a1,2) of different frequencies are coupled to the common transition of the atoms.

Download Full Size | PDF

For the driven atom-cavity system, the high order nonlinearities can be seen from the fields equations through adiabatic elimination of the atomic variables under the symmetrical cavity detunings Δc1 = −Δc2 = Δc, and symmetrical parameters g1,2 = g, and κ1,2 = κ. For equal input intensity Iin=4|gε1,2|2κ2γ2, we obtain the steady state equation for the equal field intensity I=|ga1,2|2γ2 as

Iin=I[(1+𝒜)2+(Δ¯c+𝒟)2],
where 𝒜 and 𝒟 represent respectively the absorption and the dispersion due to the atoms and depend strongly on the cavity field intensity I through the relations
𝒜=8C4Δ¯2+32I+1,𝒟=16CΔ¯4Δ¯2+32I+1.
Here we have defined the cooperator parameter C=g2Nκγ and the other parameters Δ¯=Δγ and Δ¯c=Δck. When Δ = 0, we have 𝒜 ≠ 0 and 𝒟 = 0, which means that the atoms have vanishing dispersion but nonvanishing absorption. For Δ ≠ 0, we can expand Eqs. (4) and (5) into a divergent series of infinitely high orders when the intensity is of considerable value. Bistability or multistability is obtainable due to the absorptive and/or dispersive nonlinearities. Now it is invalid to treat the nonlinearities in a perturbative way. Instead we will use a non-perturbative way to explore the effects of the nonlinearities on the field correlations.

2.1. Bogoliubov mode interactions with dressed atoms

First we make a unitary transformation [52] via U1=j=12exp(λjajλj*aj), where λj=εjκ+iΔcj (j = 1, 2), and linearize the cavity fields aj = 〈aj〉 + δaj. By doing this we rewrite the Hamiltonian (2) as

H=Ha+Hc+HI,
where
Ha=Δσ22+2(Ωσ21+Ω*σ12)
describes the interaction of the atoms with the classical pump fields ε1,2 and the semiclassical parts of the cavity fields 〈a1,2〉 (with the total Rabi frequency Ω=j=1,22(λj+gaj)),
Hc=Δc(δa1δa1δa2δa2)
denotes the free part of the fluctuating cavity fields, and HI represents the interaction of the atoms with the fluctuating cavity fields and will be given later.

Then we transform to a dressed states representation. We assume that the Rabi frequencies are real, equal and much stronger than the atomic and cavity decay rates Ω = |Ω| ≫ (γ, κ). By diagonalizing the Hamiltonian Ha, we obtain the dressed states that are expressed in terms of the bare atomic states as [53]

|+=sinθ|1+cosθ|2|=cosθ|1sinθ|2,
where we have defined tan(2θ)=ΩΔ, 0<θ<π2. These dressed states |±〉 correspond respectively to the eigenvalues λ± = (Δ ± d)/2, d=Δ2+Ω2. In the dressed representation the Hamiltonian Ha of the atoms becomes the free form Ha = ħ(λ+σ++ + λσ−−). Applying the dressed states transformation to the decay terms of the atoms we obtain the steady state populations N± = 〈σ±±〉 as
N+=Nsin4θcos4θ+sin4θ,N=Ncos4θcos4θ+sin4θ.
When Δ = 0 the atoms are equally populated in the two dressed states. When Δ > 0 (Δ < 0) we have N > N+ (N < N+). We will consider the general case Δ ≠ 0.

In the following calculations, while we give dependence of quantum correlations on the scaled atom-driving detunging ΔΩ, the cavities are tuned to be resonant with the Rabi sidebands Δc1 = −Δc2 = d. All results we will obtain below hold when ||Δc1,2| − d| ≪ d. An increase in the cavity detunings will spoil the quantum correlations. If the deviations of the cavity detunings from Rabi resonances are comparable to or larger than the level spacing d, the two cavity fields are not in selective interactions with the higher and lower sidebands and the present treatment does no longer hold. Throughout the present work we assume that the above conditions for Rabi resonances are guaranteed.

Following the above treatment, we write the total free Hamiltonian H0 of atom-field system as H0 = Ha + Hc. By using the dressed atomic states, and making a further unitary transformation with U2 = exp (−iH0t/ħ) and taking a rotating-wave approximation, we rewrite the interaction Hamiltonian HI as

HI=g(a1cos2θa2sin2θ)σ++H.c.,
where we have have used g1,2 = g and dropped the symbol δ for the simplicity and clearness. From Eq. (11) it is easily seen that each mode is in resonant interaction with the corresponding dressed transition, as shown in Fig. 2. In order to describe the physical mechanism more clearly, we introduce a pair of Bogoliubov modes for the cavity fields
b1=a1coshra2sinhr,b2=a2coshra1sinhr,
which act as a new pair of independent boson modes, i.e., [bj,bk]=δjk, [bj, bk] = 0. Then the Hamiltonian in Eq. (11) is rewritten as
HI=G(b1σ++σ+b1)forΔ>0,HI=G(b2σ++σ+b2)forΔ<0,
where we have defined the new coupling constant G=gcos(2θ) and the squeezing parameter tanh r = tan2 θ for Δ > 0, and G=gcos(2θ) and tanh r = cot2 θ for Δ < 0. The Hamiltonian (13) shows that the dressed transition |+〉 ↔ |−〉 is simply accompanied by emission and absorption of either of the two Bogoliubov modes.

 figure: Fig. 2

Fig. 2 The individual cavity fields a1,2 are both in resonant interactions with the dressed atoms. The two fields combine into the Bogoliubov mode b1 for Δ > 0 (N > N+) or b2 for Δ < 0 (N+ > N).

Download Full Size | PDF

2.2. Dissipation and two-mode squeezing

Now we use the interactions of the Bogoliubov modes with the dressed atoms to analyze the field correlations and the responsible mechanism.

2.2.1. Dissipation and two-photon processes

We assume that the atomic variables decay much more rapidly than the cavity fields, i.e., γκ. Then by tracing out the atomic variables, we derive the master equation for the reduced density operator of the cavity modes ρ̃ = Tra ρ as

ρ˜˙=𝒦aρ˜+cρ˜,
where 𝒦aρ˜=iTra[HI,ρ] describes the atom-field interaction. Following the standard technique [2,3], we obtain the atomic contribution part of the master equation for Δ > 0 as
𝒦aρ˜=A(2b1ρ˜b1b1b1ρ˜ρ˜b1b1)+B(2b1ρ˜b1b1b1ρ˜ρ˜b1b1),
where A = G2N/Γ, B = G2N+/Γ, and Γ=12γ+14γsin2(2θ). The A and B terms describe the absorption and gain of Bogoliubov mode b1 by the dressed atoms, respectively.

When ABκ, i.e., the absorption (i.e., the dissipation) by the atoms dominates [3] over the gain (i.e., the amplification). Since the Bogoliubov mode b1 undergoes dissipation, and it evolves into thermal vacuum state. It should be noted that b2 mode decouples from the atoms. In contrast, for the case of Δ < 0, the Bogoliubov mode b2 undergoes dissipation and evolves into thermal vacuum state while b1 mode decouples from the atoms. Expanding Eq. (15) in terms of a1,2 we have the terms

β1a1ρ˜a2+β2ρ˜a1a2+β3a1a2ρ˜+(12)+H.c.,
where the parameters β1–3 depend on the absorption and gain coefficients A and B but are not necessarily written out here. Such terms describe the two-photon processes as in a two-mode squeezed vacuum [54]. When the dissipation dominates, the two-photon processes are stable and thus responsible for quantum noise squeezing if existent. This is the very effect of the dissipation by the dressed atoms on the field correlations. In what follows we will show clearly that the dissipation based two-photon processes determine the two-mode squeezing and entanglement.

