This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Effect of the Atomic Dipole-Dipole Interaction on the Phase Diagrams of Field-Matter Interactions I: Variational procedure

S. Cordero sergio.cordero@nucleares.unam.mx    O. Castaños    R. López-Peña    E. Nahmad-Achar Instituto de Ciencias Nucleares, Universidad Nacional Autónoma de México, Apartado Postal 70-543, 04510 Cd.Mx., Mexico
Abstract

We establish, within the second quantization method, the general dipole-dipole Hamiltonian interaction of a system of nn-level atoms. The variational energy surface of the nn-level atoms interacting with \ell-mode fields and under the Van Der Waals forces is calculated with respect the tensorial product of matter and electromagnetic field coherent states. This is used to determine the quantum phase diagram associated to the ground state of the system and quantify the effect of the dipole-dipole Hamiltonian interaction. By considering real induced electric dipole moments, we find the quantum phase transitions for 22- and 33-level atomic systems interacting with 11- and 22- modes of the electromagnetic field, respectively. The corresponding order of the transitions is established by means of Ehrenfest classification; for some undetermined cases, we propose two procedures: the difference of the expectation value of the Casimir operators of the 22-level subsystems, and by maximizing the Bures distance between neighbor variational solutions.

I Introduction

Recently, we have studied the quantum phase diagrams of a system of nn-level atoms interacting with \ell electromagnetic modes in a cavity, under the dipolar aproximation Cordero et al. (2021); López-Peña et al. (2021).

When the inter-atomic distance of a cold atomic gas is comparable to the wavelength of the electromagnetic field, the dipole-dipole coupling between the atoms becomes important and yields relevant collective effects Weiner et al. (1999). These Van der Waals forces, due to dipole-dipole interactions of the induced electric dipole moments, become important and must be taken into account. However, one needs to be careful about the long or short character of the dipolar potential for many particle systems, as one can find in theoretical and experimental studies of ultra-cold boson systems Jones et al. (2006); Lahaye et al. (2009); Astrakharchik and Lozovik (2008).

The dipole-dipole interaction decays as 1/r31/r^{3}, with rr the distance between particles, and is thus of a different nature as the matter-field interaction considered in earlier works (cf., e.g., Cordero et al. (2015) and references therein). Energy transfer between the particles (atoms, molecules) is one of the important consequences of this interaction. For Rydberg atoms it is particularly interesting, as they have high principal quantum numbers nn, while the dipole moment scales as n2n^{2} in atomic units Gerry and Knight (2005).

A review of theoretical and experimental work on the dipole-dipole interaction between Bose-Einstein condensates has been presented in Lahaye et al. (2009). The trapping of cooled polar molecules and other atomic species Doyle et al. (2004) was important for attracting the attention to study these type of interactions, and the long-range dipole-dipole interaction in low-density atomic vapors was detected in Yu et al. (2019), confirming that the interaction is indeed long-range, and that it is present at any density.

According to the previous discussion, the interaction between atoms might be relevant to the determination of the quantum phase diagrams for the system constituted by nn-level atoms interacting with \ell-modes of electromagnetic radiation in a cavity. The main objective of this work is to quantify the effect of the atomic dipole-dipole induced interaction on the properties of the ground state of the system. The original contributions of this work are the following: To establish the general dipole-dipole interaction Hamiltonian for a system of nn-level atoms interacting with \ell-modes of electromagnetic radiation in a cavity. To calculate the associated energy surface, which allows us to determine the variational ground state, playing a fundamental role in finding the quantum phase diagrams of the system. The cases for 22- and 33- level atomic configurations are worked out explicitly, determining the quantum phase diagrams together with the corresponding order of the transitions. It is remarkable that, even for a finite number of atoms, the surface of maximum Bures distance is able to detect the phase transitions where the Ehrenfest method does not. Additionally we have found that the quantum phases continue to be dominated by a set of monochromatic regions as it was the case for noninteracting atoms, at least when the induced electric dipolar moments are real.

The paper is organized as follows: Section II derives the model for a system of NaN_{a} identical nn-level atoms interacting with \ell modes of an electromagnetic field, including the atomic dipole-dipole interaction, and particularizes it for 22- and 33-level atoms. Section III constructs the variational energy surface from a complete set of test states which approach the quantum ground state, or any other quantum excited state. Here we focus on the ground state and study its phase diagram. This is applied in section IV to the case of 22-level atoms, and in section V to the case of 33-level atoms in their different atomic configurations, finding the critical values of the coupling parameters which minimize the energy, and determining the phase diagram in both attractive and repulsive scenarios of the atomic dipole-dipole interaction. In cases where a phase transition exists which defies the Ehrenfest classification, new criteria are proposed, one based on the second Casimir operator and another one based on the maximum Bures distance between neighboring states. Finally, section VI summarizes some conclusions. Two appendices present the matter collective operators and the atomic dipole-dipole operator in explicit form.

II Model

Refer to caption
Refer to caption
Figure 1: (Color online) Left: Schematic of the atomic dipole-dipole interaction. xi(i=a,b)x_{i}\ (i=a,b) denote the position of the dipoles di=eri\vec{d}_{i}=e\vec{r_{i}}, and R\vec{R} their separation. Right: Schematic depiction of the atomic transitions |j,k|j,k|\,j,k\rangle\leftrightarrow|\,j^{\prime},k^{\prime}\rangle due to the dipole-dipole interaction.

We consider a system of NaN_{a} identical nn-level atoms interacting with \ell modes of a radiation field, placed in a cavity. The Hamiltonian is composed of three terms

𝑯=𝑯D+𝑯mf+𝑯dd,{\bm{H}}={\bm{H}}_{D}+{\bm{H}}_{mf}+{\bm{H}}_{dd}\,, (1)

where 𝑯D{\bm{H}}_{D} is the diagonal contribution given by (=1\hbar=1)

𝑯D=s=1Ωs𝝂s+k=1nωk𝑨kk;{\bm{H}}_{D}=\sum_{s=1}^{\ell}\Omega_{s}\ {\bm{\nu}}_{s}+\sum_{k=1}^{n}\omega_{k}{\bm{A}}_{kk}\,; (2)

the dipolar matter-field interaction is of the form Cordero et al. (2013)

𝑯mf=1Nas=1j<knμjk(s)(𝑨jk+𝑨kj)(𝒂s+𝒂s).{\bm{H}}_{mf}=-\frac{1}{\sqrt{N_{a}}}\sum_{s=1}^{\ell}\sum_{j<k}^{n}\mu_{jk}^{(s)}\left({\bm{A}}_{jk}+{\bm{A}}_{kj}\left)\right({\bm{a}}_{s}^{\dagger}+{\bm{a}}_{s}\right)\,. (3)

In these expressions, Ωs\Omega_{s} and 𝝂s{\bm{\nu}}_{s} are the field frequency and photon number operator, respectively, of mode ss; ωk\omega_{k} denotes the energy of the kk-th atomic level with the convention ωj<ωk\omega_{j}<\omega_{k} for j<kj<k); 𝒂s{\bm{a}}_{s}^{\dagger} and 𝒂s{\bm{a}}_{s} are the field creation and annihilation operators; and 𝑨jk{\bm{A}}_{jk} is the atomic transition operator between levels kk and jj, which in a bosonic representation 𝑨jk=𝒃j𝒃k{\bm{A}}_{jk}={\bm{b}}_{j}^{\dagger}{\bm{b}}_{k} plays the role of the collective matter operator obeying the unitary algebra in nn dimensions, U(n)U(n); here, 𝒃j{\bm{b}}^{\dagger}_{j} creates an atom in level jj and 𝒃k{\bm{b}}_{k} annihilates one in level kk (cf. Eq. (54). The dipolar matter-field coupling intensity between field mode ss and atomic dipole formed by levels jj and kk is denoted by μjk(s)\mu_{jk}^{(s)}.

Atoms do not have permanent dipole moments in their ground state, as the center of charge of the electronic cloud coincides with that of the nucleus. In the presence of an electromagnetic field, however, these centers are displaced and the induced transition dipole moments are responsible for an atomic dipole-dipole interaction. In second quantization this dipole-dipole interaction takes the form

𝑯dd=12(Na1)j,k,j,kj,k|𝑾ab|j,k𝒃j𝒃k𝒃j𝒃k,{\bm{H}}_{dd}=\frac{1}{2(N_{a}-1)}\,\sum_{j,k,j^{\prime},k^{\prime}}\,\langle j,\,k\,|\,{\bm{W}}_{ab}\,|\,j^{\prime},k^{\prime}\rangle\,{\bm{b}}^{\dagger}_{j}\,{\bm{b}}^{\dagger}_{k}\,{\bm{b}}_{j^{\prime}}\,{\bm{b}}_{k^{\prime}}\,, (4)

where bosonic creation 𝒃j{\bm{b}}_{j}^{\dagger} and annihilation 𝒃k{\bm{b}}_{k} operators were used. The factor 1/21/2 in eq. (4) compensates the double accounting in the summation, as the particles are indistinguishable. The factor (Na1)(N_{a}-1) is included to have an interaction linear in the number of particles. The first index in the bra and the ket states corresponds to the first particle, while the second index corresponds to the second particle.

