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

Collective atomic-population-inversion and stimulated radiation for two-component Bose-Einstein condensate in an optical cavity

Open Access Open Access

Abstract

In this paper we investigate the ground-state properties and related quantum phase transitions for the two-component Bose-Einstein condensate in a single-mode optical cavity. Apart from the usual normal and superradiant phases, multi-stable macroscopic quantum states are realized by means of the spin-coherent-state variational method. We demonstrate analytically the stimulated radiation from a collective state of atomic population inversion, which does not exist in the normal Dicke model with single-component atoms. It is also revealed that the stimulated radiation can be generated only from one component of atoms and the other remains in the ordinary superradiant state. However, the order of superradiant and stimulated-radiation states is interchangeable between two components of atoms by tuning the relative atom-field couplings and the frequency detuning as well.

© 2017 Optical Society of America

1. Introduction

The Dicke model (DM), which describes an ensemble of two-level atoms interacting with a single-mode quantized field [1], plays a important role in the study of Bose-Einstein condensate (BEC) trapped in an optical cavity [2–4]. It successfully illustrates the collective and coherent radiations [1]. A second-order phase transition from a normal phase (NP) to a superradiant phase (SP) was revealed long ago by increase of the atom-field coupling from weak to strong regime [5–7].

In order to realize experimentally the quantum phase transition (QPT) predicted in the DM the collective atom-photon coupling strength ought to be in the same order of magnitude as the energy level-space of atoms. This condition is far beyond atom-field coupling region in the conventional atom-cavity system. Recently the QPT was achieved with a BEC trapped in a high-finesse optical cavity [2–4]. Thus the cavity BEC has been regarded as a promising platform to explore the exotic many-body phenomena [8–21].

It was recently revealed that an extended DM with multi-mode cavity fields [22, 23] exhibits interesting phenomena, which have important applications in quantum information and simulation [24–33]. Moreover both abelian and non-abelian gauge potentials [34] are generated in the two-mode DM, from which the spin-orbit-induced anomalous Hall effect [35] is produced as well. With spatial variation of the atom-photon coupling strength various quantum phases have been predicted such as the crystallization, spin frustration [36], spin glass [37–39] and Nambu-Goldstone mode [40]. It is shown that the strong-coupling [41] may lead to the revival of atomic inversion in a time scale associated with the cavity-field period [42]. The optomechanical DM has been also proposed in order to detect the extremely weak forces [43–48].

Recently the dynamics induced by atom-pair tunneling [49–51] was revealed. The QPT was investigated [52, 53] in two-component BECs by means of the semiclassical approximation. It was demonstrated that coupled two-component BECs in an optical cavity [54] display optical [54], fluid [55], multi-stabilities and capillary instability [55–58]. Substantial many-particle entanglement is also possible in a two-component condensate with spin degree of freedom [59–61] and interference between two BECs has been observed [62]. Particularly, variety of topological excitations is admitted in multi-component and spinor BECs such as domain walls, abelian and non-abelian vortices, monopoles, skyrmions, knots, and D-brane solitons [63].

The QPT in DM has been extensively studied [1–3, 14, 21, 40, 64–67] based on variational method with the help of Holstein-Primakoff transformation [7,14,21,40,64,65,68] to convert the pseudospin operators into a one-mode bosonic operator in the thermodynamic limit. The ground-state properties were also revealed in terms of the catastrophe formalism [69], the dynamic approach [70, 71], and the spin coherent-state variational method [48, 66, 67, 72–74], in which both the normal (⇓) and inverted (⇑) pseudospin [68, 75, 76] can be taken into account giving rise to the multi-stable macroscopic quantum states.

In the present paper, we investigate macroscopic (or collective) quantum states for two-component BECs in a single-mode optical cavity by means of the spin coherent state variational method in order to reveal the rich structure of phase diagrams and the related QPTs. Particularly the collective state of atomic population inversion, namely the inverted pseudospin (⇑), is demonstrated along with the stimulated radiation, which does not exists in the usual DM.

2. Collective population inversion and stimulated radiation beyond the normal and superradiant phases

We consider two ensembles of ultracold atoms, which are coupled simultaneously to an optical cavity mode of frequency ω as depicted in Fig. 1. Effective Hamiltonian of the system has the form [77] of two-component DM in the unit convention ħ = 1,

H=ωaa+l=1,2ωlJlz+l=1,2glNl(a+a)(Jl++Jl).
Where Jlz (Jl± = Jlx ±iJly, l = 1, 2) is the collective pseudospin operator with spin quantum-number sl = Nl/2. Nl denotes the atom number of l-th component and ωl is the atomic frequency. a (a) is the photon creation (annihilation) operator and gl is the atom-field coupling strength.

 figure: Fig. 1

Fig. 1 Schematic diagram for two ensembles of ultracold atoms (blue and green) with transition frequencies ω1, ω2 in an optical cavity of frequency ω.

Download Full Size | PDF

3. Spin coherent-state variational method

In this paper we provide analytic solutions for the macroscopic quantum state (MQS) for the spin-boson system in terms of the recently developed spin coherent variational method [48, 73, 78, 79]. The meaning of MQS in the present paper is that the variational wave function is considered as a product of boson and spin coherent states seen in the followings. We begin with the partial average of the system Hamiltonian in the trial wave function |α⟩, which is assumed as the boson coherent state of cavity mode such that a|α⟩= α|α⟩. After the average in the boson coherent state we obtain an effective Hamiltonian of the pseudospin operators only,