2.2.2. Approximately analytic two-mode variances

To show clearly the effects of dissipation on the field correlations, we first give the approximately analytic description. For this purpose we assume that dissipation by the dressed atoms dominates over the vacuum reservoir and neglect temporarily the cavity losses. We define the quadrature operators δxo=12(o+o) and δpo=i2(oo), o = a1,2, b1,2. By using the Bogoliubov transformation relations, we can express the two-mode quadrature operators in terms of the Bogoliubov modes as

δX=δxa1δxa2=e4(δxb1δxb2),δP=δpa1+δpa2=er(δpb1+δpb2).
Then the corresponding variances are calculated as
(δX)2=e2r(1+:(δxb1)2:),(δP)2=e2r(1+:(δpb1)2:),
where we have used the fact that b2 mode decouples from the system and remains in vacuum state. In Eq. (18) we use “ :: ” for the fluctuations in the normal operator order. If the variance satisfies relation 〈(δX)2〉 < 1 or 〈(δP)2〉 < 1 the two-mode squeezing happens. According to the continuous variable bipartite entanglement criterion proposed by Duan et al. [5], the two-mode squeezing and entanglement occurs if the variances of Einstein-Podolsky-Rosen-like operators satisfy the inequality 〈(δX)2〉 + 〈(δP)2〉 < 2.

Due to the dissipation by the atoms, the b1 mode evolves into its thermal vacuum state. From Eqs. (14) and (15) we can obtain nonvanishing correlation b1b1=N+NN+ for Δ > 0. Similarly we have the nonvanishing correlation b2b2=NN+N for Δ < 0. Thus we have the quantum fluctuations :(δxbj)2:=12bj2+bj2+2bjbj=bjbj for Δ > 0 (j = 1) and Δ < 0 (j = 2). Finally the two-mode correlations are obtained as

(δX)2=(δP)2=e2rN+NN+forΔ>0,(δX)2=(δP)2=e2rNN+NforΔ<0.
It is easy to see that the conditions for entanglement in the quadratures X and P are the same as for the squeezing. We plot the variance 〈(δX)2〉 versus the scaled detuning ΔΩ in Fig. 3. The cavity detunings are adjusted to Rabi resonances according to Δc1=Δc2=Δ2+Ω2, as stated above. The two-mode squeezing exists in whole region except at Δ = 0. However, the above analytic calculation is just valid far away from Δ = 0. At the point Δ = 0, the atoms are equally populated in two dressed states, i.e., N+ = N, the absorption is equal to the gain and no net dissipation occurs. Near the point Δ = 0 (as shown by shadow in Fig. 3), the condition ABκ is not well satisfied, i.e., the absorption of b1 mode by atoms no longer plays the dominant role. Therefore, the analytic solution does not hold within the region around Δ = 0. As numerical verification, we include the cavity losses in the system dynamics and calculate the field correlations as follows.

 figure: Fig. 3

Fig. 3 Approximately analytic variance 〈(δX)2〉 versus the scaled detuning ΔΩ. Within the shadow at ΔΩ=0, the conditions for approximately analytic solution are not well met.

Download Full Size | PDF

2.2.3. Numerical verification and discussion

In order to include the cavity losses and to calculate the quantum correlations, we can apply a standard Langevin approach by means of the generalized P representation [54–56]. The individual operators are arranged in the normal ordering a1, a2, a2, a1, and the correspondences between the c numbers and the operators are defined as αjaj, αj*aj (j = 1, 2). The c number Langevin equations for the individual modes are derived from Eqs. (14) and (15) as

α˙1=λ11α1+λ12α2*+Fα1,α˙2=λ21α1*+λ22α2+Fα2,
where we have defined the parameters λ11=κ2g2Γ(NN+)cos4θ, λ22=κ2g2Γ(N+N)sin4θ, and λ12=λ21=g241(NN+)sin2(2θ). The F’s are noise terms with zero means and correlations 〈Fo (t)Fo′ (t′)〉 = 2Doo′ δ(tt′). The diffusion coefficients for individual modes can be calculated from the reduced master equation as follows
2Dα1*α1=2g2ΓN+cos4θ,2Dα1*α2=2g2ΓNsin4θ,2Dα1α2=g2N4Γsin2(2θ),2Dα1*α2*=2Dα1α2.
Substituting Eqs. (20) and (21) into the set of equations ddtoo=dodto+ododt+2Doo, o, o′ = α1,2, α1,2* and making a stability analysis, we obtain the quantum correlations 〈oo′〉, which are used for the final variances 〈(δX)2〉 and 〈(δP)2〉.

In Fig. 4, we show the numerical results for different rates of cavity losses. It is seen that the squeezing exists almost in the entire region except at Δ = 0. This is in a good agreement with the analytic solution. Near the point Δ = 0, the populations trend to be equal, N+N. The absorption term proportional to A is no longer dominant over the amplification term proportional to B. Therefore, the Bogoliubov mode dissipation by the atoms is greatly degraded and can’t suppress the cavity loss. As Δ approaches zero, the squeezing reduces and disappears finally at Δ = 0. This shows that the squeezing and entanglement don’t occur when the dispersive nonlinearity vanishes completely at Δ = 0. The best squeezing arrives at 50%.

 figure: Fig. 4

Fig. 4 The variance 〈(δX)2〉 versus the scaled detuning ΔΩ for different rates of cavity losses κ = 0.2γ (red dashed), κ = 0.1γ (black dotted), and κ = 0.01γ (blue solid). We have used the other parameter gN=10γ.

Download Full Size | PDF

Finally, we should note the essentially different ways between the monochromatic and bichromatic interactions with atoms for establishing the two-photon processes, which are responsible for the squeezing. For the present case, the two-photon processes are established by using two fields simultaneously to a single atomic transition. In sharp contrast, for the monochromatic case, the two-photon processes are created by using two cascaded atomic transitions (e.g., the three-level atoms in Λ configurations), each of which is coupled to one monochromatic field. For the bichromatic case on a single transition, although only either of two Bogoliubov modes mediates interactions and the best squeezing is limited to 50%, the parameter range for the squeezing is the almost entire region except Δ = 0. For the cascaded transitions with monochromatic field, although two Bogoliubov modes mediate interactions and the squeezing can be enhanced for especial parameters, the parameter range for squeezing is greatly limited [20]. Combining the above two cases we can use the three-level atoms for cascade bichromatic interactions and for four-mode squeezing, as will be shown in the following section.

3. Cascade bichromatic interactions with three-level atoms

In our setup, four optical cavities intersect at a common spot, where an ensemble of N three-level Λ-type atoms is placed, as shown in Fig. 5(a). The atom has two metastable states |1〉 and |2〉 and one excited state |3〉. The two cavity fields (the annihilation operators aj and aj+2, the creation operators aj and aj+2, and the frequencies ωcj and ωcj+2) are driven by the classical fields of the same frequency ωj, and coupled to the electronic dipole-allowed transition |j〉 − |3〉 (j = 1, 2), as shown in Fig. 5(b). The master equation for density operator ρ of the atom-field system is written in the dipole approximation and in an appropriate rotating frame as [2, 3] ρ˙=i[H,ρ]+ρ, where the Hamiltonian of the system is written as

H=j=12{Δjσjj+[(gjaj+gj+2aj+2)σ3j+H.c.]}+j=14[Δcjajaj+i(εjajεj*aj)],
where σkl=μ=1Nσklμ ( σklμ=|kμlμ|; k, l = 1, 2, 3) are the collective projection operators for k = l and the collective spin-flip operators for kl. The Δ’s terms describe the free terms due to the detunings, Δj = ω3jωj and Δcj(j+2) = ωcj(j+2)ωj (j = 1, 2). The g’s terms give the interactions of the cavity fields with the atoms, and g1–4 are the coupling strengths between atoms and cavity fields. The ε’s terms show the external driving, and ε1–4 stand for the classical driving amplitudes. The ℒρ term describes the decay due to the vacuum environment and takes the form
ρ=j=12γj2σj3ρ+j=14κj2ajρ,
where contained in the first summation term are the atomic relaxations with rates γj (j = 1, 2) and included in the second summation are the cavity losses due to the vacuum environment term with rates κj (j = 1 – 4), and o ρ = 2oρooρoo, o = σj3, aj.

 figure: Fig. 5

Fig. 5 (a) Four optical cavities intersect at a common spot, where an atomic ensemble of three-level atoms is placed. (b) The interaction of the four cavity fields with the atoms in Λ configuration.