The set of operators that appear in (4) may be rewritten in terms of the collective matter operators, by means the bosonic commutation relation, as

𝒃j𝒃k𝒃j𝒃k\displaystyle{\bm{b}}^{\dagger}_{j}\,{\bm{b}}^{\dagger}_{k}\,{\bm{b}}_{j^{\prime}}\,{\bm{b}}_{k^{\prime}} =\displaystyle= 𝑨jj𝑨kkδjk𝑨jk\displaystyle{\bm{A}}_{jj^{\prime}}{\bm{A}}_{kk^{\prime}}-\delta_{j^{\prime}k}{\bm{A}}_{jk^{\prime}} (5)
:=\displaystyle:= 𝑨jj𝑨kk,\displaystyle{\bm{A}}_{jj^{\prime}}\oslash{\bm{A}}_{kk^{\prime}}\,,

where we have defined the oslash product between collective matter operators, which removes the self-interaction terms (see appendix A for more details).

The dipole-dipole interaction 𝑯dd{\bm{H}}_{dd} is obtained from the classical expression Jackson (1999) through the standard quantization procedure, which has the form

𝑾ab=𝒅a𝒅b3(n^𝒅a)(n^𝒅b)4πϵ0R3,{\bm{W}}_{ab}=\frac{\vec{{\bm{d}}}_{a}\cdot\vec{{\bm{d}}}_{b}-3(\hat{n}\cdot\vec{{\bm{d}}}_{a})(\hat{n}\cdot\vec{{\bm{d}}}_{b})}{4\pi\epsilon_{0}R^{3}}\,, (6)

where 𝒅i=e𝒓i,(i=a,b)\vec{{\bm{d}}}_{i}=e\,\vec{{\bm{r}}}_{i},\ (i=a,b) are the induced vector operators of the electric dipole moments, RR is the separation between the dipoles and n^=R/R\hat{n}=\vec{R}/R (with R=xbxa\vec{R}=\vec{x}_{b}-\vec{x}_{a}) the unitary vector in the direction from one dipole to another, at positions xa\vec{x}_{a} and xb\vec{x}_{b} (see Fig. 1). ϵ0\epsilon_{0} is the permittivity of vacuum, and for induced magnetic moments μ1,μ2\mu_{1},\,\mu_{2}, the expression is the same with the replacements diμid_{i}\to\mu_{i}, and ϵ0μ0\epsilon_{0}\to\mu_{0}, the magnetic permeability of vacuum (cf., e.g., Cohen-Tannoudji et al. (heim)). Without loss of generality, we will consider here that the induced electric dipoles are real.

Thus, the two-body matrix elements in Hilbert space are

gjjkk=14πϵ0j,k|𝒅a𝒅b3(n^𝒅a)(n^𝒅b)R3|j,k,g_{jj^{\prime}kk^{\prime}}=\frac{1}{4\,\pi\,\epsilon_{0}}\left\langle j,\,k\,\left|\,\frac{\vec{{\bm{d}}}_{a}\cdot\vec{{\bm{d}}}_{b}\,-3\,(\hat{n}\cdot\vec{{\bm{d}}}_{a})(\hat{n}\cdot\vec{{\bm{d}}}_{b})}{R^{3}}\,\right|\,j^{\prime},\,k^{\prime}\right\rangle\,, (7)

The indices of the dipole-dipole coefficient gjjkkg_{jj^{\prime}kk^{\prime}} refer to the two dipoles involved in the bra-ket (7).

For indistinguishable particles, and identifying the expansion components djk=j|𝒅a,b|k\vec{d}_{jk}=\langle j|\vec{{\bm{d}}}_{a,b}|k\rangle of the dipolar operator in terms of the collective matter operators, viz. the dipolar operator given by 𝒅=jkndjk𝑨jk\vec{{\bm{d}}}=\sum_{j\neq k}^{n}\vec{d}_{jk}{\bm{A}}_{jk}, the matrix element in Eq. (7) reads

gjjkk=djjdkk3(n^djj)(n^dkk)4πϵ0R3,g_{jj^{\prime}kk^{\prime}}=\frac{\vec{d}_{jj^{\prime}}\cdot\vec{d}_{kk^{\prime}}-3(\hat{n}\cdot\vec{d}_{jj^{\prime}})(\hat{n}\cdot\vec{d}_{kk^{\prime}})}{4\pi\epsilon_{0}R^{3}}\,, (8)

where RR stands for the average distance between pairs of atoms. The hermiticity of Eq.(4) follows from the relations

gjklm=glmjk,gjklm=gkjml.g_{jklm}=g_{lmjk}\,,\qquad g_{jklm}=g_{kjml}^{*}\,. (9)

Also, for real dipolar vectors djk=dkj\vec{d}_{jk}=\vec{d}_{kj}, one has gjklm=gjkml=gkjlmg_{jklm}=g_{jkml}=g_{kjlm}.

Finally, using the oslash operator introduced above, we may write the dipole-dipole interaction in a simplified form as

𝑯dd=12(Na1)jknlmngjklm𝑨jk𝑨lm.{\bm{H}}_{dd}=\frac{1}{2(N_{a}-1)}\sum_{j\neq k}^{n}\sum_{l\neq m}^{n}g_{jklm}{\bm{A}}_{jk}\oslash{\bm{A}}_{lm}\,. (10)

Inserting the different contributions into (10), one may write the atomic dipole-dipole term in the Hamiltonian as (see appendix B)

𝑯dd=12!jkn𝑾jk2levels+12!jkln𝑾jlk3levels+14!jklmn𝑾jklm4levels,\displaystyle{\bm{H}}_{dd}=\frac{1}{2!}\sum_{j\neq k}^{n}{\bm{W}}_{jk}^{2-{\rm levels}}+\frac{1}{2!}\sum_{j\neq k\neq l}^{n}{\bm{W}}_{jlk}^{3-{\rm levels}}+\frac{1}{4!}\sum_{j\neq k\neq l\neq m}^{n}{\bm{W}}_{jklm}^{4-{\rm levels}}\,, (11)

where the operator 𝑾jk2levels{\bm{W}}_{jk}^{2-{\rm levels}} stands for the dipole-dipole contribution of the pair djkdjk\vec{d}_{jk}\rightleftharpoons\vec{d}_{jk}; the operator 𝑾jlk3levels{\bm{W}}_{jlk}^{3-{\rm levels}} for that of the pair of dipoles djldlk\vec{d}_{jl}\rightleftharpoons\vec{d}_{lk} (here the atomic level ωl\omega_{l} plays the role of an intermediate level, so a prohibited dipolar transition djk=0\vec{d}_{jk}=0 is possible via the permitted dipolar transitions djl0\vec{d}_{jl}\neq 0 and dlk0\vec{d}_{lk}\neq 0); and the operator 𝑾jklm4levels{\bm{W}}_{jklm}^{4-{\rm levels}} corresponds to the contribution of isolated dipoles djkdlm\vec{d}_{jk}\rightleftharpoons\vec{d}_{lm} (which do not share an energy level). The upper index denotes the number of different atomic levels which contribute to the interaction; hence, the terms 𝑾jlk3levels{\bm{W}}_{jlk}^{3-{\rm levels}} and 𝑾jklm4levels{\bm{W}}_{jklm}^{4-{\rm levels}} are zero for nn-level atoms with n=2n=2 and n3n\leq 3, respectively. The set of transitions included in each interaction term is given in table 1, and shown schematically in figure 2. These terms are given in appendix B. Also, the factors 1/2!1/2! and 1/4!1/4! in expression (11) eliminate the double summation due to index reordering.