Hsp(α)=α|H|α=ωα*α+l=1,2ωlJlz+l=1,2glNl(α*+α)(Jl++Jl),
which is going to be diagonalized in terms of spin coherent-state transformation. A spin coherent state can be generated from the maximum Dicke states |s, ±s⟩ (Jz|s, ±s⟩ = ±s|s, ±s⟩) with a spin coherent-state transformation [74,80]. For the l-th component pseudospin operator we have two orthogonal coherent states defined by
|±nl=R(nl)|s,±sl,
which are called north- and south- pole gauges respectively. As a matter of fact the spin coherent states are actually the eigenstates of the spin projection operator Jl · nlnl⟩ = ±jnl⟩, where nl = (sin θl cos φl, sin θl sin φl, cos θl) is the unit vector with the directional angles θl and φl. In the spin coherent states the spin operators satisfy the minimum uncertainty relation, for example, ΔJ+ΔJ = ⟨Jz⟩/2 so that |±n⟩ are called the MQSs. The unitary operator is explicitly given by
R(nl)=eθl2(Jl+JeiφlJleiφl).
Since pseudospin operators for two components of atoms commute each other, the entire trail-wave-function is the direct product of two-component spin coherent states
|ψs=|±n1|±n2,
which is required to be the energy eigenstate of the effective Hamiltonian of pseudospin operator such that
Hsp(α)|ψs=E(α)|ψs.
Where
|ψs=U|±s1|±s2,
with
U=R(n1)R(n2),
being the total unitary operator of spin coherent-state transformation. It is a key point to take into account of both spin coherent states |±n⟩ for revealing the multi-stable MQSs. Applying the unitary transformation U = R(n2)R(n1) to the energy eigenequation Eq. (3) we have
H˜sp(α)|±s1|±s2=E(α)|±s1|±s2,
where
H˜sp(α)=UHsp(α)U.
Under the spin coherent-state transformation the spin operators Jlz, Jl+, Jl (l = 1, 2) become [80]
J˜lz=Jlzcosθl+12sinθl(Jl+eiφl+Jleiφl),J˜l+=Jl+cos2θl2Jle2iφlsin2θl2Jlzeiφlsinθl,J˜l=Jlcos2θl2Jl+e2iφlsin2θl2Jlzeiφlsinφl.
Then the effective spin Hamiltonian can be diagonalized under the conditions
ωl2eiφlsinθl+g˜lα(cos2θl2e2iφlsin2θl2)=0,ωl2eiφlsinθl+g˜lα(cos2θl2e2iφlsin2θl2)=0,
from which the angle parameters θl, φl are determined in principle. Thus we obtain the energy function
E(α)=ω|α|2±l=1,2Nl2Al(α,θl,φl),
where
Al(α,θl,φl)=ωlcosθlg˜lα(eiφl+eiφl)sinθl,
with g˜lα=glNl(α*+α). The total trial-wave-function is
|ψ=|α|ψs,
and corresponding energies are found as local minima of the energy function E (α), in which the complex eigenvalue of boson coherent state is parametrized as
α=γeiϕ.
By solving the Eq. (6) and eliminating the angle parameters θl, φl, φ we derive the scaled-energy as a function of one variational-parameter γ only
Eω(γ)=γ2±l=1,2,Nl2(ωlω)2+16γ2Nl(glω)2.
The local minima of energy function Eq. (8) can be determined in terms of the variation with respect to the parameter γ.

4. Multi-stable states and phase diagram

In our formalism both the normal (⇓) and inverted (⇑) pseudospin states [68, 75] are taken into account to reveal the multiple stable states. Thus there exist four combinations of two-spin states labeled by ⇊ (both normal spins), ⇈ (both inverted spins), ⇵ and ⇅ (first-spin normal, second-spin inverted and vas versa). For the configuration of both normal spins the dimensionless energy is

E(γ)ω=γ2l=1,2Nl2(ωlω)2+16γ2Nl(glω)2.
In the following evaluations we assume the equal atom numbers for the two components that N1 = N2 = N/2. The atomic frequencies are parametrized according to the cavity frequency ω and atom-field detuning Δ
ω1=ωΔ,ω2=ω+Δ.
The ground-state is obtained from the variation of average energy
ε=E(γ)Nω
with respect to the variational parameter γ. The energy extremum condition is found as
εγ=2γp(γ)=0,
where
p(γ)=1l=1,24gl2ω2Fl(γ),
and
Fl(γ)=(ωlω)2+32(glω)2γ2N.
The extremum condition Eq. (10) possesses always a zero photon-number solution γ = 0, which is stable if the second-order derivative of energy function,
2(ε (γ 2=0)γ2=2[14ω(g12ω1+g22ω2)],
is positive. Therefore a phase boundary is determined from 2(ε(γ2=0)/γ2=0, which gives rise to the relation of two critical coupling values
g1,c2ω1+g2,c2ω2=ω4.
When
g12ω1+g22ω2<ω4,
we have a stable zero photon-number solution, which we call the NP denoted by N. The energy function for the configuration ⇵ is
ε=γ2N14[F1(γ)F2(γ)]
The energy extremum condition ∂ε/∂γ = 2γ p (γ) = 0 with
p(γ)=14ω2[g12F1(γ)g22F2(γ)],
has the zero photon-number solution, which is stable when the second-order derivative
2(ε(γ2=0))γ2=2N[14ω(g12ω1g22ω2)]
is positive. Thus we have the NP (denoted by N) region when
g12ω1g22ω2<ω4.
Correspondingly for the configuration ⇅ the energy function is
ε=γ2N+14[F1(γ)F2(γ)].
The energy extremum condition is ∂ε/∂γ = γp(γ) = 0 with
p(γ)=1+4ω2(g12F1(γ)g22F2(γ)).
Again the stable zero photon-number solution denoted by N requires
g22ω2g12ω1<ω4.
The energy function for the configuration ⇈ is
ε=E(γ)ωN=γ2N+14l=1,2Fl(γ).
The extremum condition is
(ε )γ=2γ p (γ )=0,
with
p (γ )=1+l=1,24gl2ω2Fl(γ ).
The zero photon-number solution is stable denoted by N since the second-order derivative
2ε(γ2=0)γ2=2N[1+4ω(g12ω1+g22ω2)]>0,
is always positive. The nonzero-photon solution can be obtained from the extremum condition.
pk(γsk)=0
for the four configurations k =⇊, ⇵, ⇅, ⇈. The extremum condition Eq. (14) is able to be solved numerically. We display in Fig. 2(a) the stable nonzero photon solutions γsk, which are called the superradiant states, and the corresponding energies ε(γsk) as shown in Fig. 2(b) for k =⇊ (black line), ⇵ (olive line), ⇅ (blue line) respectively. For the dimensionless coupling g2/ω = 0.2 [Figs. 2(a1) and 2(b1)] both solutions γs and γs of the extremum equation are stable with a positive sloop [Fig. 2(a1)], namely a positive second-order derivative of the energy function with respect the variation parameter γ. The corresponding energies are local minima [Fig. 2(b1)]. γs indicates the solution of stimulated radiation from the state of atomic population inversion for the second-component of atoms. Increasing the coupling strength to g2/ω = 0.4 [Figs. 2(a2) and 2(b2)] and 0.7 we have only one stable solution γs. While the two stable solutions appear again for g2/ω = 0.9 [Figs. 2(a4) and 2(b4)]. It is interesting to see a fact that the the stimulated radiation becomes the first-component of atoms i.e. γs. The superradiant states are denoted respectively by S, S and S in the following phase diagrams. The new observation with the spin coherent-state variational-method is that besides the ground states we also obtained the stable MQSs of higher energies. Fig. 3 depicts the phase diagram in g1g2 plane with the resonance condition ω1 = ω2 = ω. The phase boundaries gc gc⇵ gc are determined from the following three relations respectively
g2=121(2g1ω)2,g2=12(2g1ω)2+1,g2=12(2g1ω)21.
In the region denoted by NPts (bounded by the critical line gc) there exist triple zero-photon states, in which N with lowest energy is the ground state. This region is separated into two areas (pink and yellow) with only one state difference that the state N in one area is replaced by N in the other. We see the simultaneous spin-flip from the state N to N by adjusting the ratio of two coupling constants from g2/g1 < 1 (yellow region) to g2/g1 > 1 (pink region). The notation, for example, SPco (S, N, N) (cyan area) means the SP region characterized by the superradiant ground-state S coexisting with the first (N) and second (N) excited states of zero photons). The phase diagram is symmetric with respect to the line g2/g1 = 1, which separates the SP region to two areas. Below the symmetric line (green area) only the first excited state is changed to N by the coupling-variation induced spin flip. The critical line gc is a boundary, above which the first excited state becomes surperradiant state S (cyan region) in the upper area of the symmetric line. While gc is the corresponding boundary for the first excited states N and S (olive area). The superradiant states S, S, which are new observation for the two-component BECs, are seen to be the stimulated radiation from the higher-energy atomic levels. The stable population inversion state N for both components exists in the whole region. The multi-stable MQSs observed in this paper agree with the dynamic study of nonequilibrium QPTs [68,75].