Download Full Size | PDF

For the bistable or multistable system, the high order nonlinearities can be seen from the fields equations through adiabatic elimination of the atomic variables under the symmetrical detunings Δ1 = −Δ2 = −Δ and Δcj = −Δcj+2 = Δc (j = 1, 2), and equal parameters g1–4 = g, γ1,2 = γ, and κ1–4 = κ. For equal input intensity Iin=|gεl|2κ2γ2, and the collective field intensity I=|gal|2γ2 (l = 1 – 4) satisfies the nonlinear equation Iin = I[(1 + 𝒜)2 + (Δ̄c + 𝒟)2], where 𝒜 and 𝒟 represent respectively the nonlinear absorption and dispersion by the atoms and depend strongly on the cavity field intensity I through the relations

𝒜=CΔ¯2Δ¯4+Δ¯2+Δ¯2I+I2,𝒟=CΔ¯(Δ¯2+I)Δ¯4+Δ¯2+Δ¯2I+I2.
Here we have defined C=g2Nκγ, Δ¯=Δγ, and Δ¯c=Δcκ. On the exact resonance Δ = 0, we have 𝒜 = 𝒟 = 0, which means that the atoms are transparent to the fields. For Δ ≠ 0, when expanding the above nonlinear equation we can obtain a divergent series of infinitely high orders for the intensity I of considerable value. Bistability or multistability exists due to the nonlinearities. The cross nonlinearities also appear between the modes aj and aj+2 (j = 1, 2). It is obviously invalid to treat the nonlinearities in a perturbative way. In the near-resonant regime (|Δ||gaj+aj+2|), the nonlinearity to absorption ratio is much larger than unity, i.e., |𝒟|𝒜I|Δ¯|1. However, it is hard to see any effects of the nonlinearities on the quantum correlations between any two modes. This means a gap between the nonlinearities and the multimode quantum correlations. Our purpose is to identify that the gap is the dissipation of Bogoliubov-like collective modes by the atoms.

3.1. Bogoliubov-like four-mode interactions with dressed atoms

In order to extract the mechanism hidden behind the nonlinearities, we merge the nonlinearities into the dressed atomic states and analyze the dressed atom-photon interactions as follows. First, we make a unitary transformation [52] via U1=j=14exp(λjajλj*aj), where λj=εjκ+iΔcj (j = 1 – 4). Then by linearizing the cavity fields aj = 〈aj〉 + δaj, we rewrite the Hamiltonian as H = Ha + Hc + HI, where

Ha=j=12[(1)j+1Δσjj+(Ωj+Ωj+2)σ3j+(Ωj*+Ωj+2*)σj3]
describes the interaction of the atoms with the classical pump fields and the semiclassical parts of the cavity fields with the half of Rabi frequency Ωj = λj + gaj〉, j = 1 – 4,
Hc=j=12Δc(δajδajδaj+2δaj+2)
denotes the free part of the fluctuating fields, and HI represents the interaction of the atoms with the fluctuating fields and will be given later.

We transfer the common phase of Ωj + Ωj+2 to the atomic operators and are left with Ω¯j=|Ωj+Ωj+2ei(φj+2φj)|=Ω2|1+eiΦ|, where we have assmued that Rabi frequencies have their phases φj+2φj = Φ, j = 1, 2. We also assume that the amplitudes of Rabi frequencies are equal and much stronger than the atomic and cavity decay rates, Ω̄j ≫ (γ, κ). By diagonalizing the Hamiltonian Ha, we obtain the dressed states, which are expressed in terms of the bare atomic states as [53]

|+=1+sinθ2|1+1sinθ2|2+cosθ2|3,|0=cosθ2|1+cosθ2|2+sinθ|3,|=1sinθ2|1+1+sinθ2|2cosθ2|3,
where we have have definde sinθ=Δd, cosθ=2Ω¯d, d=Δ2+2Ω¯2, and Ω̄1,2 = Ω̄. These dressed states |0〉 and |±〉 have respectively their eigenvalues λ0,± = 0, ±ħd, which means equal spacings. In the dressed states representation the Hamiltonian Ha becomes a free part Ha = ħd(σ++σ−−). Expressing the atomic decay term in terms of the dressed states we obtain the steady state populations Nj = 〈σjj〉 (j = 0, ±) as
N0=Ncos4θ1+3sin4θ,N±12(NN0).
It is worth noting that on the exact resonance Δ = 0 (sin θ = 0, cos θ = 1), the atoms are trapped in the dark state |0=12(|1+|2), i.e., N0 = N, N± = 0. In this case one gets the maximal coherence σ12=N2 but vanishing nonlinearities, as pointed out in the above section. On the off-resonant case we have N0 > N± when |Δ|Ω<1, and N0 < N± when |Δ|Ω>1.

Here we focus on the off-resonance case, Δ ≠ 0. As in the case of two-level atoms, we also assume that the conditions for Rabi resonances are guaranteed, Δc1 = −Δc2 = d. The total free Hamiltonian H0 of atom-field system is written as H0 = Ha + Hc. At the same time we write the interaction Hamiltonian HI in terms of the dressed atomic states, and make a further unitary transformation with U2 = exp (−iH0t/ħ) and a rotating-wave approximation. Then we obtain the interaction Hamiltonian

HI=12g[cos2θ(a1a2)+sinθ(1+sinθ)a3+sinθ(1sinθ)a4]σ+0+12g[cos4θ(a3a4)+sinθ(1sinθ)a1+sinθ(1+sinθ)a2]σ0+H.c.,
where we have also dropped the symbol δ for simplicity and clarity. The atom-field interaction is pictorially shown in Fig. 6. In what follows we will be going to discuss the effects of the four-mode atom-field interaction on the quantum correlations. It is seen from Eq. (29) that all of the four cavity fields a1–4 are simultaneously or collectively coupled to each of the dressed atomic transitions |0〉 ↔ |±〉. The simultaneous or collective interactions of the four cavity fields with the dressed atoms establish the cascaded interactions between the different cavity modes. We describe such interactions as follows.

 figure: Fig. 6

Fig. 6 The dressed atomic transitions for N0 > N±. The left blue and red lines with arrows mean respectively the annihilation of the a1,2 modes and the creation of the a3,4 modes and constitute the collective b1 mode. The right red and blue lines with arrows show the annihilation of the a3,4 modes and the creation of the a1,2 modes and constitute the collective b2 mode.

Download Full Size | PDF

We define four Bogoliubov-like modes for the parameter regime (i) |Δ|Ω¯<1 as follows,

b1=p(a1a2)ua3va4,b2=p(a3a4)+va1+ua2,b3=wa3+za4q(a1a2),b4=za1+wa2+q(a3a4),
where we have used the parameters p=cosθ/2(12tan2θ), q=psinθ/cos(2θ), u = p tan θ(1 + sin θ)/ cos θ, v = p tan θ(1 − sin θ)/ cos θ, w = q[1 − tan2 θ(1 − sin θ)]/ sin θ, and z = q[1 − tan2 θ(1 + sin θ)]/ sin θ. These new modes are independent of each other, i.e., [bj, bk] = 0, [bj,bk]=δjk, j, k = 1 – 4. Then the Hamiltonian in Eq. (29) is rewritten as
H=G(b1σ+0+b2σ0)+H.c.,
where G=g/(2p) denotes the strength for the interactions of the Bogoliubov modes with the dressed atoms. For (ii) 1<|Δ|Ω¯<2, with substitutions of cos2 θ − 2 sin2 θ and b1–4 for −(cos2 θ − 2 sin2 θ) and b14 respectively in Eqs. (30) and (31), the Hamiltonian is changed into H = ħG(−b1σ0+ + b2σ0−) + H.c.. For (iii) |Δ|Ω¯>2, the Hamiltonian and b1,2 are the same as in case (ii) except for the substitutions of b3,4 for −b3,4. Pictorially, the interactions of the cavity fields with the dressed atoms are shown in Fig. 7(a). Depending on the atomic populations, either the transitions from |0〉 to |±〉 or the reverse transitions are accompanied with the absorption of the b1,2 modes respectively. At the same time, the other two Bogoliubov-like modes b3,4 are decoupled from the atoms. In what follows our derivations are given only for the region (i), and the other regions can be treated in the same way.

 figure: Fig. 7

Fig. 7 (a) Interactions of the collective b1,2 modes with the dressed atoms via the transitions (i) from |0〉 to |±〉 (N0 > N±) (ii) from |±〉 to |0〉 (N± > N0). (b) The interactions between any two individual fields. All four fields are in a quadrilateral loop of two-photon interactions and in the X lines of two quantum-beat interactions. Any three fields are in a triangle of two two-photon interactions and one quantum-beat interaction.