Table 1: Contribution of the atomic transitions to the terms 𝑾jk2levels,𝑾jkl3levels{\bm{W}}_{jk}^{\rm 2-levels},\,{\bm{W}}_{jkl}^{\rm 3-levels} and 𝑾jklm4levels{\bm{W}}_{jklm}^{\rm 4-levels} of the atomic dipole-dipole interaction. See also the accompanying figure 2.
aaAtomic levels aa aa Interaction aa aaaaAtomic Transitions aaaa
ωj,ωk\omega_{j}\,,\omega_{k} 𝑾jk2levels{\bm{W}}_{jk}^{\rm 2-levels} jkj\rightrightarrows k jkj\leftleftarrows k
jkjj\rightarrow k\rightarrow j kjkk\leftarrow j\leftarrow k
ωj,ωk,ωl\omega_{j}\,,\omega_{k}\,,\omega_{l} 𝑾jkl3levels{\bm{W}}_{jkl}^{\rm 3-levels} jklj\leftarrow k\rightarrow l jklj\rightarrow k\leftarrow l
jklj\leftarrow k\leftarrow l jklj\rightarrow k\rightarrow l
ωj,ωk;ωl,ωm\omega_{j}\,,\omega_{k}\,;\omega_{l}\,,\omega_{m} 𝑾jklm4levels{\bm{W}}_{jklm}^{\rm 4-levels} jk;lmj\rightleftarrows k;\,l\rightleftarrows m jk;mlj\rightleftarrows k;\,m\rightleftarrows l
jl;kmj\rightleftarrows l;\,k\rightleftarrows m jl;mkj\rightleftarrows l;\,m\rightleftarrows k
aaaa jm;klj\rightleftarrows m;\,k\rightleftarrows laaaa aaaa jm;lkj\rightleftarrows m;\,l\rightleftarrows kaaaa
Refer to caption
Figure 2: (Color online) The set of transitions involved in each term, 𝑾jk2levels,𝑾jkl3levels{\bm{W}}_{jk}^{\rm 2-levels},\,{\bm{W}}_{jkl}^{\rm 3-levels} and 𝑾jklm4levels{\bm{W}}_{jklm}^{\rm 4-levels}, of the dipole-dipole operator are shown schematically. Each transition is indicated by arrows of the same color. So for two-level atoms (diagram on the left), the blue lines denote the transition jkj\rightrightarrows k, the green lines the transition jkj\leftleftarrows k, the indigo line the transition jkjj\rightarrow k\rightarrow j, and the orange line the transition kjkk\leftarrow j\leftarrow k. They are also cumulative; thus, for instance, for the transitions in 𝑾jkl3levels{\bm{W}}_{jkl}^{\rm 3-levels} we have those shown in the diagram on the left plus those in the diagram in the middle; similarly, 𝑾jklm4levels{\bm{W}}_{jklm}^{\rm 4-levels} contains all the transitions in the three diagrams.

Amongst the parameters in the Hamiltonian, we are free to choose ω1=0\omega_{1}=0 and ωn=1\omega_{n}=1, i.e., the energies are normalized to the highest atomic level. We also consider systems where only one field mode promotes the transition between a given pair of atomic levels; this constriction is imposed by the condition Cordero et al. (2016)

ifμjk(s)0thenμjk(s)=0for allss;\textrm{if}\quad\mu_{jk}^{(s)}\neq 0\quad\textrm{then}\quad\mu_{jk}^{(s^{\prime})}=0\quad\textrm{for all}\quad s^{\prime}\neq s\,; (12)

Since the interaction (11) involves the dipole-dipole contribution gjklmg_{jklm}, we have gjklm0g_{jklm}\neq 0 only when μjk(s)0\mu_{jk}^{(s)}\neq 0 and μlm(s)0\mu_{lm}^{(s^{\prime})}\neq 0 for at least one of the modes Ωs\Omega_{s} and Ωs\Omega_{s^{\prime}}. The first and second order Casimir operators [Eqs. (52 and 53)], whose eigenvalues are functions of the number of atoms NaN_{a} and the number of levels nn, are of course constants of motion.

As an example, we now write explicitly the contribution of 𝑯dd{\bm{H}}_{dd} for two- and three-level atoms:

22-level atoms

For a system of two-level atoms, the Hamiltonian (1) reads

𝑯=𝑯D+𝑯mf+𝑾jk2levels,{\bm{H}}={\bm{H}}_{D}+{\bm{H}}_{mf}+{\bm{W}}_{jk}^{\rm 2-levels}\,, (13)

where we fix j<kj<k for the atomic levels ωj<ωk\omega_{j}<\omega_{k} respectively.

33-level atoms

33-level atoms present three different atomic configurations (Ξ\Xi, Λ\Lambda and VV), according to which atomic transitions are prohibited.

  • For the Ξ\Xi-configuration, the dipolar transition d13=0\vec{d}_{13}=0 is prohibited, and the Hamiltonian takes the form

    𝑯Ξ\displaystyle{\bm{H}}_{\Xi} =\displaystyle= 𝑯D+𝑯mf\displaystyle{\bm{H}}_{D}+{\bm{H}}_{mf} (14)
    +\displaystyle+ 𝑾122levels+𝑾232levels+𝑾1233levels.\displaystyle{\bm{W}}_{12}^{2-{\rm levels}}+{\bm{W}}_{23}^{2-{\rm levels}}+{\bm{W}}_{123}^{3-{\rm levels}}\,.

    The intermediate atomic level ω2\omega_{2} may promote the transition ω1ω3\omega_{1}\rightleftharpoons\omega_{3}. The set of nonzero dipolar-dipolar strengths is {g1212,g1221,g2323,g2332,g1232,g1223}\{g_{1212},g_{1221},g_{2323},g_{2332},g_{1232},g_{1223}\} together with their complex conjugates, obtained as gjklm=gkjmlg_{jklm}=g_{kjml}^{*}, cf. Eq. (9).

  • For the Λ\Lambda-configuration it is the dipolar transition d12=0\vec{d}_{12}=0 which is prohibited, and the Hamiltonian takes the form

    𝑯Λ\displaystyle{\bm{H}}_{\Lambda} =\displaystyle= 𝑯D+𝑯mf\displaystyle{\bm{H}}_{D}+{\bm{H}}_{mf} (15)
    +\displaystyle+ 𝑾132levels+𝑾232levels+𝑾1323levels.\displaystyle{\bm{W}}_{13}^{2-{\rm levels}}+{\bm{W}}_{23}^{2-{\rm levels}}+{\bm{W}}_{132}^{3-{\rm levels}}\,.

    The atomic level ω3\omega_{3} serves as an intermediate level which may promote the transition ω1ω2\omega_{1}\rightleftharpoons\omega_{2}. The set of nonzero dipolar-dipolar strengths is {g1313,g1331,g2323,g2332,g1323,g1332}\{g_{1313},g_{1331},g_{2323},g_{2332},g_{1323},g_{1332}\} together with their complex conjugates obtained by Eq. (9).

  • For the VV-configuration the dipolar transition d23=0\vec{d}_{23}=0 is prohibited, and the Hamiltonian is

    𝑯V\displaystyle{\bm{H}}_{V} =\displaystyle= 𝑯D+𝑯mf\displaystyle{\bm{H}}_{D}+{\bm{H}}_{mf} (16)
    +\displaystyle+ 𝑾122levels+𝑾132levels+𝑾2133levels.\displaystyle{\bm{W}}_{12}^{2-{\rm levels}}+{\bm{W}}_{13}^{2-{\rm levels}}+{\bm{W}}_{213}^{3-{\rm levels}}\,.

    The atomic level ω1\omega_{1} acts here as an intermediate and may promote the transition ω2ω3\omega_{2}\rightleftharpoons\omega_{3}. The set of nonzero dipolar-dipolar strengths is {g1212,g1221,g1313,g1331,g2131,g2113}\{g_{1212},g_{1221},g_{1313},g_{1331},g_{2131},g_{2113}\} together with their complex conjugates obtained by Eq. (9).

In a recent work Civitarese et al. (2010) the case where equal contributions of the form gjppj=gjppk=gg_{jppj}=g_{jppk}=g for all j,kj,\,k, was considered (other terms were neglected). It was shown that the dipole–-dipole interactions act against the appearance of atomic squeezing, and also that an increase in the mean value of the number of photons of the initial state smears out the effect.

III Variational energy surface

The variational solution involves a test state which approaches the quantum ground or desired excited state, and which depends on a set of parameters ziz_{i}. The corresponding energy surface is obtained by taking the expectation value of the Hamiltonian and minimizing with respect to the parameters ziz_{i} of the test state. In this work we focus on the ground state and take as test state the direct product of coherent states for both the matter and the field contributions. Clearly, this test state presents no matter-field entanglement, but it yields a good description of the minimum energy surface, as well as some expectation values of the physical quantities, and the phase diagram together with the order of the phase transitions.