 figure: Fig. 2

Fig. 2 Graphical solutions of the extremum equation pk (γsk) = 0 for k = ⇊ (black line), k (olive line), and k =⇅ (blue line) with g1/ω = 0.6 and g2/ω = 0.2 (a1), 0.4 (a2), 0.7 (a3), 0.9 (a4). The corresponding average energy curves ε are plotted in the lower panel (b1–b4). γ¯2=γ2/N denotes the mean photon number.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Phase diagram in the resonance condition ω1 = ω2 = ω. The notations NPts(N, N,N) and NPts (N, N,N) mean the NP with triple states, in which N is the ground state. SPco (S, N, N) [SPco (S, N,N)] means the SP characterized by the ground state S, which coexists with N (N) and N. SPco (S, S,N) [SPco (S, S,N)] is also the coexisting SP, in which the first excited-state is a superradiant state S (S).

Download Full Size | PDF

We now consider the phase diagram for the atom-field detuning ω1 = ω − Δ and ω2 = ω + Δ with Δ ∈ [−0.9, 0.9] and the atom-field coupling imbalance parameter δ given by

g1=g,g2=(1+δ)g.
Substituting atom-field coupling Eq. (15) into the corresponding ground-state energy function we obtain the phase diagram of g-Δ space displayed in Fig. 4 for the imbalance parameter δ = 0 [Fig. 4(a)], 0.5 [Fig. 4(b)], −0.5 [Fig. 4(c)]. The phase boundary line gc for the normal state N is found from Eq. (11)
gc=12(ω2Δ2)ω[2ω+(ωΔ)(2δ+δ2)],
The phase diagram for δ = 0 as depicted in Fig. 4(a) is symmetric with respect to the horizontal line Δ = 0. The triple-state NP region denoted by NPts (N, N, N) (yellow) and NPts (N, N, N) (pink) is located on the left-hand side of the critical line gc, which shifts towards the lower value direction of the atom-field coupling g [21, 68] with the increase of absolute value of detuning |Δ| seen from Fig. 4(a). SPco (S, N,N) (green region) and SPco (S, N,N) (cyan) denote the SP characterized by the ground-state S coexisting with the normal states N, N and N respectively. The QPT from the NP of ground-state N to the SP ground-state S by the variation of atom-field coupling g is the standard DM type for the fixed atom-field detuning Δ. The phase boundary lines gc, gc, which separate the states S and S, are respectively determined from Eqs. (12, 13)
gc=12(ω2Δ2)ω[(ω+Δ)(ωΔ)(1+δ)2]=12(ω2Δ2)ω[2Δ(ωΔ)(2δ+δ2)],
and
gc=12(ω2Δ2)ω[(ωΔ)(1+δ)2(ω+Δ)]=12(ω2Δ2)ω[(ωΔ)(2δ+δ2)2Δ].
The superradiant region denoted by SPco S, S N (olive area) is above the the critical line gc, while SPco (S, S,N) (blue) is located below the critical line gc. We see that the second excited-state varies from the normal state N to the superradiant state S by the increase of detuning Δ. The difference of upper and lower half-plane of the phase diagram is made only by the first excited-states N, S and N, S with the interchange of spin polarizations between two components. This boundary line, which separates the regions with different first excited-states, moves upward and downward respectively for δ = 0.5 [Fig. 4(b)], −0.5 [Fig. 4(c)].

 figure: Fig. 4

Fig. 4 Phase diagram in g-Δ space with the atom-photon coupling parameter δ = 0 (a), δ = 0.5 (b), and δ = −0.5 (c). The boundary line, which separates the regions with different first-excited-states (N, S and N, S), moves upward and downward respectively for δ = 0.5 (b), −0.5 (c).

Download Full Size | PDF

5. Mean photon number, atomic population and average energy from viewpoint of phase transition

The mean photon numbers in the states N and S can be evaluated directly from the average of photon number-operator in the corresponding wave functions |ψ〉 = |α〉 |ψs〉 in Eq. (7) with spin-state |ψs (−s, −s)〉 = U|−s1 |−s2 given in Eq. (4). The result is obviously