Download Full Size | PDF

3.2. Dissipation and four-mode squeezing

In what follows we use the above Hamiltonian to discuss the four-mode correlations. We first show a quadrilateral loop of two-photon processes, then give the apprroximately analytic solution, and finally present our numerical results.

3.2.1. Dissipation and a quadrilateral of two-photon processes

We assume that the atomic variables decay much more rapidly than the cavity fields, i.e., γκ. Then by tracing out the atomic variables, we derive the master equation for the reduced density operator of the cavity modes ρ̃ = Tra ρ as ρ˜˙=𝒦aρ˜+cρ˜, where 𝒦aρ˜=iTra[HI,ρ] describes the atom-field interaction. Following the standard technique [2,3], we derive the atomic contribution part for (i) |Δ|Ω¯<1 as

𝒦aρ˜=A(b1ρ˜b1b1b1ρ˜+b2ρ˜b2b2b2ρ˜)+B(b1ρ˜b1b1b1ρ˜+b2ρ˜b2b2b2ρ˜)+D(b1ρ˜b2+b2ρ˜b1)D1(ρ˜b1b2+ρ˜b2b1)D2(b1b2ρ˜+b2b1ρ˜)+H.c.,
where A = G2ΓN0/Γ̃2, B = G2ΓN+/Γ̃2, D1 = G2γc N+/̃Γ2, D2 = G2γc N0/Γ̃2, D = D1 + D2, Γ=γ(112cos4θ), γc=γ8sin2(2θ), and Γ˜2=Γ2γc2. The A’s terms stand for the absorption of b1,2 by the dressed atoms, the B’s terms represent the gain, and the D’s terms are due to the coherence transfer between the degenerate dressed atomic transitions, as shown in Fig. 6. These coefficients satisfy the relations (D, D1,2) ≪ (A, B) due to γc ≪ Γ.

When AB ≫ (κ, γc), i.e., the absorption by the atoms dominates [3], and thus the collective modes b1,2 undergo dissipation and reduce to the thermal vacuum states. This is the very effect of the dissipation by the dressed atoms. Expanding the dominant terms (those proportional to A and B), we can find two types of interactions between the different individual modes ak,l. One type is the two-photon interactions as in the squeezed vacuum reservoir case [2,3] and described by the terms

β1akρ˜al+β2ρ˜akal+β3akalρ˜+(kl)+H.c.,
where kl = 13, 14, 23, 24. The parameters β1–3 depend on the absorption and gain coefficients A and B and are not necessarily written out here. The other type is the quantum-beat interactions as in a quantum-beat laser [2,3] and described by the terms
β1akρ˜al+β2ρ˜alak+β3alakρ˜+(kl)+H.c.,
where kl = 12, 34. The parameters β′1–3, like β1–3, are also dependent on the absorption and gain coefficients A and B. These interactions are interlinked to each other in the characteristic ways as follows.
  1. All four modes are in a quadrilateral loop of two-photon interactions, as shown in Fig. 7(b). We use ”𝕋” for two-photon interactions and ”𝔹” for quantum-beat interactions. The quadrilateral loop of two-photon processes is
    a1𝕋a3𝕋a2𝕋a4𝕋a1.
    At the same time, the two pairs of modes on the diagonal lines are respectively in the quantum-beat interactions
    a1𝔹a2,a3𝔹a4.
  2. Any three of the four cavity fields are in interactions with each other in a triangle [Fig. 7(b)], in which both lines correspond to the two-photon interactions and the third one stands for the quantum-beat interaction. There are four such triangles for the four-mode system
    a1𝕋a3𝕋a2𝔹a1,a1𝕋a4𝕋a2𝔹a1,a3𝕋a1𝕋a4𝔹a3,a3𝕋a2𝕋a4𝔹a3.

The loop interactions are the most remarkable feature of the multimode atom-field interactions. All of these interactions happen either in the absorption of collective modes (described by the A term in Eq. (32)) or in the amplification of the collective modes (described by the B term) [2, 3]. When the absorption dominates over the amplification, the four-mode system will evolve into a multimode steady state. As will be shown below, the dissipation based two-photon processes will determine the existence of four-mode squeezing.

3.2.2. Approximately analytic four-mode variances

In order to show the dissipation effects on the multimode correlations, we rewrite the multimode annihilation operator b1 in terms of collective annihilation and creation operators as

b1=coshra1a22ξsinhrua3+va4u2+v2,
where we have defined the two-collective-mode squeezing parameter
r=arctanh|sinθ1+sin2θ|cos2θfor|Δ|Ω¯<1,r=arctanhcos2θ|sinθ1+sin2θ|for|Δ|Ω¯>1,
and the step function ξ = ξ(Δ) = 1 for Δ > 0, and ξ = −1 for Δ < 0. In what follows we show that the two-collective-mode squeezing parameter r will determine the four-mode squeezing.

To show this clearly, we first consider the dominant terms (the A and B terms in Eq. (32)) and neglect temporarily those minor terms (the cavity relaxations and the coherence transfer). Under such conditions, we can give the approximately analytic expression of the four-mode quantum correlations. We introduce the quadrature operators δxo=12(o+o) and δpo=i2(oo), o = a1–4. By using the reversible transformation from Eq. (30), we can express the individual modes a1–4 (or b1–4) in terms of collective modes b1–4. On the basis of the above two-collective-mode Bogolubov transform in Eq. (38) we focus on the four-mode operator

X=xa1xa22ξuxa3+vxa4u2+v2
and express its variance as
(δX2)=e2r(1+:(δxb1)2:+sin2θ1+sin2θ:(δxb2)2:),
where we have used the fact that the modes b3,4 are decoupled from the atoms and remain in vacuum state. Due to the dissipation by the atoms, the b1,2 modes evolve into the thermal vacuum states. From the master equation we can obtain nonvanishing correlations for the b1,2 modes b1b1=b2b2=N+N0N+. Thus we have :(δxbj)2:=12bj2+bj2+2bjbj=bjbj, j = 1, 2. The four-mode correlation is obtained for |Δ|Ω<1 as
(δX)2=e2r(1+1+2sin2θ1+sin2θN+N0N+).
It is clear to see that the variance depends on the scaled detuning ΔΩ¯. When Δ = 0 (N+ = 0, r = 0), the variance equals the standard quantum limit, i.e., 〈(δX)2〉 = 1, which corresponds to absence of the multimode squeezing. Only in the off-resonant situation can the multimode squeezing exist. Similarly, for |Δ|Ω>1 we obtain the variance as
(δX)2=e2r(1+1+2sin2θ1+sin2θN0N+N0).

If the variance satisfies the relation 〈(δX)2〉 < 1, the multimode squeezing exists. We take Φ = 0 and plot the variance 〈(δX)2〉 versus the scaled detuning ΔΩ in Fig. 8. The four-mode squeezing exists in the almost entire regime except at the specific points, i.e., |Δ|Ω=(0,1,2). At |Δ|Ω=0, the atoms are trapped in the dark state and the cavity fields are decoupled from the atoms. No squeezing occurs at the exact dark resonance. At |Δ|Ω=2, the two-collective-mode squeezing parameter is not well defined. This problem is avoided when we treat the individual modes, because this definition is no longer necessary. At |Δ|Ω=1, the atoms are equally populated in three dressed states i.e., N0 = N±, the absorption is equal to the gain and no net dissipation occurs. In fact, about |Δ|Ω=1 (as shown by shadows in Fig. 8), the condition ABκ is not well met, i.e., the absorption by the atoms does no more play a dominant role. The approximately analytic solution does not hold about |Δ|Ω=1. As will be shown in the following subsection, the best achievable squeezing is 50% and appears still near |Δ|Ω=1. We should also point out that we can obtain the same variance for other collective modes

P=pa1pa22+ξupa3+vpa4u2+v2,X=xa3xa42+ξvxa1+uxa2u2+v2,P=pa3pa42ξvpa1+upa2u2+v2,
which are defined in terms of the collective annihilation and creation operators.

 figure: Fig. 8

Fig. 8 Approximately analytic four-mode variance 〈(δX)2〉 versus the scaled detuning ΔΩ for a fixed phase Φ = 0. Four-mode squeezing occurs in almost entire region except at |Δ|Ω=0,1,2. Within two shadows at |Δ|Ω=1, the conditions for approximately analytic solution are not well met.