Coherent matter state

The coherent matter state is defined as Iachello and Arima (1987)

|γ=1Na![𝚪]Na|0m,|\vec{\gamma}\rangle=\frac{1}{\sqrt{N_{a}!}}\left[{\bm{\Gamma}}^{\dagger}\right]^{N_{a}}|0\rangle_{m}\,, (17)

where γ=(γ1,,γn)\vec{\gamma}=(\gamma_{1},\dots,\gamma_{n}) and γ:=(|γ1|2+|γ2|2++|γn|2)1/2||\vec{\gamma}||:=(|\gamma_{1}|^{2}+|\gamma_{2}|^{2}+\cdots+|\gamma_{n}|^{2})^{1/2}. The operator 𝚪{\bm{\Gamma}}^{\dagger} is

𝚪=γ1𝒃1+γ2𝒃2++γn𝒃nγ,{\bm{\Gamma}}^{\dagger}=\frac{\gamma_{1}{\bm{b}}_{1}^{\dagger}+\gamma_{2}{\bm{b}}_{2}^{\dagger}+\cdots+\gamma_{n}{\bm{b}}_{n}^{\dagger}}{||\vec{\gamma}||}\,, (18)

and, using the bosonic realization [𝒃j,𝒃k]=δjk[{\bm{b}}_{j},{\bm{b}}_{k}^{\dagger}]=\delta_{jk} it is immediate that the relationship [𝚪,𝚪]=1[{\bm{\Gamma}},{\bm{\Gamma}}^{\dagger}]=1 is fulfilled; hence, the state (17) is normalised. It is straightforward to show that

[𝒃k,𝚪]=γkγ,\left[{\bm{b}}_{k},{\bm{\Gamma}}^{\dagger}\right]=\frac{\gamma_{k}}{||\vec{\gamma}||}\,, (19)

which for any number of atoms generalizes to

[𝒃k,(𝚪)Na]=Naγkγ(𝚪)Na1.\left[{\bm{b}}_{k},({\bm{\Gamma}}^{\dagger})^{N_{a}}\right]=N_{a}\frac{\gamma_{k}}{||\vec{\gamma}||}({\bm{\Gamma}}^{\dagger})^{N_{a}-1}\,. (20)

The relations above are useful in order to find the matrix elements of the collective matter operators. The linear contribution is

γ|𝑨jk|γ=Naγjγkγ2,\langle\vec{\gamma}|{\bm{A}}_{jk}|\vec{\gamma}\rangle=N_{a}\frac{\gamma_{j}^{*}\gamma_{k}}{||\vec{\gamma}||^{2}}\,, (21)

and the quadratic contribution

γ|𝑨jk𝑨lm|γ=\displaystyle\langle\vec{\gamma}|{\bm{A}}_{jk}{\bm{A}}_{lm}|\vec{\gamma}\rangle= Na(Na1)γjγkγlγmγ4\displaystyle N_{a}(N_{a}-1)\frac{\gamma_{j}^{*}\gamma_{k}\gamma_{l}^{*}\gamma_{m}}{||\vec{\gamma}||^{4}} (22)
+δklNaγjγmγ2,\displaystyle+\delta_{kl}N_{a}\frac{\gamma_{j}^{*}\gamma_{m}}{||\vec{\gamma}||^{2}}\,,

where the last term corresponds to the self-interactions, and vanishes in the dipole-dipole interaction.

Coherent field state

The coherent field state for \ell modes is given by the direct product of coherent states for each mode, as follows Klauder and Skagerstam (1985); Ali et al. (2014),

|α:=|α1|α1|α,|\vec{\alpha}\rangle:=|\alpha_{1}\rangle\otimes|\alpha_{1}\rangle\otimes\cdots\otimes|\alpha_{\ell}\rangle\,, (23)

where α={α1,,α}\vec{\alpha}=\{\alpha_{1},\dots,\alpha_{\ell}\}. For each mode s=1,,s=1,\dots,\ell the coherent state satisfies 𝒂s|αs=αs|αs{\bm{a}}_{s}|\alpha_{s}\rangle=\alpha_{s}|\alpha_{s}\rangle, and hence

α|𝒂s|α=αs,α|𝒂s|α=αs,\langle\vec{\alpha}|{\bm{a}}_{s}|\vec{\alpha}\rangle=\alpha_{s}\,,\qquad\langle\vec{\alpha}|{\bm{a}}_{s}^{\dagger}|\vec{\alpha}\rangle=\alpha_{s}^{*}\,, (24)

while the expectation value of the number operator for each mode is

α|𝒂s𝒂s|α=α|𝝂s|α=|αs|2.\langle\vec{\alpha}|{\bm{a}}_{s}^{\dagger}{\bm{a}}_{s}|\vec{\alpha}\rangle=\langle\vec{\alpha}|{\bm{\nu}}_{s}|\vec{\alpha}\rangle=|\alpha_{s}|^{2}\,. (25)

From the expressions above, and writing for the complete test state the direct product of the coherent states for field and matter,

|α,γ:=|α|γ,|\vec{\alpha},\vec{\gamma}\rangle:=|\vec{\alpha}\rangle\otimes|\vec{\gamma}\rangle\,, (26)

the variational energy surface per atom :=α,γ|𝑯|α,γ/Na{\cal E}:=\langle\vec{\alpha},\vec{\gamma}|{\bm{H}}|\vec{\alpha},\vec{\gamma}\rangle/N_{a}, as a function of αs=Rseiθs\alpha_{s}=R_{s}e^{i\theta_{s}}, γk=ϱkeiϕk\gamma_{k}=\varrho_{k}e^{i\phi_{k}}, and parameters of the Hamiltonian, reads

\displaystyle{\cal E} =\displaystyle= 1Nas=1ΩsRs2+k=1nωkϱk2γ2\displaystyle\frac{1}{N_{a}}\sum_{s=1}^{\ell}\Omega_{s}\ R_{s}^{2}+\sum_{k=1}^{n}\omega_{k}\frac{\varrho_{k}^{2}}{||\vec{\gamma}||^{2}} (27)
\displaystyle- 4Nas=1j<knμjk(s)ϱjϱkRsγ2cos(ϕjk)cos(θs)\displaystyle\frac{4}{\sqrt{N_{a}}}\sum_{s=1}^{\ell}\sum_{j<k}^{n}\mu_{jk}^{(s)}\frac{\varrho_{j}\varrho_{k}R_{s}}{||\vec{\gamma}||^{2}}\cos(\phi_{jk})\cos(\theta_{s})
+\displaystyle+ 1Naα,γ|𝑯dd|α,γ,\displaystyle\frac{1}{N_{a}}\langle\vec{\alpha},\vec{\gamma}|{\bm{H}}_{dd}|\vec{\alpha},\vec{\gamma}\rangle\,,\quad

where ϕjl=ϕlϕj\phi_{jl}=\phi_{l}-\phi_{j}. The last term in (27) corresponds to the atomic dipole-dipole interaction per particle dd{\cal E}_{dd}, and has the form

dd\displaystyle{\cal E}_{dd} =\displaystyle= 1γ4j<kRe[gjkjke2iϕjk+gjkkj]ϱj2ϱk2+2γ4j<k;jpkRe[gjpkpei(ϕjp+ϕkp)+gjppkeiϕjk]ϱjϱkϱp2\displaystyle\frac{1}{||\vec{\gamma}||^{4}}\sum_{j<k}{\rm Re}\,[g_{jkjk}\,e^{2i\phi_{jk}}+\,g_{jkkj}]\,\varrho_{j}^{2}\varrho_{k}^{2}+\frac{2}{||\vec{\gamma}||^{4}}\sum_{j<k;j\neq p\neq k}{\rm Re}\,[g_{jpkp}\,e^{i(\phi_{jp}+\phi_{kp})}+g_{jppk}\,e^{i\phi_{jk}}]\,\varrho_{j}\varrho_{k}\varrho_{p}^{2} (28)
+\displaystyle+ 2γ4j<k<l<mRe[gjklmei(ϕjk+ϕlm)+gjkmlei(ϕjkϕlm)+gjlkmei(ϕjl+ϕkm)+gjlmkei(ϕjlϕkm)\displaystyle\frac{2}{||\vec{\gamma}||^{4}}\sum_{j<k<l<m}{\rm Re}\,[g_{jklm}\,e^{i(\phi_{jk}+\phi_{lm})}+g_{jkml}\,e^{i(\phi_{jk}-\phi_{lm})}+g_{jlkm}\,e^{i(\phi_{jl}+\phi_{km})}+g_{jlmk}\,e^{i(\phi_{jl}-\phi_{km})}
+gjmklei(ϕjm+ϕkl)+gjmlkei(ϕjmϕkl)]ϱjϱkϱlϱm.\displaystyle+g_{jmkl}\,e^{i(\phi_{jm}+\phi_{kl})}+g_{jmlk}\,e^{i(\phi_{jm}-\phi_{kl})}]\,\varrho_{j}\varrho_{k}\varrho_{l}\varrho_{m}\,.

Here, we used the fact that gjklm=glmjkg_{jklm}=g_{lmjk} and gjklm=gkjmlg_{jklm}=g_{kjml}^{*} in order to simplify the expression.