np()=α|aa|αN={0,g<gc,γ2N,g>gc..
While the atomic population imbalance becomes
Δna()=ψs(s,s)|(J1z+J2z)|ψs(s,s)N=14l=1,2ωlωFl(γ),
which reduces to the well-known standard Dicke-model value
Δna()=12,
at the critical line gc and also the NP state N. The average energy in ground states N and S is given by
ε={0.5,g<gc,γ2N14l=1,2Fl(γ),g>gc..
For the states Nk and Sk with opposite spin-polarizations k =⇵,⇅ the average photon number is
np(Nk)=0;np(Sk)=γk2N.
The atomic population imbalance becomes
Δna(Nk)=0
for the zero-photon states Nk. While the atomic population imbalance for the superradiant states Sk is seen to be
Δna(S)=14ω[ω1F1(γ)+ω2F2(γ)],Δna(S)=14ω[ω1F1(γ)ω2F2(γ)].
The average energies εk (Sk) of the superradiant states Sk for k = ⇵,⇅, can be obtained from the energy functions with the corresponding solutions γk, which lead to εk (Nk) = 0. For the inverted-spin state of zero photon the atomic population imbalance is Δna(N) = 0.5 and the average energy is found as
ε(N)=14ω(ω1+ω2)
The stable nonzero-photon state does not exists for this configuration of both inverted spins. The average photon number np, atomic population imbalance Δna, and the average energy ε are plotted in Fig. 5 as functions of the atom-field coupling strength g in the red and blue detuning Δ = ±0.6 with δ = 0. Below the critical point gc we have triple stable (zero-photon) states denoted by NPts (N, N, N) [or NPts (N, N, N)], in which N (black line) is the ground state with lowest energy. Between the critical points gc and gc (or gc) the superradiant ground-state S (black line) coexists with the states N [olive lines in Figs. 5(a1)–5(c1)], or N [blue lines in Figs. 5(a2)–5(c2)], and N (red lines). The QPT from the NP (N) to the SP (S) is the standard DM type, which takes place at the critical point gc. From Fig. 5(c1) we see that the states N and S (olive lines) of opposite spin-polarizations are the first excited-states in the case Δ = 6. While the states N and S with interchange of the spin polarizations between two components become the first excited states seen from Fig. 5(c2) (blue lines) for the negative detuning Δ = −6. We observe for the first time the phase transition at the critical point gc (gc) from the normal state N (N) to the suprradiant state S (S), which is the stimulated radiation from the collective states of atomic population inversion for one component of BECs seen from Figs. 5 and 6. The ground state does not change at the critical point gc (or gc), which separates the normal state N (or N) and the superradiant one S (or S), which are the collective excited-states of the system. For the given frequency detuning Δ = ±0.6 (Fig. 5) the critical points can be evaluated precisely for Eq. (16, 17, 18), gc =2/5=0.2828 and gc=gc=2/15=0.365148. The normal state N (red line) of atomic population inversion for both components does not involve in radiation process.

 figure: Fig. 5

Fig. 5 Variations of the average photon number np (a), atom population imbalance Δna (b), and average energy ε (c) with respect to the coupling constant g = g1 = g2 in the atom-field frequency detuning Δ = 0.6 (1) and Δ = −0.6 (2).

Download Full Size | PDF

We display the variation curves of average photon-number np as shown in Figs. 6(a1) and 6(a2), atom population imbalance Δna as shown in Figs. 6(b1) and 6(b2), and energy ε as depicted in Figs. 6(c1) and 6(c2) with respect the coupling constant g for the imbalance parameter δ = ±0.5 at the resonance condition Δ = 0. The QPT from normal state N to the superradiant state S takes place at the critical point gc=5/5=0.447214. In the case δ = −5 as depicted in Figs. 6(a1)–6(c1), namely the second component has lower coupling value, an additional transition appears between the collective excited-states N and S at the critical point gc=3/3=0.577350. This transition is from the normal state of atomic population inversion to the superradiant state for the second component realized from atom population imbalance and the energy in Figs. 5(b1) and 5(c1). By adjusting the imbalance parameter to δ = 0.5 as shown Figs. 6(a2)–6(c2), the transition becomes from N to S for the first component. The collective stimulated-radiation shifts to the first component, which has lower atom-field coupling than the second component in this case. The transition critical point is found as gc=5/5=0.447214.

 figure: Fig. 6

Fig. 6 The average photon number np (a), atomic population Δna (b), and average energy ε (c) curves for the imbalance parameter δ = −0.5 (1), δ = 0.5 (2) in the resonance condition Δ = 0.0. The stimulated radiation shifts from one component to the other by adjusting the relative coupling constants.

Download Full Size | PDF

6. Conclusion and discussion

In summary, multiple MQSs are derived analytically for two-component BECs in a single-mode cavity by means of the spin coherent-state variational method. The rich phase diagrams are presented with the variation of atom-field coupling imbalance between two components and the atom-field frequency detuning. Indeed the ground states display a typical Dicke-model QPT from the NP to SP for both components in the normal spin-states. When the atom-field coupling imbalance between two components increases the normal spin-state with relatively lower coupling-value flips to the inverted spin-state, the radiation from this state is the stimulated radiation from atomic population-inversion levels. The stimulated radiation can be also generated from manipulation of atom-field frequency detuning. In the specific cases when one of the coupling constants vanishes or two couplings are equal the ground-states and related QPT reduce to that of an ordinary Dicke model. The controllable stimulated radiation may have technical applications in the laser physics. The spin coherent-state variational method is a powerful tool in the study of macroscopic quantum properties for the atom-ensemble and cavity-field system, since it takes into account both the normal and inverted pseudospins, which result in multiple MQSs in agreement with the semiclassical dynamics of nonequilibrium QPT in the Dicke model [76]. In addition a one-parameter variational energy-function is able to be derived in this formalism, so that one can evaluate the second-order derivative to determine rigorously the local minima of energy functions in consistence with the numerical simulation [48, 68, 73, 75, 77].

Funding

National Natural Science Foundation of China (Grant Nos. 11275118, 11404198, 91430109, 61505100); the Scientific and Technological Innovation Programs of Higher Education Institutions in Shanxi Province (STIP) (Grant No. 2014102); the Launch of the Scientific Research of Shanxi University (Grant No. 011151801004); the National Fundamental Fund of Personnel Training (Grant No. J1103210); Natural Science Foundation of Shanxi Province(Grant No. 2015011008).