Download Full Size | PDF

3.2.3. Numerical verification and discussion

Now we present a numerical calculation by including the vacuum damping and the coherence transfer. We can apply a standard Langevin approach by means of the generalized P representation [54–56]. The individual operators are arranged in the normal ordering a1, a2, a3, a4, a4, a3, a2, a1, and the correspondences between the c numbers and the operators are defined as αjaj, αj*aj (j = 1 – 4). The c number Langevin equations for individual modes are derived as

α˙1=λ11α1+λ12α2+λ13α3*+λ14α4*+Fα1,α˙2=λ21α1+λ22α2+λ23α3*+λ24α4*+Fα2,α˙3=λ31α1*+λ32α2*+λ33α3+λ34α4+Fα3,α˙4=λ41α1*+λ42α2*+λ43α3+λ44α4+Fα4,
where we have defined the parameters
λ11=λ44=κ/2(AB)β1,λ22=λ33=κ/2(AB)β2,λ23=λ32=(D1D2)β2,λ14=λ41=(D1D2)β1,λ13=λ42=(AB)cosθsin(2θ)+(D1D2)cos2θ,λ31=λ24=(AB)cosθsin(2θ)+(D1D2)cos2θ,λ12=λ43=(AB)cos2θ+(D1D2)cosθsin(2θ),λ21=λ34=(AB)cos2θ(D1D2)cosθsin(2θ),
with β1 = (1 − sin θ)2(1 + 2 sin θ) and β2 = β1(−θ). The F’s are noise terms with zero means and correlations 〈Fo (t)Fo′ (t′)〉 = 2Doo′ δ(tt′). The nonvanishing diffusion coefficients can be calculated from the reduced master equation for individual modes as follows
Dα1*α1=Aη12+Bcos4θDη1cos2θ,Dα3*α3=Aη22+Bcos4θ+Dη2cos2θ,Dα1α4=(A+B)η1cos2θD1cos4θD2η12,Dα1α3=(A+BD2)cos2θsin2θ+D1cos4θ,Dα2α3=(A+B)η2cos2θD1cos4θD2η22,Dα1α2*=(Asin2θBcos2θ)cos2θDcos2sin2θ,Dα2*α2=Dα3*α3,Dα4*α4=Dα1*α1,Dα2α4=Dα1α3,Dα3α4*=Dα1α2*,
where η1 = sin θ(1 − sin θ) and η2 = sin θ(1 + sin θ). From Eqs. (45)(47) we derive the set of equations for quantum correlations ddtoo=dodto+ododt+2Doo, o, o′ = aj, αj*, j = 1–4. By setting time derivatives at left side of equations to zero and making a stability analysis, we obtain the steady state solutions.

In Fig. 9 we show the numerical results for different rates of cavity losses. It is seen that the four-mode squeezing is obtainable for the almost entire regime except at |Δ|Ω=0,1. This is in a good agreement with the analytic calculation. A difference is the pair of reverse dips at |Δ|Ω=1. About these two positions, the populations trend to be the same, N0N±. The absorption term proportional to A is no longer dominant over the amplification term proportional to B. The condition ABκ is not met, i.e., the absorption (A) by the atoms is not sufficient to suppress the amplification (B) and the vacuum fluctuations (κ). Therefore, as |Δ|Ω trends to 1, the dissipation effects are greatly degraded. However, the best achievable squeezing is about 50% and appears still close to |Δ|Ω=1. In addition, the requirement of |Δ|Ω2 is no longer needed since we directly use the individual modes. Except at those specific points, the other entire region is possible for the occurrence of the squeezing. In addition, we show in Fig. 10 the phase dependence of the four-mode correlation when the condition Ω̄ ≫ (γ, κ) is well satisfied. The squeezing is existent in the entire regime except at |Δ|Ω¯=0,1 (i.e., at |Δ|Ω=0,12|1+eiΦ|).

 figure: Fig. 9

Fig. 9 The numerical four-mode variance 〈(δX)2〉 versus the scaled detuning ΔΩ for different rates of cavity losses κ = 0.2γ (red dashed), κ = 0.1γ (black dotted), κ = 0.01γ (blue solid). The parameters are chosen as gN=10γ and Φ = 0.

Download Full Size | PDF

 figure: Fig. 10

Fig. 10 The numerical four-mode variance 〈(δX)2〉 versus the scaled detuning ΔΩ for varying phases: Φ = 0 (blue dashed), Φ=π2 (red dotted), Φ=2π3 (black solid). The chosen parameters are gN=10γ and κ = 0.01γ.

Download Full Size | PDF

So far it has been understood that the dissipation-dominant two-photon interactions are responsible for the above four-mode squeezing. Finally, we should note the almost entire parameter range for squeezing except several specific points. This is in sharp contrast to the monochromatic case [20], where the dissipation mechanism does not happen for 1<ΔΩ<23 and ΔΩ>1, or for 23<ΔΩ<1 and ΔΩ<1 and so no squeezing occurs. We explain this remarkable difference by comparing with the correlation of a multimode quantity

Y=12[xa1xa2ξ(xa3+xa4)],
where ξ(Δ) is dependent on the sign of Δ. We plot the variance 〈(δY)2〉 in Fig. 11. By comparing with Fig. 10 we note three differences as follows. (i) All curves in Fig. 11 shift upward and the best squeezing reduces to 37%. (ii) There two symmetric parameter regimes in which no squeezing occurs. (iii) The range of ΔΩ for the four-mode squeezing changes as the phase Φ is varied. We note that the strengths of the two-photon processes akal (kl = 13, 14, 23, 24) are strongly dependent on the parameter ΔΩ. The Bogoliubov-like mode dissipation is best for the squeezing of two collective modes a1a22 and ua3+va4u2+v2 (or a3a42 and ua1+va2u2+v2). In other words, u and v are optimization parameters for the four-mode squeezing of the four-mode operator X as in Eqs. (40) and (44). It is for this reason that the squeezing degree and the ΔΩ region for squeezing are both reduced to a certain degree.

 figure: Fig. 11

Fig. 11 The numerical four-mode variance 〈(δY)2〉 versus the scaled detuning ΔΩ for the same parameters as in Fig. 10.

Download Full Size | PDF

4. Conclusion

We have revealed that the Bogoliubov mode dissipation is responsible for the noise squeezing of fields under bichromatic excitation. The two-level atomic system is used as a fundamental model, and the three-level atomic system is considered for cascade bichromatic interactions. By transferring the atom-field nonlinearities to the dressed atoms we separate out the dissipative interactions of Bogoliubov modes with dressed-atoms. The dissipative interactions lead to the two-photon processes of two involved fields in the two-level system, and to a quadrilateral loop of the two-photon processes of four involved fields in the three-level system. Therefore it is the dissipation based two-photon interactions that result in the two-mode squeezing and the four-mode squeezing in the above two systems, respectively.

Funding

National Natural Science Foundation of China (NSFC) (Grants Nos. 11474118, 61178021, and 11204099); National Basic Research Program of China (Grant No. 2012CB921604).

References and links

1. C. M. Caves, K. S. Thorne, R. W. P. Drever, V. D. Sandberg, and Zimmermann, “On the rrieasureri-ient of a weak classical force coupled to a quantum-mechanical oscillator. I. Issues of principle,” Rev. Mod. Phys. 52, 341–392 (1980). [CrossRef]  

2. D. F. Walls and G. J. Milburn, Quantum Optics (Springer, 1995).

3. M. O. Scully and M. S. Zubairy, Quantum Optics (Cambridge University, 1997).

4. A. Einstein, B. Podolsky, and N. Rosen, “Can quantum-mechanical description of physical reality be considered complete?” Phys. Rev. 47, 777–780 (1935). [CrossRef]  

5. 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]  

6. S. L. Braunstein and P. van Loock, “Quantum information with continuous variables,” Rev. Mod. Phys. 77, 513–577 (2005). [CrossRef]  

7. R. Horodecki, P. Horodecki, M. Horodecki, and K. Horodecki, “Quantum entanglement,” Rev. Mod. Phys. 81, 865–942 (2009). [CrossRef]  

8. M. D. Reid, D. F. Walls, and B. J. Dalton, “Squeezing of quantum fluctuations via atomic coherence effects,” Phys. Rev. Lett. 55, 1288–1290 (1985). [CrossRef]   [PubMed]  

9. R. Guzmán, J. C. Retamal, E. Solano, and N. Zagury, “Field squeeze operators in optical cavities with atomic ensembles,” Phys. Rev. Lett. 96, 010502 (2006). [CrossRef]   [PubMed]  