By simple inspection, one may note that the energy surface has minima at the critical values θsc=0,π\theta_{s}^{c}=0,\pi and Rsc=NarscR_{s}^{c}=\sqrt{N_{a}}\,r_{s}^{c}, with

rsc=2j<knμjk(s)Ωsϱjcϱkcγc2cos(ϕjkc)cos(θsc),rsc0,r_{s}^{c}=2\sum_{j<k}^{n}\frac{\mu_{jk}^{(s)}}{\Omega_{s}}\frac{\varrho_{j}^{c}\varrho_{k}^{c}}{||\vec{\gamma}^{c}||^{2}}\cos(\phi_{jk}^{c})\cos(\theta_{s}^{c})\,,\quad r_{s}^{c}\geq 0\,, (29)

and ϕjlc=ϕlcϕjc\phi_{jl}^{c}=\phi_{l}^{c}-\phi_{j}^{c}.

In a similar fashion, for the fixed values ϱ1=1\varrho_{1}=1 and ϕ1=0\phi_{1}=0, and supposing, without loss of generality, real values for the dipolar (μ\mu) and dipole-dipole (gjklmg_{jklm}) strengths, one finds the critical values for the phase ϕj\phi_{j} to be ϕjc=0,π\phi_{j}^{c}=0,\,\pi.

After substitution of the critical values θsc,ϕjc\theta_{s}^{c},\,\phi_{j}^{c}, and fixing ϱ1=1\varrho_{1}=1 and ϕ1=0\phi_{1}=0, we obtain a family of energy surfaces E(ϱ;θc,ϕc)E(\varrho;\theta^{c},\phi^{c}) for ϱ=(ϱ2ϱn),θc=(θ1cθc)\varrho=(\varrho_{2}\,\dots\,\varrho_{n}),\,\theta^{c}=(\theta^{c}_{1}\,\dots\,\theta^{c}_{\ell}) and ϕc=(ϕ2cϕnc)\phi^{c}=(\phi^{c}_{2}\,\dots\,\phi^{c}_{n}); appropriate values for θsc\theta_{s}^{c} and ϕjc\phi_{j}^{c} should be selected in order to satisfy rsc0r_{s}^{c}\geq 0 in Eq. (29). The minimum energy surface is then obtained by calculating the critical points ϱjc\varrho_{j}^{c}, which is done numerically in general.

IV Two-level atoms

For two-level atoms the expression of the energy surface reads

\displaystyle{\cal E} =\displaystyle= Ωsrs2+ωjϱj2+ωkϱk2ϱj2+ϱk24μjkrsϱjϱkcos(θs)cos(ϕjk)ϱj2+ϱk2\displaystyle\Omega_{s}\,r_{s}^{2}+\frac{\omega_{j}\,\varrho_{j}^{2}+\omega_{k}\,\varrho_{k}^{2}}{\varrho_{j}^{2}+\varrho_{k}^{2}}-4\frac{\mu_{jk}\,r_{s}\varrho_{j}\varrho_{k}\cos(\theta_{s})\cos(\phi_{jk})}{\varrho_{j}^{2}+\varrho_{k}^{2}} (30)
+\displaystyle+ [gjkjkcos(2ϕjk)+gjkkj]ϱj2ϱk2(ϱj2+ϱk2)2,\displaystyle[g_{jkjk}\,\cos(2\,\phi_{jk})+\,g_{jkkj}]\,\frac{\varrho_{j}^{2}\varrho_{k}^{2}}{(\varrho_{j}^{2}+\varrho_{k}^{2})^{2}}\,,

with j<kj<k. The critical values of the corresponding energy surface Eq. (27) must satisfy

μjkcos(ϕjkc)cos(θsc)=|μjk|,\mu_{jk}\cos(\phi_{jk}^{c})\cos(\theta_{s}^{c})=|\mu_{jk}|\,, (31)

and, fixing ϕj=0\phi_{j}=0 and ϱj=1\varrho_{j}=1, one finds two solutions

ϱkc=0,andϱkc=xjk2yjkxjk2yjk+2.\displaystyle\varrho_{k}^{c}=0,\quad\textrm{and}\quad\varrho_{k}^{c}=\sqrt{\frac{x_{jk}^{2}-y_{jk}}{x_{jk}^{2}-y_{jk}+2}}\,. (32)

Here, we have used a dimensionless matter-field coupling intensity xjkx_{jk}, defined as

xjk=μjkμjkc,μjkc=12Ωsωjk,x_{jk}=\frac{\mu_{jk}}{\mu_{jk}^{c}}\,,\qquad\mu_{jk}^{c}=\frac{1}{2}\sqrt{\Omega_{s}\,\omega_{jk}}\,, (33)

with μjkc\mu_{jk}^{c} the critical value of the coupling constant when the atomic dipole-dipole interaction is neglected, and where ωjk=|ωkωj|\omega_{jk}=|\omega_{k}-\omega_{j}|; we have also defined

yjk=ωjk+gωjk,g=gjkkj+gjkjk,y_{jk}=\frac{\omega_{jk}+g}{\omega_{jk}}\,,\quad g=g_{jkkj}+g_{jkjk}\,, (34)

to simplify the notation.

For values xjk2yjk0x_{jk}^{2}-y_{jk}\leq 0 one finds only one critical value ϱk=0\varrho_{k}=0, for which the energy surface has the constant value ωj\omega_{j}. When xjk2yjk>0x_{jk}^{2}-y_{jk}>0 we have two critical values (32), in this case the energy surface has a dependence on the matter-field dipolar strength xjkx_{jk}. After minimizing one finds

Emin={ωj;xjk2<yjkωj[xjk2yjk]24(xjk2yjk+1)ωjk;xjk2yjk,E_{\rm min}=\left\{\begin{array}[]{l l}\omega_{j};&x_{jk}^{2}<y_{jk}\\[8.53581pt] \omega_{j}-\displaystyle{\frac{[x_{jk}^{2}-y_{jk}]^{2}}{4(x_{jk}^{2}-y_{jk}+1)}}\omega_{jk};&x_{jk}^{2}\geq y_{jk}\end{array}\right.\,, (35)

This relationship, EminE_{\rm min} vs. xjkx_{jk}, is shown in figure 3. The solid line (black) corresponds to the case g=0g=0 without atomic dipole-dipole interaction; repulsive g=0.05g=0.05 (dashed line, blue) and attractive g=0.05g=-0.05 (dotted line, green) cases are also shown. Due to the minuteness of this interaction when compared with the dipolar matter-field interaction, the difference for dissimilar values of gg is difficult to appreciate. We have zoomed around the value xjk=1x_{jk}=1 (see figure inset) where the transition into the collective region appears, in order to make this difference clear. In what follows, we will consider unnaturally large values for the atomic dipole-dipole coupling parameter gg so that its effect may be appreciated; when studying actual realistic systems these values (and their effects) must be scaled down accordingly.

Refer to caption
Figure 3: (colour online) Minimum energy as a function of the matter-field coupling xjkx_{jk}, for two-level atoms interacting with a single mode of an electromagnetic field. The solid line corresponds to the case without dipole-dipole interaction g=0g=0; the repulsive case g=0.05g=0.05 (dashed line) and the attractive case g=0.05g=-0.05 (dotted line). Inset shows a zoom around the value xjk=1x_{jk}=1 where the transition appears. The parameters are ωj=0,ωk=1\omega_{j}=0\,,\omega_{k}=1 for the atomic levels, and Ω=1\Omega=1 for the field frequency.
Refer to caption
Refer to caption
Figure 4: (colour online) (a) Minimum energy as a function of the matter-field coupling xjkx_{jk}, for two-level atoms interacting with a single mode of an electromagnetic field. The solid line corresponds to the case g=0g=0 without dipole-dipole interaction; the repulsive case g=0.5g=0.5 (dash-dot line), and two attractive cases g=0.5g=-0.5 (dashed line) and g=2g=-2 (dotted line) are also shown (the latter in a regime of very strong attractive interaction). (b) Minimum energy (dotted line) and its first (dashed line) and second (solid line) derivatives. The parameters are ωj=0,ωk=1\omega_{j}=0\,,\omega_{k}=1 for the atomic levels, and Ω=1\Omega=1 for the field frequency.