References and links

1. R. H. Dicke, “Coherence in Spontaneous Radiation Processes,” Phys. Rev. 93, 99 (1954). [CrossRef]  

2. K. Baumann, C. Guerlin, F. Brennecke, and T. Esslinger, “Dicke quantum phase transition with a superfluid gas in an optical cavity,” Nature (London) 464, 1301 (2010). [CrossRef]  

3. K. Baumann, R. Mottl, F. Brennecke, and T. Esslinger, “Exploring symmetry breaking at the Dicke quantum phase transition,” Phys. Rev. Lett. 107, 140402 (2011). [CrossRef]   [PubMed]  

4. H. Ritsch, P. Domokos, F. Brennecke, and T. Esslinger, “Cold atoms in cavity-generated dynamical optical potentials,” Rev. Mod. Phys. 85, 553 (2013). [CrossRef]  

5. Y. K. Wang and F. T. Hioe, “Phase transition in the Dicke model of superradiance,” Phys. Rev. A , 7, 831 (1973). [CrossRef]  

6. F. T. Hioe, “Phase transitions in some generalized Dicke models of superradiance,” Phys. Rev. A 8, 1440 (1973). [CrossRef]  

7. C. Emary and T. Brandes, “Chaos and the quantum phase transition in the Dicke model,” Phys. Rev. E 67, 066203 (2003). [CrossRef]  

8. Y. Colombe, T. Steinmetz, G. Dubois, F. Linke, D. Hunger, and J. Reichel, “Strong atom-field coupling for Bose-Einstein condensates in an optical cavity on a chip,” Nature (London) 450, 272 (2007). [CrossRef]  

9. F. Brennecke, T. Donner, S. Ritter, T. Bourdel, M. Köhl, and T. Esslinger, “Cavity QED with a Bose-Einstein condensate,” Nature (London) 450, 268 (2007). [CrossRef]  

10. R. Puebla, A. Relaño, and J. Retamosa, “Excited-state phase transition leading to symmetry-breaking steady states in the Dicke model,” Phys. Rev. A 87, 023819 (2013). [CrossRef]  

11. J. Larson, B. Damski, G. Morigi, and M. Lewenstein, “Mott-insulator states of ultracold atoms in optical resonators,” Phys. Rev. Lett. 100, 050401 (2008). [CrossRef]   [PubMed]  

12. S. Morrison and A. S. Parkins, “Dynamical quantum phase transitions in the dissipative Lipkin-Meshkov-Glick model with proposed realization in optical cavity QED,” Phys. Rev. Lett. 100, 040403 (2008). [CrossRef]   [PubMed]  

13. J. M. Zhang, W. M. Liu, and D. L. Zhou, “Mean-field dynamics of a Bose Josephson junction in an optical cavity,” Phys. Rev. A 78, 043618 (2008). [CrossRef]  

14. G. Chen, X. G. Wang, J. -Q. Liang, and Z. D. Wang, “Exotic quantum phase transitions in a Bose-Einstein condensate coupled to an optical cavity,” Phys. Rev. A 78, 023634 (2008). [CrossRef]  

15. J. Larson and J. -P. Martikainen, “Ultracold atoms in a cavity-mediated double-well system,” Phys. Rev. A 82, 033606 (2010). [CrossRef]  

16. L. Zhou, H. Pu, H. Y. Ling, K. Zhang, and W. P. Zhang, “Spin dynamics and domain formation of a spinor Bose-Einstein condensate in an optical cavity,” Phys. Rev. A 81, 063641 (2010). [CrossRef]  

17. M. J. Bhaseen, M. Hohenadler, A. O. Silver, and B. D. Simons, “Polaritons and Pairing Phenomena in Bose-Hubbard Mixtures,” Phys. Rev. Lett. 102, 135301 (2009). [CrossRef]   [PubMed]  

18. A. O. Silver, M. Hohenadler, M. J. Bhaseen, and B. D. Simons, “Bose-Hubbard models coupled to cavity light fields,” Phys. Rev. A 81, 023617 (2010). [CrossRef]  

19. G. Szirmai, D. Nagy, and P. Domokos, “Excess noise depletion of a Bose-Einstein Condensate in an optical cavity,” Phys. Rev. Lett. 102, 080401 (2009). [CrossRef]   [PubMed]  

20. G. Szirmai, D. Nagy, and P. Domokos, “Quantum noise of a Bose-Einstein condensate in an optical cavity, correlations, and entanglement,” Phys. Rev. A 81, 043639 (2010). [CrossRef]  

21. N. Liu, J. L. Lian, J. Ma, L. T. Xiao, G. Chen, J-Q. Liang, and S. T. Jia, “Light-shift-induced quantum phase transitions of a Bose-Einstein condensate in an optical cavity,” Phys. Rev. A 83, 033601 (2011). [CrossRef]  

22. B. V. Thompson, “A canonical transformation theory of the generalized Dicke model,” J. Phys. A 10, 89 (1977). [CrossRef]  

23. D. Tolkunov and D. Solenov, “Quantum phase transition in the multimode Dicke model,” Phys. Rev. B 75, 024402 (2007). [CrossRef]  

24. M. Mariantoni, F. Deppe, A. Marx, R. Gross, F. K. Wilhelm, and E. Solano, “Two-resonator circuit quantum electrodynamics: A superconducting quantum switch,” Phys. Rev. B 78, 104508 (2008). [CrossRef]  

25. D. G. Norris, E. J. Cahoon, and L. A. Orozco, “Atom detection in a two-mode optical cavity with intermediate coupling: Autocorrelation studies,” Phys. Rev. A 80, 043830 (2009). [CrossRef]  

26. M. L. Terraciano, R. Olson Knell, D. G. Norris, J. Jing, A. Fernández, and L. A. Orozco, “Photon burst detection of single atoms in an optical cavity,” Nat. Phys. 5, 480 (2009). [CrossRef]  

27. P. Nataf and C. Ciuti, “Protected quantum computation with multiple resonators in ultrastrong coupling circuit QED,” Phys. Rev. Lett. 107, 190402 (2011). [CrossRef]   [PubMed]  

28. J. Q. You and F. Nori, “Atomic physics and quantum optics using superconducting circuits,” Nature (London) 474, 589 (2011). [CrossRef]  

29. H. Wang, M. Mariantoni, R. C. Bialczak, M. Lenander, E. Lucero, M. Neeley, A. D. O’Connell, D. Sank, M. Weides, J. Wenner, T. Yamamoto, Y. Yin, J. Zhao, J. M. Martinis, and A. N. Cleland, “Deterministic entanglement of photons in two superconducting microwave resonators,” Phys. Rev. Lett. 106, 060401 (2011). [CrossRef]   [PubMed]  

30. Y. Eto, A. Noguchi, P. Zhang, M. Ueda, and M. Kozuma, “Projective measurement of a single nuclear spin qubit by using two-mode cavity QED,” Phys. Rev. Lett. 106, 160501 (2011). [CrossRef]   [PubMed]  