10. V. A. Sautenkov, Y. V. Rostovtsev, and M. O. Scully, “Switching between photon-photon correlations and Raman anticorrelations in a coherently prepared Rb vapor,” Phys. Rev. A 72, 065801 (2005). [CrossRef]  

11. A. Dantan, J. Cviklinski, E. Giacobino, and M. Pinard, “Spin squeezing and light entanglement in coherent population trapping,” Phys. Rev. Lett. 97, 023605 (2006). [CrossRef]   [PubMed]  

12. A. Sinatra, “Quantum correlations of two optical fields close to electromagnetically induced transparency,” Phys. Rev. Lett. 97, 253601 (2006). [CrossRef]  

13. A. Sinatra, J. F. Roch, K. Vigneron, Ph. Grelu, J. Ph. Poizat, K. Wang, and P. Grangier, “Quantum-nondemolition measurements using cold trapped atoms: comparison between theory and experiment,” Phys. Rev. A 57, 2980–2995 (1998). [CrossRef]  

14. P. Barberis-Blostein and N. Zagury, “Field correlations in electromagnetically induced transparency,” Phys. Rev. A 70, 053827 (2004). [CrossRef]  

15. Y. Wu, M. G. Payne, E. W. Hagley, and L. Deng, “Preparation of multiparty entangled states using pairwise perfectly efficient single-probe photon four-wave mixing,” Phys. Rev. A 69, 063803 (2004). [CrossRef]  

16. X. Liang, X. M. Hu, and C. He, “Creating multimode squeezed states and Greenberger-Horne-Zeilinger entangled states using atomic coherent effects,” Phys. Rev. A 85, 032329 (2012). [CrossRef]  

17. S. Pielawa, G. Morigi, D. Vitali, and L. Davidovich, “Generation of Einstein-Podolsky-Rosen-entangled radiation through an atomic reservoir,” Phys. Rev. Lett. 98, 240401 (2007). [CrossRef]   [PubMed]  

18. P. B. Li, “Generation of two-mode field squeezing through selective dynamics in cavity QED,” Phys. Rev. A 77, 015809 (2008). [CrossRef]  

19. G. L. Cheng, X. M. Hu, W. X. Zhong, and Q. Li, “Two-channel interaction of squeeze-transformed modes with dressed atoms: entanglement enhancement in four-wave mixing in three-level systems,” Phys. Rev. A 78, 033811 (2008). [CrossRef]  

20. X. M. Hu, “Entanglement generation by dissipation in or beyond dark resonances,” Phys. Rev. A 92, 022329 (2015). [CrossRef]  

21. A. S. Parkins, E. Solano, and J. I. Cirac, “Unconditional two-mode squeezing of separated atomic ensembles,” Phys. Rev. Lett. 96, 053602 (2006). [CrossRef]   [PubMed]  

22. E. G. Dalla Torre, J. Otterbach, E. Demler, V. Vuletic, and M. D. Lukin, “Dissipative preparation of spin squeezed atomic ensembles in a steady state,” Phys. Rev. Lett. 110, 120402 (2013). [CrossRef]   [PubMed]  

23. H. Krauter, C. A. Muschik, K. Jensen, W. Wasilewski, J. M. Petersen, J. I. Cirac, and E. S. Polzik, “Entanglement generated by dissipation and steady state entanglement of two macroscopic objects,” Phys. Rev. Lett. 107, 080503 (2011). [CrossRef]   [PubMed]  

24. M. O. Scully, K. Wódkiewicz, M. S. Zubairy, J. Bergou, N. Lu, and J. Myer ter Vehn, “Two-Photon Correlated-Spontaneous-Emission Laser: Quantum Noise Quenching and Squeezing,” Phys. Rev. Lett. 60, 1832–1835 (1988). [CrossRef]   [PubMed]  

25. X. Zhang and X. M. Hu, “Entanglement between collective fields via atomic coherence effects,” Phys. Rev. A 81, 013811 (2010). [CrossRef]  

26. B. Blind, P. R. Fontana, and P. Thomann, “Resonance fluorescence spectrum of intense amplitude modulated laser light,” J. Phys. B: At. Mol. Opt. Phys. 13, 2717–2727 (1980). [CrossRef]  

27. Y. Zhu, Q. Wu, A. Lezama, D. J. Gauthier, and T. W. Mossberg, “Resonance fluorescence of tvvo-level atoms under strong bichromatic excitation,” Phys. Rev. A 41, R6574–R6576 (1990). [CrossRef]  

28. C. C. Yu, J. R. Bochinski, T. M. V. Kordich, T. W. Mossberg, and Z. Ficek, “Driving the driven atom: spectral signatures,” Phys. Rev. A 56, R4381–R4384 (1997). [CrossRef]  

29. G. S. Agarwal, Y. Zhu, D. J. Gauthier, and T. W. Mossberg, “Spectrum of radiation from two-level atoms under intense bichromatic excitation,” J. Opt. Soc. Am. B 8, 1163–1167 (1991). [CrossRef]  

30. Z. Ficek and H. S. Freedhoff, “Resonance-Auorescence and absorption spectra of a two-level atom driven by a strong bichromatic field,” Phys. Rev. A 48, 3092–3104 (1993). [CrossRef]   [PubMed]  

31. Z. Ficek and T. Rudolph, “Quantum interference in a driven two-level atom,” Phys. Rev. A 60, R4245–R4248 (1999). [CrossRef]  

32. Z. Ficek, J. Seke, A. V. Soldatov, and G. Adam, “Fluorescence spectrum of a two-level atom driven by a multiple modulated field,” Phys. Rev. A 64, 013813 (2001). [CrossRef]  

33. Z. Ficek, J. Seke, A. V. Soldatov, and G. Adam, “Phase control of subharmonic resonances,” Opt. Commun. 182, 143–150 (2000). [CrossRef]  

34. Z. Ficek, J. Seke, A. V. Soldatov, G. Adam, and N. N. Bogoubov Jr, “Multilevel coherence effects in a two-level atom driven by a trichromatic field,” Opt. Commun. 217, 299–309 (2003). [CrossRef]  

35. D. L. Aronstein, R. S. Bennink, R. W. Boyd, and C. R. Stroud Jr., “Comment on Resonance-fluorescence and absorption spectra of a two-level atom driven by a strong bichromatic field,” Phys. Rev. A 65, 067401 (2002). [CrossRef]  

36. Yu. A. Zinin and N. V. Sushilov, “Absorption and dispersion spectra of a polychromatic field in a two-level medium driven by a strong polychromatic pumping field,” Phys. Rev. A 51, 3916–3921 (1995). [CrossRef]   [PubMed]  

37. M. F. Van Leeuwen, S. Papademetriou, and C. R. Stroud Jr., “Autler-Townes effect for an atom in a 100% amplitude-modulated laser field. I. a dressed-atom approach,” Phys. Rev. A 53, 990–996 (1996). [CrossRef]   [PubMed]  

38. S. Papademetriou, M. F. Van Leeuwen, and C. R. Stroud Jr., “Autler-Townes effect for an atom in a 100% amplitude-modulated laser field. II. experimental results,” Phys. Rev. A 53, 997–1003 (1996). [CrossRef]   [PubMed]  

39. A. D. Greentree, C. Wei, and N. B. Manson, “Polychromatic excitation of a two-level system,” Phys. Rev. A 59, 4083–4086 (1999). [CrossRef]  

40. T. H. Yoon, M. S. Chung, and H. W. Lee, “Absorption spectra of two-level atoms interacting with a strong polychromatic pump field and an arbitrarily intense probe field,” Phys. Rev. A 60, 2547–2553 (1999). [CrossRef]  

41. T. H. Yoon, S. A. Pulkin, J. R. Park, M. S. Chung, and H. W. Lee, “Theoretical analysis of resonances in the polarization spectrum of a two-level atom driven by a polychromatic field,” Phys. Rev. A 60, 605–613 (1999). [CrossRef]  

42. J. Wang, Y. Zhu, K. J. Jiang, and M. S. Zhan, “Bichromatic electromagnetically induced transparency in cold rubidium atoms,” Phys. Rev. A 68, 063810 (2003). [CrossRef]  