The minimum energy for different (larger) values of the dipolar coupling strength is plotted in figure 4. For values of gg such that yjk>0y_{jk}>0, the critical points xjkc=±yjkx_{jk}^{c}=\pm\sqrt{y_{jk}} divide the normal region xjk2<(xjkc)2x_{jk}^{2}<(x_{jk}^{c})^{2} from the collective region xjk2>(xjkc)2x_{jk}^{2}>(x_{jk}^{c})^{2}. One should note that for the case without dipole-dipole interaction, g=0g=0, (solid line in figure 4(a)) the critical points occur at (xjkc)2=1(x_{jk}^{c})^{2}=1, while in the attractive case g<0g<0 (dashed line in figure 4(a)) one has (xjkc)2<1(x_{jk}^{c})^{2}<1, i.e., the normal region decreases. Correspondingly, for the repulsive case g>0g>0 (dot-dash line in figure 4(a)) the normal region increases, as we have (xjkc)2>1(x_{jk}^{c})^{2}>1. The anomalous behaviour is the strong attractive regime, this is characterised by values of gg such that yjk0y_{jk}\leq 0, when the normal region vanishes completely (dotted line in Fig. 4(a)). It is important to note that, for large matter-field coupling xjk2yjkx_{jk}^{2}\gg y_{jk}, the minimum energy surface EminE_{\rm min} tends to that without the atomic dipole-dipole interaction; in other words, the effect of the dipole-dipole terms on the energy surface is seen mainly in a vicinity of the normal region.

The order of the transition may be determined using the Ehrenfest classification Gilmore (1993), which involves the derivatives of the energy surface. We exemplify the case g=0.5g=0.5 in figure 4(b), showing, respectively, the first (dashed-line) and second derivatives (solid line) of the energy. Since the second derivative presents a discontinuity at the critical point xjkcx_{jk}^{c}, a second order transition occurs at that location.

V Three-level atoms

For three-level atomic systems interacting dipolarly with a two-mode electromagnetic field in a cavity, the atomic dipole-dipole interaction can be obtained from expression (10) or (59). For the case of real induced dipole moments one has only to consider the real coupling strengths g1212,g1313,g2323g_{1212},\,g_{1313},\,g_{2323} for two-level interactions, and g1213,g1232,g1323g_{1213},\,g_{1232},\,g_{1323} for those associated to three-level interactions. Thus the induced dipole-dipole interaction for three-level atoms takes the form,

𝑯dd\displaystyle\bm{H}_{dd} =\displaystyle= g12122(Na1){(𝑨12+𝑨21)2𝑨11𝑨22}+g13132(Na1){(𝑨13+𝑨31)2𝑨11𝑨33}\displaystyle\frac{g_{1212}}{2\,(N_{a}-1)}\left\{(\bm{A}_{12}+\bm{A}_{21})^{2}-\bm{A}_{11}-\bm{A}_{22}\right\}+\frac{g_{1313}}{2\,(N_{a}-1)}\left\{(\bm{A}_{13}+\bm{A}_{31})^{2}-\bm{A}_{11}-\bm{A}_{33}\right\}
+g23232(Na1){(𝑨23+𝑨32)2𝑨22𝑨33}+g1213Na1{𝑨12𝑨13+𝑨31𝑨21+𝑨13𝑨21+𝑨12𝑨31}\displaystyle+\frac{g_{2323}}{2\,(N_{a}-1)}\left\{(\bm{A}_{23}+\bm{A}_{32})^{2}-\bm{A}_{22}-\bm{A}_{33}\right\}+\frac{g_{1213}}{N_{a}-1}\left\{\bm{A}_{12}\,\bm{A}_{13}+\bm{A}_{31}\,\bm{A}_{21}+\bm{A}_{13}\,\bm{A}_{21}+\bm{A}_{12}\,\bm{A}_{31}\right\}\,
+g1232Na1{𝑨12𝑨32+𝑨23𝑨21+𝑨23𝑨12+𝑨21𝑨32}+g1323Na1{𝑨13𝑨23+𝑨32𝑨31+𝑨32𝑨13+𝑨31𝑨23}.\displaystyle+\frac{g_{1232}}{N_{a}-1}\left\{\bm{A}_{12}\,\bm{A}_{32}+\bm{A}_{23}\,\bm{A}_{21}+\bm{A}_{23}\,\bm{A}_{12}+\bm{A}_{21}\,\bm{A}_{32}\right\}\,+\frac{g_{1323}}{N_{a}-1}\left\{\bm{A}_{13}\,\bm{A}_{23}+\bm{A}_{32}\,\bm{A}_{31}+\bm{A}_{32}\,\bm{A}_{13}+\bm{A}_{31}\,\bm{A}_{23}\right\}\,.

Notice that for the different atomic configurations one has at most three real parameters; in the case of the Λ\Lambda configuration, for instance, we have the coupling strengths g1313,g2323g_{1313},\,g_{2323} and g1323g_{1323}.

The corresponding variational energy surface for the dipole-dipole interaction may be obtained by taking the expectation value of (V) with respect the variational state |γ1,γ2,γ3|α1,α2|\gamma_{1},\,\gamma_{2},\,\gamma_{3}\rangle\otimes|\alpha_{1},\,\alpha_{2}\rangle, or from the general expression (28) by considering real induced dipole moments together with three-level atomic systems and a two-mode electromagnetic field. The resulting expressions for the Λ\Lambda, VV and Ξ\Xi atomic configuration are given by

dd(Λ)\displaystyle{\cal E}^{(\Lambda)}_{dd} =\displaystyle= g1313ρ32(1+cos2ϕ3)(1+ρ22+ρ32)2+g2323ρ32ρ22(1+cos2(ϕ3ϕ2))(1+ρ22+ρ32)2+2g1323ρ32ρ2(cos(2ϕ3ϕ2)+cosϕ2)(1+ρ22+ρ32)2,\displaystyle\frac{g_{1313}\,\rho^{2}_{3}\,(1+\cos{2\,\phi_{3}})}{(1+\rho^{2}_{2}+\rho^{2}_{3})^{2}}+\frac{g_{2323}\,\rho^{2}_{3}\,\rho^{2}_{2}\,(1+\cos{2(\phi_{3}}-\phi_{2}))}{(1+\rho^{2}_{2}+\rho^{2}_{3})^{2}}+\frac{2\,g_{1323}\,\rho^{2}_{3}\,\rho_{2}\,(\cos(2\,\phi_{3}-\phi_{2})+\cos{\phi_{2}})}{(1+\rho^{2}_{2}+\rho^{2}_{3})^{2}}\,, (37)
dd(V)\displaystyle{\cal E}^{(V)}_{dd} =\displaystyle= g1212ρ22(1+cos2ϕ2)(1+ρ22+ρ32)2+g1313ρ32(1+cos2ϕ3)(1+ρ22+ρ32)2+4g1213ρ2ρ3cosϕ2cosϕ3(1+ρ22+ρ32)2,\displaystyle\frac{g_{1212}\,\rho^{2}_{2}\,(1+\cos{2\,\phi_{2}})}{(1+\rho^{2}_{2}+\rho^{2}_{3})^{2}}+\frac{g_{1313}\,\rho^{2}_{3}\,(1+\cos{2\,\phi_{3}})}{(1+\rho^{2}_{2}+\rho^{2}_{3})^{2}}+\frac{4\,g_{1213}\,\rho_{2}\,\rho_{3}\,\cos\phi_{2}\,\cos{\phi_{3}}}{(1+\rho^{2}_{2}+\rho^{2}_{3})^{2}}\,, (38)
dd(Ξ)\displaystyle{\cal E}^{(\Xi)}_{dd} =\displaystyle= g1212ρ22(1+cos2ϕ2)(1+ρ22+ρ32)2+g2323ρ22ρ32(1+cos2(ϕ3ϕ2))(1+ρ22+ρ32)2+2g1232ρ22ρ3(cos(2ϕ2ϕ3)+cosϕ3)(1+ρ22+ρ32)2.\displaystyle\frac{g_{1212}\,\rho^{2}_{2}\,(1+\cos{2\,\phi_{2}})}{(1+\rho^{2}_{2}+\rho^{2}_{3})^{2}}+\frac{g_{2323}\,\rho^{2}_{2}\,\rho^{2}_{3}\,(1+\cos{2\,(\phi_{3}}-\phi_{2}))}{(1+\rho^{2}_{2}+\rho^{2}_{3})^{2}}+\frac{2\,g_{1232}\,\rho^{2}_{2}\,\rho_{3}\,(\cos(2\,\phi_{2}-\phi_{3})+\cos{\phi_{3}})}{(1+\rho^{2}_{2}+\rho^{2}_{3})^{2}}\,. (39)

For systems of 33-level atoms interacting with two modes of electromagnetic field, the critical values of the phases (vide supra) are θsc=0,π\theta_{s}^{c}=0,\,\pi and ϕkc=0,π\phi_{k}^{c}=0,\,\pi, for which the relationship μjk(s)cos(θsc)cos(ϕjkc)>0\mu_{jk}^{(s)}\cos(\theta_{s}^{c})\cos(\phi_{jk}^{c})>0 is satisfied, and where we defined ϕjkc=ϕkcϕjc\phi_{jk}^{c}=\phi_{k}^{c}-\phi_{j}^{c}. Also, the critical values rscr_{s}^{c} associated to the field are given as functions of the critical values ϱkc\varrho_{k}^{c} of the matter [cf. Eq. (29)]. These values must be calculated numerically, except when the dipole-dipole interaction is neglected, since in this latter case we have an analytical solution Cordero et al. (2015).