31. M. Mariantoni, H. Wang, R. C. Bialczak, M. Lenander, E. Lucero, M. Neeley, A. D. O’Connell, D. Sank, M. Weides, J. Wenner, T. Yamamoto, Y. Yin, J. Zhao, J. M. Martinis, and A. N. Cleland, “Photon shell game in three-resonator circuit quantum electrodynamics,” Nat. Phys. 7, 287 (2011). [CrossRef]  

32. D. J. Egger and F. K. Wilhelm, “Multimode circuit quantum electrodynamics with hybrid metamaterial transmission lines,” Phys. Rev. Lett. 111, 163601 (2013). [CrossRef]   [PubMed]  

33. C.-P. Yang, Q.-P. Su, S.-B. Zheng, and S. Y. Han, “Generating entanglement between microwave photons and qubits in multiple cavities coupled by a superconducting qutrit,” Phys. Rev. A 87, 022320 (2013). [CrossRef]  

34. J. Larson and S. Levin, “Effective abelian and non-abelian gauge potentials in cavity QED,” Phys. Rev. Lett. 103, 013602 (2009). [CrossRef]   [PubMed]  

35. J. Larson, “Analog of the spin-orbit-induced anomalous Hall effect with quantized radiation,” Phys. Rev. A 81, 051803 (2010). [CrossRef]  

36. S. Gopalakrishnan, B. L. Lev, and P. M. Goldbart, “Emergent crystallinity and frustration with Bose-Einstein condensates in multimode cavities,” Nat. Phys. 5, 845 (2009). [CrossRef]  

37. S. Gopalakrishnan, B. L. Lev, and P. M. Goldbart, “Frustration and glassiness in spin models with cavity-mediated interactions,” Phys. Rev. Lett. 107, 277201 (2011). [CrossRef]  

38. P. Strack and S. Sachdev, “Dicke quantum spin glass of atoms and photons,” Phys. Rev. Lett. 107, 277202 (2011). [CrossRef]  

39. M. Buchhold, P. Strack, S. Sachdev, and S. Diehl, “Dicke-model quantum spin and photon glass in optical cavities: Nonequilibrium theory and experimental signatures,” Phys. Rev. A 87, 063622 (2013). [CrossRef]  

40. J. T. Fan, Z. W. Yang, Y. W. Zhang, J. Ma, G. Chen, and S. T. Jia, “Hidden continuous symmetry and Nambu-Goldstone mode in a two-mode Dicke model,” Phys. Rev. A , 89, 023812 (2014). [CrossRef]  

41. A. Wickenbroc, M. Hemmerling, G. R. M. Robb, C. Emary, and F. Renzoni, “Collective strong coupling in multi-mode cavity QED,” Phys. Rev. A 87, 043817 (2013). [CrossRef]  

42. D. O. Krimer, M. Liertzer, S. Rotter, and H. E. Türeci, “Route from spontaneous decay to complex multimode dynamics in cavity QED,” arXiv: 1306.4787.

43. T. J. Kippenberg and K. J. Vahala, “Cavity optomechanics: Back-action at the mesoscale,” Science 321, 1172 (2008) [CrossRef]   [PubMed]  

44. F. Marquardt and S. M. Girvin, “Optomechanics,” Physics 2, 40 (2009). [CrossRef]  

45. I. Favero and K. Karrai, “optomechanics of deformable optical cavities,” Nature Photonics 3, 201 (2009). [CrossRef]  

46. M. Aspelmeyer, S. Gröblacher, K. Hammerer, and N. Kiesel, “Quantum optomechanics—throwing a glance,” J. Opt. Soc. Am. B 27, A189 (2010). [CrossRef]  

47. C. A. Regal and K. W. Lehnert, “From cavity electromechanics to cavity optomechanics,” J. Phys.: Conf. Ser. 264, 012025 (2011).

48. Z. M. Wang, J. L. Lian, J.-Q. Liang, Y. M. Yu, and W. M. Liu, “Collapse of the superradiant phase and multiple quantum phase transitions for Bose-Einstein condensates in an optomechanical cavity,” Phys. Rev. A 93, 033630 (2016). [CrossRef]  

49. J.-Q. Liang, J.-L. Liu, W.-D. Li, and Z.-J. Li, “Atom-pair tunneling and quantum phase transition in the strong-interaction regime,” Phys. Rev. A 79, 033617 (2009). [CrossRef]  

50. H. Cao and L. B. Fu, “Quantum phase transition and dynamics induced by atom-pair tunnelling of Bose-Einstein condensates in a double-well potential,” Eur. Phys. J. D 66, 97 (2012). [CrossRef]  

51. Y. C. Zhang, X. F. Zhou, G. C. Guo, X. X. Zhou, H. Pu, and Z. W. Zhou, “Two-component polariton condensate in an optical microcavity,” Phys. Rev. A 89, 053624 (2014). [CrossRef]  

52. E. Timmermans, “Phase separation of Bose-Einstein condensates,” Phys. Rev. Lett. 81, 5718 (1998). [CrossRef]  

53. H. Pu and N.P. Bigelow, “Properties of two-species Bose condensates,” Phys. Rev. Lett. 80, 1130 (1998). [CrossRef]  

54. Y. Dong, J. W. Ye, and H. Pu, “Multistability in an optomechanical system with a two-component Bose-Einstein condensate,” Phys. Rev. A 83, 031608 (2011). [CrossRef]  

55. K Sasaki, N. Suzuki, and H. Saito, “Capillary instability in a two-component Bose-Einstein condensate,” Phys. Rev. A 83, 053606 (2011). [CrossRef]  

56. R. A. Barankov, “Boundary of two mixed Bose-Einstein condensates,” Phys. Rev. A 66, 013612 (2002). [CrossRef]  

57. B. V. Schaeybroeck, “Interface tension of Bose-Einstein condensates,” Phys. Rev. A 78, 023624 (2008). [CrossRef]  

58. K. Sasaki, N. Suzuki, D. Akamatsu, and H. Saito, “Rayleigh-Taylor instability and mushroom-pattern formation in a two-component Bose-Einstein condensate,” Phys. Rev. A 80, 063611 (2009). [CrossRef]  

59. A. Søensen, L.-M. Duan, J. I. Cirac, and P. Zoller, “Many-particle entanglement with Bose-Einstein condensates,” Nature(London) 409, 63 (2001). [CrossRef]  

60. D. Gordon and C. M. Savage, “Creating macroscopic quantum superpositions with Bose-Einstein condensates,” Phys. Rev. A 59, 4623 (1999). [CrossRef]  