43. X. M. Hu, J. H. Zou, X. Li, D. Du, and G. L. Cheng, “Amplitude and phase control of trichromatic electromagnetically induced transparency,” J. Phys. B: At. Mol. Opt. Phys. 38, 683–692 (2005). [CrossRef]  

44. X. M. Hu, G. L. Cheng, J. H. Zou, X. Li, and D. Du, “Double switching from normal to anomalous dispersion via trichromatic phase manipulation of electromagnetically induced transparency,” Phys. Rev. A 72, 023803 (2005). [CrossRef]  

45. P. B. Sellin, C. C. Yu, J. R. Bochinski, and T. W. Mossberg, “Intrinsically irreversible multiphoton laser gain mechanisms,” Phys. Rev. Lett. 78, 1432–1435 (1997). [CrossRef]  

46. J. R. Bochinski, C. C. Yu, T. Loftus, and T. W. Mossberg, “Vacuum-mediated multiphoton transitions,” Phys. Rev. A 63, 051402(R) (2001). [CrossRef]  

47. J. H. Zou, X. M. Hu, G. L. Cheng, X. Li, and D. Du, “Inhibition of two-photon absorption in a three-level system with a pair of bichromatic fields,” Phys. Rev. A 72, 055802 (2005). [CrossRef]  

48. W. Lange, H. Walther, and G. S. Agarwal, “Decay of bichromatically driven atoms in a cavity,” Phys. Rev. A 50, R3593–R3596 (1994). [CrossRef]  

49. X. M. Hu, Q. Xu, J. Y. Li, X. X. Li, W. X. Shi, and X. Zhang, “Bichromatic and trichromatic manipulation of spontaneous emission in a three-level system,” Opt. Commun. 260, 196–202 (2006). [CrossRef]  

50. G. Yu. Kryuchkyan, M. Jakob, and A. S. Sargsian, “Resonance fluorescence in a bichromatic field as a source of nonclassical light,” Phys. Rev. A 57, 2091–2095 (1998). [CrossRef]  

51. M. Jakob and G. Yu. Kryuchkyan, “Squeezing in the resonance fluorescence of a bichromatically driven two-level atom,” Phys. Rev. A 58, 767–770 (1998). [CrossRef]  

52. G. S. Agarwal, W. Lange, and H. Walther, “Intense-field renormalization of cavity-induced spontaneous emission,” Phys. Rev. A 48, 4555–4568 (1993). [CrossRef]   [PubMed]  

53. C. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg, Atom-Photon Interactions (Wiley, 1992).

54. C. W. Gardiner and P. Zoller, Quantum Noise, 2nd ed. (Springer, 2000). [CrossRef]  

55. P. D. Drummond and C. W. Gardiner, “Generalised P-representations in quantum optics,” J. Phys. A 13, 2353–2368 (1980). [CrossRef]  

56. P. D. Drummond and D. F. Walls, “Quantum theory of optical bistability. II. atomic fluorescence in a high-Q cavity,” Phys. Rev. A 23, 2563–2579 (1981). [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 (11)

Fig. 1
Fig. 1 (a) An ensemble of two-level atoms is placed at the intersection of two cavities, into which two external coherent fields are input respectively. (b) The two cavity fields (annihilation operators a1,2) of different frequencies are coupled to the common transition of the atoms.
Fig. 2
Fig. 2 The individual cavity fields a1,2 are both in resonant interactions with the dressed atoms. The two fields combine into the Bogoliubov mode b1 for Δ > 0 (N > N+) or b2 for Δ < 0 (N+ > N).
Fig. 3
Fig. 3 Approximately analytic variance 〈(δX)2〉 versus the scaled detuning Δ Ω . Within the shadow at Δ Ω = 0 , the conditions for approximately analytic solution are not well met.
Fig. 4
Fig. 4 The variance 〈(δX)2〉 versus the scaled detuning Δ Ω for different rates of cavity losses κ = 0.2γ (red dashed), κ = 0.1γ (black dotted), and κ = 0.01γ (blue solid). We have used the other parameter g N = 10 γ .
Fig. 5
Fig. 5 (a) Four optical cavities intersect at a common spot, where an atomic ensemble of three-level atoms is placed. (b) The interaction of the four cavity fields with the atoms in Λ configuration.
Fig. 6
Fig. 6 The dressed atomic transitions for N0 > N±. The left blue and red lines with arrows mean respectively the annihilation of the a1,2 modes and the creation of the a3,4 modes and constitute the collective b1 mode. The right red and blue lines with arrows show the annihilation of the a3,4 modes and the creation of the a1,2 modes and constitute the collective b2 mode.
Fig. 7
Fig. 7 (a) Interactions of the collective b1,2 modes with the dressed atoms via the transitions (i) from |0〉 to |±〉 (N0 > N±) (ii) from |±〉 to |0〉 (N± > N0). (b) The interactions between any two individual fields. All four fields are in a quadrilateral loop of two-photon interactions and in the X lines of two quantum-beat interactions. Any three fields are in a triangle of two two-photon interactions and one quantum-beat interaction.
Fig. 8
Fig. 8 Approximately analytic four-mode variance 〈(δX)2〉 versus the scaled detuning Δ Ω for a fixed phase Φ = 0. Four-mode squeezing occurs in almost entire region except at | Δ | Ω = 0 , 1 , 2 . Within two shadows at | Δ | Ω = 1 , the conditions for approximately analytic solution are not well met.
Fig. 9
Fig. 9 The numerical four-mode variance 〈(δX)2〉 versus the scaled detuning Δ Ω for different rates of cavity losses κ = 0.2γ (red dashed), κ = 0.1γ (black dotted), κ = 0.01γ (blue solid). The parameters are chosen as g N = 10 γ and Φ = 0.
Fig. 10
Fig. 10 The numerical four-mode variance 〈(δX)2〉 versus the scaled detuning Δ Ω for varying phases: Φ = 0 (blue dashed), Φ = π 2 (red dotted), Φ = 2 π 3 (black solid). The chosen parameters are g N = 10 γ and κ = 0.01γ.
Fig. 11
Fig. 11 The numerical four-mode variance 〈(δY)2〉 versus the scaled detuning Δ Ω for the same parameters as in Fig. 10.

Equations (48)