In this work we calculate the critical values for the three atomic configurations (Ξ\Xi, Λ\Lambda and VV) and obtain the corresponding separatrix; we fix in all cases the double resonant condition, i.e., the field frequencies are given by Ω1=ωjk\Omega_{1}=\omega_{jk} and Ω2=ωlm\Omega_{2}=\omega_{lm}. The atomic levels satisfy the condition ω1<ω2<ω3\omega_{1}<\omega_{2}<\omega_{3} with ω1=0\omega_{1}=0 and ω3=1\omega_{3}=1. We take (j,k,l,m)=(1,2,2,3)(j,k,l,m)=(1,2,2,3) and the value ω2=3/4\omega_{2}=3/4 for the Ξ\Xi-configuration, (j,k,l,m)=(1,3,2,3)(j,k,l,m)=(1,3,2,3) and ω2=1/4\omega_{2}=1/4 for the Λ\Lambda-configuration, and (j,k,l,m)=(1,2,1,3)(j,k,l,m)=(1,2,1,3) and ω2=3/4\omega_{2}=3/4 for the VV-configuration. The values considered for the dipolar-dipolar strength gjklmg_{jklm}, assuming real dipolar vectors djk=dkj\vec{d}_{jk}=\vec{d}_{kj}, are given in table 2.

Table 2: Values for the dipole-dipole strength g±sg_{\pm s} used in the numerical calculation of the minimum energy surface. The indices are (j,k,l,m)=(1,2,2,3)(j,k,l,m)=(1,2,2,3) for the Ξ\Xi-configuration, (j,k,l,m)=(1,3,2,3)(j,k,l,m)=(1,3,2,3) for the Λ\Lambda-configuration, and (j,k,l,m)=(1,2,1,3)(j,k,l,m)=(1,2,1,3) for the VV-configuration. We have used the relationship gjklm=gjkmlg_{jklm}=g_{jkml} assuming real dipolar vectors djk\vec{d}_{jk}.
  aaaagjkjkg_{jkjk}   aaaaglmlmg_{lmlm}   aaaagjklmg_{jklm}
g±1g_{\pm 1}   aaa±0.1\pm 0.1   aaa±0.04\pm 0.04   aaa±14105\pm 14\,\sqrt{10^{-5}}
g±2g_{\pm 2}   aaa±0.3\pm 0.3   aaa±0.2\pm 0.2   aaa±143/2×102\pm 14\,\sqrt{3/2}\times 10^{-2}
g±3g_{\pm 3}   aaa±1.0\pm 1.0   aaa±0.4\pm 0.4   aaa±140105\pm 140\,\sqrt{10^{-5}}
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 5: (colour online) VV-configuration with fixed values of g3g_{3}. (a) shows the minimum energy surface, (b) its first derivative [Eq. (40)], (c) its second derivative [Eq. (41)], and (d) the difference between the second order Casimir operators of the subsystems. Parameters used are discussed in the text.

In order to exemplify how to obtain the separatrix, we consider explicitly the particular case of the VV-configuration with a repulsive dipole-dipole strength g3g_{3}. The set of critical points ϱ2c\varrho_{2}^{c} and ϱ3c\varrho_{3}^{c} are evaluated numerically and inserted into the expression for the minimum energy; the result is shown in Fig. 5(a). The normal region, where Emin=0E_{min}=0, is colored in black. The separatrix is found by calculating the first derivatives of the energy surface, as

δE=:Exjk+Exlm,\delta E=:\frac{\partial E}{\partial x_{jk}}+\frac{\partial E}{\partial x_{lm}}\,, (40)

which is shown in Fig. 5(b). It is a continuous surface. We calculate the second order derivative as

δ2E=:δExjk+δExlm,\delta^{2}E=:\frac{\partial\delta E}{\partial x_{jk}}+\frac{\partial\delta E}{\partial x_{lm}}\,, (41)

which is discontinuous [cf. Fig. 5(c)]. The loci form a separatrix which splits the normal from the collective region; in fact, this discontinuity shows that a second order transition occurs at these points for the VV-configuration.

In Fig. 5(c), the slight undulation (observed by a small change in the orange hue of the surface) within the collective region in the second derivative of the minimum energy surface, is a signature of a kind of transition due to a change of subspaces formed by 22-level atoms, as was discussed recently for the case without dipole-dipole interaction g=0g=0 Cordero et al. (2015), from one subspace in which one of the radiation modes dominates to another subspace where the other mode dominates. This change grows as g0g\to 0, and gives a discontinuity when g=0g=0. However, for values g0g\neq 0 the second derivative remains continuous, as well as derivatives of higher order; in other words, the Ehrenfest classification does not provide a criterion to determine that the transition exists. In this work, we propose to consider the second order Casimir operator corresponding to each 22-level subsystem in order to label this transition (vide infra).

The second order Casimir operator for a system of NaN_{a} particles of nn-levels is given by

j,k=1n𝑨kj𝑨jk=Na(Na+n1).\sum_{j,k=1}^{n}{\bm{A}}_{kj}{\bm{A}}_{jk}=N_{a}\,(N_{a}+n-1)\,. (42)

In particular, when only two levels are considered, we may define

Cjk=:𝑨jj𝑨jj+𝑨jk𝑨kj+𝑨kj𝑨jk+𝑨kk𝑨kk,C_{jk}=:{\bm{A}}_{jj}{\bm{A}}_{jj}+{\bm{A}}_{jk}{\bm{A}}_{kj}+{\bm{A}}_{kj}{\bm{A}}_{jk}+{\bm{A}}_{kk}{\bm{A}}_{kk}\,, (43)

which coincides with the second order Casimir operator for 22-levels. Therefore, the expectation value ψ|Cjk|ψ\langle\psi|C_{jk}|\psi\rangle will be close to Na(Na+1)N_{a}(N_{a}+1) when the bulk of the contribution to the state |ψ|\psi\rangle is given by the basis of the sub-system of the two levels (j,k)(j,k). Since the variational solution is independent of NaN_{a}, we fix for this calculation Na=2N_{a}=2 and consider the absolute value of the difference of the second order Casimir operator of each subsystem

δC=:|ψ|CjkClm|ψ|,\delta C=:|\langle\psi|C_{jk}-C_{lm}|\psi\rangle|\,, (44)

where |ψ|\psi\rangle stands for the ground state.

This quantity is plotted in figure 5(d), showing that it is sensitive to the transition in the collective region. The points in the collective region where a transition occurs are given by δC=0\delta C=0, indicating that the bulk of the ground state changes from one sub-space to the other.

Another criterion that we have proposed Cordero et al. (2021); López-Peña et al. (2021) in order to find transitions not detectable through the Ehrenfest classification, is to use the Bures distance in the total product space of nn-level atoms and \ell-mode radiation field, defined by Bures (1969); Uhlmann (1976)

DB=21|α,γ|α,γ|2,D_{B}=\sqrt{2}\sqrt{1-|\langle\vec{\alpha},\vec{\gamma}|\vec{\alpha}^{\prime},\vec{\gamma}^{\prime}\rangle|^{2}}\,, (45)

for states

α,γ|α,γ=e(|α|2+|α|22αα)/2(γγγγ)Na,\langle\vec{\alpha},\vec{\gamma}|\vec{\alpha}^{\prime},\vec{\gamma}^{\prime}\rangle=e^{-(|\vec{\alpha}|^{2}+|\vec{\alpha}^{\prime}|^{2}-2\vec{\alpha}^{*}\cdot\vec{\alpha}^{\prime})/2}\left(\frac{\vec{\gamma}^{*}\cdot\vec{\gamma}^{\prime}}{||\vec{\gamma}||\,||\vec{\gamma}^{\prime}||}\right)^{N_{a}}\,, (46)

and maximize it for neighboring states. As a general procedure, one selects various points around a circumference of radius ε\varepsilon about each point pp in parameter space, in order to find the state with maximum distance to pp (cf. Cordero et al. (2021); López-Peña et al. (2021) for details). In our case, it was sufficient to calculate it for four points about each pp in order to get a qualitative behavior of the surface of maximum Bures distance.

Figure 6 shows, for Na=5N_{a}=5 (a) and for Na=5000N_{a}=5000 (b), the surface of maximum Bures distance between neighboring states. Note that the transition within the collective regions stands out, and for Na=5000N_{a}=5000 we reach the maximum distance of 2\sqrt{2} for variational states in the thermodynamic limit. We may refer to this as a transition of the kind continuous unstable, in the sense that this transition tends to a first order one in the limit g0g\to 0.