61. A. Micheli, D. Jaksch, J. I. Cirac, and P. Zoller, “Many-particle entanglement in two-component Bose-Einstein condensates,” Phys. Rev. A 67, 013607 (2003). [CrossRef]  

62. M. R. Andrews, C. G. Townsend, H.-J. Miesner, D. S. Durfee, D. M. Kurn, and W. Ketterle, “Observation of interference between two Bose condensates,” Science 275, 637 (1997). [CrossRef]   [PubMed]  

63. M. Eto, K. Kasamatsu, M. Nitta, H. Takeuchi, and M. Tsubota, “Interaction of half-quantized vortices in two-component Bose-Einstein condensates,” Phys. Rev. A 83, 063603 (2011). [CrossRef]  

64. C. Emary and T. Brandes, “Quantum chaos triggered by precursors of a quantum phase transition: The Dicke model,” Phys. Rev. Lett. 90, 044101 (2003). [CrossRef]   [PubMed]  

65. G. Chen, J. Q. Li, and J.-Q. Liang, “Critical property of the geometric phase in the Dicke model,” Phys. Rev. A 74, 054101 (2006). [CrossRef]  

66. J. L. Lian, Y. W. Zhang, and J.-Q Liang, “Macroscopic quantum states and quantum phase transition in the Dicke model,” Chin. Phys. Lett. 29, 060302 (2012). [CrossRef]  

67. J. L. Lian, N. Liu, J.-Q Liang, G. Chen, and S. T. Jia, “Ground-state properties of a Bose-Einstein condensate in an optomechanical cavity,” Phys. Rev. A 88, 043820 (2013). [CrossRef]  

68. N. Liu, J. D. Li, and J. -Q. Liang, “Nonequilibrium quantum phase transition of Bose-Einstein condensates in an optical cavity,” Phys. Rev. A 87, 053623 (2013). [CrossRef]  

69. R. Gilmore and L.M. Narducci, “Relation between the equilibrium and nonequilibrium critical properties of the Dicke model,” Phys. Rev. A 17, 1747 (1978). [CrossRef]  

70. F. Dimer, B. Estienne, A. S. Parkins, and H. J. Carmichael, “Proposed realization of the Dicke-model quantum phase transition in an optical cavity QED system,” Phys. Rev. A 75, 013804 (2007). [CrossRef]  

71. P. Horak and H. Ritsch, “Dissipative dynamics of Bose condensates in optical cavities,” Phys. Rev. A 63, 023603 (2001). [CrossRef]  

72. O. Castaños, E. Nahmad-Achar, R. López-Peñna, and J. G. Hirsch, “No singularities in observables at the phase transition in the Dicke model,” Phys. Rev. A 83, 051601 (2011). [CrossRef]  

73. X. Q. Zhao, N. Liu, and J.-Q. Liang, “Nonlinear atom-photon-interaction-induced population inversion and inverted quantum phase transition of Bose-Einstein condensate in an optical cavity,” Phys. Rev. A 90, 023622 (2014). [CrossRef]  

74. Z.-D. Chen, J.-Q. Liang, S.-Q. Shen, and W. -F. Xie, “Dynamics and Berry phase of two-species Bose-Einstein condensates,” Phys. Rev. A 69, 023611 (2004). [CrossRef]  

75. J. Keeling, M. J. Bhaseen, and B. D. Simons, “Collective dynamics of Bose-Einstein condensates in optical cavities,” Phys. Rev. Lett. 105, 043001 (2010). [CrossRef]   [PubMed]  

76. M. J. Bhaseen, J. Mayoh, B. D. Simons, and J. Keeling, “Dynamics of nonequilibrium Dicke models,” Phys. Rev. A , 85, 013817 (2012). [CrossRef]  

77. A. B. Bhattacherjee, “Non-equilibrium dynamical phases of two-Atom Dicke model,” Phys. Lett. A 378, 3244 (2014). [CrossRef]  

78. E. Layton, Y. H. Huang, and S. I. Chu, “Cyclic quantum evolution and Aharonov-Anantlan geometric phases in SU(2) spin-coherent states,” Phys. Rev. A 41, 42 (1990). [CrossRef]   [PubMed]  

79. R. F. Fox, “Generalized coherent states,” Phys. Rev. A 59, 3241 (1999). [CrossRef]  

80. Y.-Z. Lai, J.-Q. Liang, H. J. W. Müller-Kirsten, and J. G. Zhou, “Time-dependent quantum systems and the invariant Hermitian operator,” Phys. Rev. A 53, 3691 (1996). [CrossRef]   [PubMed]  

Cited By

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

Alert me when this article is cited.


Figures (6)

Fig. 1
Fig. 1 Schematic diagram for two ensembles of ultracold atoms (blue and green) with transition frequencies ω1, ω2 in an optical cavity of frequency ω.
Fig. 2
Fig. 2 Graphical solutions of the extremum equation pk (γsk) = 0 for k = ⇊ (black line), k (olive line), and k =⇅ (blue line) with g1/ω = 0.6 and g2/ω = 0.2 (a1), 0.4 (a2), 0.7 (a3), 0.9 (a4). The corresponding average energy curves ε are plotted in the lower panel (b1–b4). γ ¯ 2 = γ 2 / N denotes the mean photon number.
Fig. 3
Fig. 3 Phase diagram in the resonance condition ω1 = ω2 = ω. The notations NPts(N, N,N) and NPts (N, N,N) mean the NP with triple states, in which N is the ground state. SPco (S, N, N) [SPco (S, N,N)] means the SP characterized by the ground state S, which coexists with N (N) and N. SPco (S, S,N) [SPco (S, S,N)] is also the coexisting SP, in which the first excited-state is a superradiant state S (S).
Fig. 4
Fig. 4 Phase diagram in g-Δ space with the atom-photon coupling parameter δ = 0 (a), δ = 0.5 (b), and δ = −0.5 (c). The boundary line, which separates the regions with different first-excited-states (N, S and N, S), moves upward and downward respectively for δ = 0.5 (b), −0.5 (c).
Fig. 5
Fig. 5 Variations of the average photon number np (a), atom population imbalance Δna (b), and average energy ε (c) with respect to the coupling constant g = g1 = g2 in the atom-field frequency detuning Δ = 0.6 (1) and Δ = −0.6 (2).
Fig. 6
Fig. 6 The average photon number np (a), atomic population Δna (b), and average energy ε (c) curves for the imbalance parameter δ = −0.5 (1), δ = 0.5 (2) in the resonance condition Δ = 0.0. The stimulated radiation shifts from one component to the other by adjusting the relative coupling constants.