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

ρ ˙ = i [ H , ρ ] + ρ ,
H = { Δ σ 11 + [ ( g 1 a 1 + g 2 a 2 ) σ 21 + H . c . ] } + j = 1 2 [ Δ c j a j a j + i ( ε j a j ε j * a j ) ] ,
ρ = γ 2 σ 12 ρ + j = 1 2 κ j 2 a j ρ ,
I in = I [ ( 1 + 𝒜 ) 2 + ( Δ ¯ c + 𝒟 ) 2 ] ,
𝒜 = 8 C 4 Δ ¯ 2 + 32 I + 1 , 𝒟 = 16 C Δ ¯ 4 Δ ¯ 2 + 32 I + 1 .
H = H a + H c + H I ,
H a = Δ σ 22 + 2 ( Ω σ 21 + Ω * σ 12 )
H c = Δ c ( δ a 1 δ a 1 δ a 2 δ a 2 )
| + = sin θ | 1 + cos θ | 2 | = cos θ | 1 sin θ | 2 ,
N + = N sin 4 θ cos 4 θ + sin 4 θ , N = N cos 4 θ cos 4 θ + sin 4 θ .
H I = g ( a 1 cos 2 θ a 2 sin 2 θ ) σ + + H . c . ,
b 1 = a 1 cosh r a 2 sinh r , b 2 = a 2 cosh r a 1 sinh r ,
H I = G ( b 1 σ + + σ + b 1 ) for Δ > 0 , H I = G ( b 2 σ + + σ + b 2 ) for Δ < 0 ,
ρ ˜ ˙ = 𝒦 a ρ ˜ + c ρ ˜ ,
𝒦 a ρ ˜ = A ( 2 b 1 ρ ˜ b 1 b 1 b 1 ρ ˜ ρ ˜ b 1 b 1 ) + B ( 2 b 1 ρ ˜ b 1 b 1 b 1 ρ ˜ ρ ˜ b 1 b 1 ) ,
β 1 a 1 ρ ˜ a 2 + β 2 ρ ˜ a 1 a 2 + β 3 a 1 a 2 ρ ˜ + ( 1 2 ) + H . c . ,
δ X = δ x a 1 δ x a 2 = e 4 ( δ x b 1 δ x b 2 ) , δ P = δ p a 1 + δ p a 2 = e r ( δ p b 1 + δ p b 2 ) .
( δ X ) 2 = e 2 r ( 1 + : ( δ x b 1 ) 2 : ) , ( δ P ) 2 = e 2 r ( 1 + : ( δ p b 1 ) 2 : ) ,
( δ X ) 2 = ( δ P ) 2 = e 2 r N + N N + for Δ > 0 , ( δ X ) 2 = ( δ P ) 2 = e 2 r N N + N for Δ < 0 .
α ˙ 1 = λ 11 α 1 + λ 12 α 2 * + F α 1 , α ˙ 2 = λ 21 α 1 * + λ 22 α 2 + F α 2 ,
2 D α 1 * α 1 = 2 g 2 Γ N + cos 4 θ , 2 D α 1 * α 2 = 2 g 2 Γ N sin 4 θ , 2 D α 1 α 2 = g 2 N 4 Γ sin 2 ( 2 θ ) , 2 D α 1 * α 2 * = 2 D α 1 α 2 .
H = j = 1 2 { Δ j σ j j + [ ( g j a j + g j + 2 a j + 2 ) σ 3 j + H . c . ] } + j = 1 4 [ Δ c j a j a j + i ( ε j a j ε j * a j ) ] ,
ρ = j = 1 2 γ j 2 σ j 3 ρ + j = 1 4 κ j 2 a j ρ ,
𝒜 = C Δ ¯ 2 Δ ¯ 4 + Δ ¯ 2 + Δ ¯ 2 I + I 2 , 𝒟 = C Δ ¯ ( Δ ¯ 2 + I ) Δ ¯ 4 + Δ ¯ 2 + Δ ¯ 2 I + I 2 .
H a = j = 1 2 [ ( 1 ) j + 1 Δ σ j j + ( Ω j + Ω j + 2 ) σ 3 j + ( Ω j * + Ω j + 2 * ) σ j 3 ]
H c = j = 1 2 Δ c ( δ a j δ a j δ a j + 2 δ a j + 2 )
| + = 1 + sin θ 2 | 1 + 1 sin θ 2 | 2 + cos θ 2 | 3 , | 0 = cos θ 2 | 1 + cos θ 2 | 2 + sin θ | 3 , | = 1 sin θ 2 | 1 + 1 + sin θ 2 | 2 cos θ 2 | 3 ,
N 0 = N cos 4 θ 1 + 3 sin 4 θ , N ± 1 2 ( N N 0 ) .
H I = 1 2 g [ cos 2 θ ( a 1 a 2 ) + sin θ ( 1 + sin θ ) a 3 + sin θ ( 1 sin θ ) a 4 ] σ + 0 + 1 2 g [ cos 4 θ ( a 3 a 4 ) + sin θ ( 1 sin θ ) a 1 + sin θ ( 1 + sin θ ) a 2 ] σ 0 + H . c . ,
b 1 = p ( a 1 a 2 ) u a 3 v a 4 , b 2 = p ( a 3 a 4 ) + v a 1 + u a 2 , b 3 = w a 3 + z a 4 q ( a 1 a 2 ) , b 4 = z a 1 + w a 2 + q ( a 3 a 4 ) ,
H = G ( b 1 σ + 0 + b 2 σ 0 ) + H . c . ,
𝒦 a ρ ˜ = A ( b 1 ρ ˜ b 1 b 1 b 1 ρ ˜ + b 2 ρ ˜ b 2 b 2 b 2 ρ ˜ ) + B ( b 1 ρ ˜ b 1 b 1 b 1 ρ ˜ + b 2 ρ ˜ b 2 b 2 b 2 ρ ˜ ) + D ( b 1 ρ ˜ b 2 + b 2 ρ ˜ b 1 ) D 1 ( ρ ˜ b 1 b 2 + ρ ˜ b 2 b 1 ) D 2 ( b 1 b 2 ρ ˜ + b 2 b 1 ρ ˜ ) + H . c . ,
β 1 a k ρ ˜ a l + β 2 ρ ˜ a k a l + β 3 a k a l ρ ˜ + ( k l ) + H . c . ,
β 1 a k ρ ˜ a l + β 2 ρ ˜ a l a k + β 3 a l a k ρ ˜ + ( k l ) + H . c . ,
a 1 𝕋 a 3 𝕋 a 2 𝕋 a 4 𝕋 a 1 .
a 1 𝔹 a 2 , a 3 𝔹 a 4 .
a 1 𝕋 a 3 𝕋 a 2 𝔹 a 1 , a 1 𝕋 a 4 𝕋 a 2 𝔹 a 1 , a 3 𝕋 a 1 𝕋 a 4 𝔹 a 3 , a 3 𝕋 a 2 𝕋 a 4 𝔹 a 3 .
b 1 = cosh r a 1 a 2 2 ξ sinh r u a 3 + v a 4 u 2 + v 2 ,
r = arctanh | sin θ 1 + sin 2 θ | cos 2 θ for | Δ | Ω ¯ < 1 , r = arctanh cos 2 θ | sin θ 1 + sin 2 θ | for | Δ | Ω ¯ > 1 ,
X = x a 1 x a 2 2 ξ u x a 3 + v x a 4 u 2 + v 2
( δ X 2 ) = e 2 r ( 1 + : ( δ x b 1 ) 2 : + sin 2 θ 1 + sin 2 θ : ( δ x b 2 ) 2 : ) ,
( δ X ) 2 = e 2 r ( 1 + 1 + 2 sin 2 θ 1 + sin 2 θ N + N 0 N + ) .
( δ X ) 2 = e 2 r ( 1 + 1 + 2 sin 2 θ 1 + sin 2 θ N 0 N + N 0 ) .
P = p a 1 p a 2 2 + ξ u p a 3 + v p a 4 u 2 + v 2 , X = x a 3 x a 4 2 + ξ v x a 1 + u x a 2 u 2 + v 2 , P = p a 3 p a 4 2 ξ v p a 1 + u p a 2 u 2 + v 2 ,
α ˙ 1 = λ 11 α 1 + λ 12 α 2 + λ 13 α 3 * + λ 14 α 4 * + F α 1 , α ˙ 2 = λ 21 α 1 + λ 22 α 2 + λ 23 α 3 * + λ 24 α 4 * + F α 2 , α ˙ 3 = λ 31 α 1 * + λ 32 α 2 * + λ 33 α 3 + λ 34 α 4 + F α 3 , α ˙ 4 = λ 41 α 1 * + λ 42 α 2 * + λ 43 α 3 + λ 44 α 4 + F α 4 ,
λ 11 = λ 44 = κ / 2 ( A B ) β 1 , λ 22 = λ 33 = κ / 2 ( A B ) β 2 , λ 23 = λ 32 = ( D 1 D 2 ) β 2 , λ 14 = λ 41 = ( D 1 D 2 ) β 1 , λ 13 = λ 42 = ( A B ) cos θ sin ( 2 θ ) + ( D 1 D 2 ) cos 2 θ , λ 31 = λ 24 = ( A B ) cos θ sin ( 2 θ ) + ( D 1 D 2 ) cos 2 θ , λ 12 = λ 43 = ( A B ) cos 2 θ + ( D 1 D 2 ) cos θ sin ( 2 θ ) , λ 21 = λ 34 = ( A B ) cos 2 θ ( D 1 D 2 ) cos θ sin ( 2 θ ) ,
D α 1 * α 1 = A η 1 2 + B cos 4 θ D η 1 cos 2 θ , D α 3 * α 3 = A η 2 2 + B cos 4 θ + D η 2 cos 2 θ , D α 1 α 4 = ( A + B ) η 1 cos 2 θ D 1 cos 4 θ D 2 η 1 2 , D α 1 α 3 = ( A + B D 2 ) cos 2 θ sin 2 θ + D 1 cos 4 θ , D α 2 α 3 = ( A + B ) η 2 cos 2 θ D 1 cos 4 θ D 2 η 2 2 , D α 1 α 2 * = ( A sin 2 θ B cos 2 θ ) cos 2 θ D cos 2 sin 2 θ , D α 2 * α 2 = D α 3 * α 3 , D α 4 * α 4 = D α 1 * α 1 , D α 2 α 4 = D α 1 α 3 , D α 3 α 4 * = D α 1 α 2 * ,
Y = 1 2 [ x a 1 x a 2 ξ ( x a 3 + x a 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.