Equations (51)

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

H = ω a a + l = 1 , 2 ω l J l z + l = 1 , 2 g l N l ( a + a ) ( J l + + J l ) .
H s p ( α ) = α | H | α = ω α * α + l = 1 , 2 ω l J l z + l = 1 , 2 g l N l ( α * + α ) ( J l + + J l ) ,
| ± n l = R ( n l ) | s , ± s l ,
R ( n l ) = e θ l 2 ( J l + J e i φ l J l e i φ l ) .
| ψ s = | ± n 1 | ± n 2 ,
H s p ( α ) | ψ s = E ( α ) | ψ s .
| ψ s = U | ± s 1 | ± s 2 ,
U = R ( n 1 ) R ( n 2 ) ,
H ˜ s p ( α ) | ± s 1 | ± s 2 = E ( α ) | ± s 1 | ± s 2 ,
H ˜ s p ( α ) = U H s p ( α ) U .
J ˜ l z = J l z cos θ l + 1 2 sin θ l ( J l + e i φ l + J l e i φ l ) , J ˜ l + = J l + cos 2 θ l 2 J l e 2 i φ l sin 2 θ l 2 J l z e i φ l sin θ l , J ˜ l = J l cos 2 θ l 2 J l + e 2 i φ l sin 2 θ l 2 J l z e i φ l sin φ l .
ω l 2 e i φ l sin θ l + g ˜ l α ( cos 2 θ l 2 e 2 i φ l sin 2 θ l 2 ) = 0 , ω l 2 e i φ l sin θ l + g ˜ l α ( cos 2 θ l 2 e 2 i φ l sin 2 θ l 2 ) = 0 ,
E ( α ) = ω | α | 2 ± l = 1 , 2 N l 2 A l ( α , θ l , φ l ) ,
A l ( α , θ l , φ l ) = ω l cos θ l g ˜ l α ( e i φ l + e i φ l ) sin θ l ,
| ψ = | α | ψ s ,
α = γ e i ϕ .
E ω ( γ ) = γ 2 ± l = 1 , 2 , N l 2 ( ω l ω ) 2 + 16 γ 2 N l ( g l ω ) 2 .
E ( γ ) ω = γ 2 l = 1 , 2 N l 2 ( ω l ω ) 2 + 16 γ 2 N l ( g l ω ) 2 .
ω 1 = ω Δ , ω 2 = ω + Δ .
ε = E ( γ ) N ω
ε γ = 2 γ p ( γ ) = 0 ,
p ( γ ) = 1 l = 1 , 2 4 g l 2 ω 2 F l ( γ ) ,
F l ( γ ) = ( ω l ω ) 2 + 32 ( g l ω ) 2 γ 2 N .
2 ( ε   ( γ   2 = 0 ) γ 2 = 2 [ 1 4 ω ( g 1 2 ω 1 + g 2 2 ω 2 ) ] ,
g 1 , c 2 ω 1 + g 2 , c 2 ω 2 = ω 4 .
g 1 2 ω 1 + g 2 2 ω 2 < ω 4 ,
ε = γ 2 N 1 4 [ F 1 ( γ ) F 2 ( γ ) ]
p ( γ ) = 1 4 ω 2 [ g 1 2 F 1 ( γ ) g 2 2 F 2 ( γ ) ] ,
2 ( ε ( γ 2 = 0 ) ) γ 2 = 2 N [ 1 4 ω ( g 1 2 ω 1 g 2 2 ω 2 ) ]
g 1 2 ω 1 g 2 2 ω 2 < ω 4 .
ε = γ 2 N + 1 4 [ F 1 ( γ ) F 2 ( γ ) ] .
p ( γ ) = 1 + 4 ω 2 ( g 1 2 F 1 ( γ ) g 2 2 F 2 ( γ ) ) .
g 2 2 ω 2 g 1 2 ω 1 < ω 4 .
ε = E ( γ ) ω N = γ 2 N + 1 4 l = 1 , 2 F l ( γ ) .
( ε   ) γ = 2 γ   p   ( γ   ) = 0 ,
p   ( γ   ) = 1 + l = 1 , 2 4 g l 2 ω 2 F l ( γ   ) .
2 ε ( γ 2 = 0 ) γ 2 = 2 N [ 1 + 4 ω ( g 1 2 ω 1 + g 2 2 ω 2 ) ] > 0 ,
p k ( γ s k ) = 0
g 2 = 1 2 1 ( 2 g 1 ω ) 2 , g 2 = 1 2 ( 2 g 1 ω ) 2 + 1 , g 2 = 1 2 ( 2 g 1 ω ) 2 1 .
g 1 = g , g 2 = ( 1 + δ ) g .
g c = 1 2 ( ω 2 Δ 2 ) ω [ 2 ω + ( ω Δ ) ( 2 δ + δ 2 ) ] ,
g c = 1 2 ( ω 2 Δ 2 ) ω [ ( ω + Δ ) ( ω Δ ) ( 1 + δ ) 2 ] = 1 2 ( ω 2 Δ 2 ) ω [ 2 Δ ( ω Δ ) ( 2 δ + δ 2 ) ] ,
g c = 1 2 ( ω 2 Δ 2 ) ω [ ( ω Δ ) ( 1 + δ ) 2 ( ω + Δ ) ] = 1 2 ( ω 2 Δ 2 ) ω [ ( ω Δ ) ( 2 δ + δ 2 ) 2 Δ ] .
n p ( ) = α | a a | α N = { 0 , g < g c , γ 2 N , g > g c . .
Δ n a ( ) = ψ s ( s , s ) | ( J 1 z + J 2 z ) | ψ s ( s , s ) N = 1 4 l = 1 , 2 ω l ω F l ( γ ) ,
Δ n a ( ) = 1 2 ,
ε = { 0.5 , g < g c , γ 2 N 1 4 l = 1 , 2 F l ( γ ) , g > g c . .
n p ( N k ) = 0 ; n p ( S k ) = γ k 2 N .
Δ n a ( N k ) = 0
Δ n a ( S ) = 1 4 ω [ ω 1 F 1 ( γ ) + ω 2 F 2 ( γ ) ] , Δ n a ( S ) = 1 4 ω [ ω 1 F 1 ( γ ) ω 2 F 2 ( γ ) ] .
ε ( N ) = 1 4 ω ( ω 1 + ω 2 )
Select as filters


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