Refer to caption
Refer to caption
Figure 6: (Color online) Surface of maximum Bures distance between neighboring states for (a) Na=5N_{a}=5 and (b) Na=5000N_{a}=5000 particles, in the atomic VV configuration. The separatrix within the collective region, which defies an Ehrenfest-type classification, is clearly noticeable. Parameters are the same as in Fig. 5.
Refer to caption
Refer to caption
Figure 7: (colour online) Separatrices for the Ξ\Xi-configuration shown as a function of the dimensionless matter-field dipolar strength xjkx_{jk}, for values of atomic dipole-dipole strength g±1g_{\pm 1} (dotted line), g±2g_{\pm 2} (dashed line) and g±3g_{\pm 3} (dot-dash line). Figure (a) shows the repulsive, figure (b) the attractive case. Parameters used are discussed in text.
Refer to caption
Refer to caption
Figure 8: (colour online) Separatrices for the Λ\Lambda-configuration shown as a function of the dimensionless matter-field dipolar strength xjkx_{jk}, for values of atomic dipole-dipole strength g±1g_{\pm 1} (dotted line), g±2g_{\pm 2} (dashed line) and g±3g_{\pm 3} (dot-dash line). Figure (a) shows the repulsive, figure (b) the attractive case. Parameters used are discussed in text.
Refer to caption
Refer to caption
Figure 9: (colour online) Separatrices for the VV-configuration shown as a function of the dimensionless matter-field dipolar strength xjkx_{jk}, for values of atomic dipole-dipole strength g±1g_{\pm 1} (dotted line), g±2g_{\pm 2} (dashed line) and g±3g_{\pm 3} (dot-dash line). Figure (a) shows the repulsive, figure (b) the attractive case. Parameters used are discussed in text.

Figure 7 shows the separatrix for the atomic Ξ\Xi-configuration, in the case of a repulsive dipole-dipole interaction 7(a), and in the case of an attractive one attractive 7(b). One notes that, in the repulsive case, the normal region NN grows as the dipole-dipole interaction grows. The regions where the bulk of the ground state is dominated by the basis of the subsystem S12S_{12} or S23S_{23} are also indicated. The order of the phase transitions are marked: a first order transition for NS23N\leftrightarrow S_{23} and second order transition for NS12N\leftrightarrow S_{12}. In the attractive case, Fig. 7(b), the normal region decreases in size as gg increases in magnitude, and it in fact vanishes for the value of g3g_{-3} where only the regions S12S_{12} and S23S_{23} subsist; for g1g_{-1} and g2g_{-2} the order of the phase transitions is the same as in the repulsive case.

A similar behavior occurs for the Λ\Lambda-configuration, Fig. 8, where the subregions in the collective regime are S13S_{13} and S23S_{23}.

Figure 9 shows the situation for atoms in the VV-configuration. In the repulsive case, Fig 9(a), a normal region exists in all the circumstances, and the transitions from the normal to the collective region are of second order. For the attractive case, Fig 9(b), in the case g1g_{-1} (dotted line) a normal region exists and we have a second order transition. In the strong attractive cases of g2g_{-2} and g3g_{-3} we only have the collective regions S12S_{12} and S13S_{13}.

VI Conclusions

We have established the general atomic dipole-dipole interaction Hamiltonian for a system of nn-level atoms interacting with \ell-modes of electromagnetic radiation in a cavity, together with the associated energy surface, which allows to determine the variational ground state (see expressions (27), (28), and (29)). For 22- and 33-level atomic configurations, we have found that for attractive (repulsive) atomic dipole-dipole interactions the normal region decreases (increases) in size. The quantum phase diagrams, together with the corresponding order of the transitions, have also been determined. For a finite or infinite number of atoms, the surface of maximum Bures distance is able to detect the transitions between the collective regions where the Ehrenfest criterion fails (see Fig. 6). In other words, we find that, in cases where the Ehrenfest criterion for the phase transitions does not give information, a criterion based on the maximum probability for prohibited transitions comes to the rescue. We have also proved that the quantum phase diagrams continue being dominated by monochromatic regions as it is the case for noninteracting atoms, at least for real induced electric dipolar moments.

Phase diagrams for 22- and 33-level atoms interacting with an external radiation field have been studied, for all the possible atomic configurations. It is seen that the atomic dipole-dipole interaction is minuscule compared with the dipolar matter-field interaction, so the atomic dipole-dipole coupling has been exaggerated in order to see its consequences. (The unnaturally large values for this coupling, taken so that its effects may be appreciated, must be scaled down accordingly when studying actual realistic systems.) Although small, energy transfer between the particles (atoms, molecules) is one of the important consequences of this interaction, as is evident in the Van der Waals forces between induced dipoles. The formation of optical lattices, and the many-body effects in systems such as atomic clocks, are also some of its consequences Lahaye et al. (2009).

The separatrices dividing normal from collective superradiant regions have been calculated and classified according to the Ehrenfest classification. However, there are separatrices present within the collective regimes, marking transitions between regions where one or another mode of the radiation field dominates the bulk of the ground state, which defy the Ehrenfest classification. In these cases, we have proposed two methods to detect, calculate, and classify them, one based on the second Casimir operator and another one using the surface of maximum Bures distance between neighboring states.

Appendix A Matter collective operators

Let 𝑨pq(γ){\bm{A}}_{pq}^{(\gamma)} denote the matter operator of the γ\gammath atom of nn-levels, which promotes the atom from level ωq\omega_{q} to level ωp\omega_{p}. Note that 𝑨qp(γ)=𝑨pq(γ){\bm{A}}_{qp}^{(\gamma)}={{\bm{A}}_{pq}^{(\gamma)}}^{\dagger}. For each atom γ\gamma these operators obey the unitary algebra uγ(n)u_{\gamma}(n) in nn dimensions (for nn-level atoms), i.e.,

q=1n𝑨qq(γ)=𝟏γ,\displaystyle\sum_{q=1}^{n}{\bm{A}}_{qq}^{(\gamma)}={\bm{1}}_{\gamma}\,, (47)
[𝑨pq(γ),𝑨rs(γ)]=δγγ(δqr𝑨ps(γ)δps𝑨rq(γ)),\displaystyle\left[{\bm{A}}_{pq}^{(\gamma)},{\bm{A}}_{rs}^{(\gamma^{\prime})}\right]=\delta_{\gamma\gamma^{\prime}}\left(\delta_{qr}{\bm{A}}_{ps}^{(\gamma)}-\delta_{ps}{\bm{A}}_{rq}^{(\gamma)}\right)\,, (48)

with 𝟏γ{\bm{1}}_{\gamma} the identity operator in the subspace γ\gamma. Also note that, for a single atom, we have

𝑨pq(γ)𝑨rs(γ)=δqr𝑨ps(γ).{\bm{A}}_{pq}^{(\gamma)}{\bm{A}}_{rs}^{(\gamma)}=\delta_{qr}\,{\bm{A}}_{ps}^{(\gamma)}\,. (49)

For NaN_{a} identical atoms, the collective matter operator is defined as

𝑨pq:=γ=1Na𝑨pq(γ),{\bm{A}}_{pq}:={\sum_{\gamma=1}^{N_{a}}}{\bm{A}}_{pq}^{(\gamma)}\,, (50)

and note that the sum over γ\gamma does not preserve the structure of the each subspace. By simple inspection, one may prove easily the follow relationships for the collective operators:

𝑨qp=𝑨pq,{\bm{A}}_{qp}={\bm{A}}_{pq}^{\dagger}\,, (51)
q=1n𝑨qq=γ=1Na𝟏γ:=Na 1,\sum_{q=1}^{n}{\bm{A}}_{qq}=\sum_{\gamma=1}^{N_{a}}{\bm{1}}_{\gamma}:=N_{a}\,{\bm{1}}\,, (52)
j,k=1n𝑨kj𝑨jk=Na(Na+n1),\sum_{j,k=1}^{n}{\bm{A}}_{kj}{\bm{A}}_{jk}=N_{a}\,(N_{a}+n-1)\,, (53)
[𝑨pq,𝑨rs]=δqr𝑨psδps𝑨rq.\left[{\bm{A}}_{pq},{\bm{A}}_{rs}\right]=\delta_{qr}{\bm{A}}_{ps}-\delta_{ps}{\bm{A}}_{rq}\,. (54)

Equations (52) and (53) are the first and second order Casimir operators; equation (54) shows that the operators 𝑨pq{\bm{A}}_{pq} obey a unitary algebra in nn dimensions, U(n):=γ=1Nauγ(n)U(n):=\oplus_{\gamma=1}^{N_{a}}u_{\gamma}(n). The weight operators are 𝑨pp{\bm{A}}_{pp} which give the number of particles in each atomic level ωp\omega_{p}, i.e., for an uncoupled state |ψ|\psi\rangle one has 𝑨pp|ψ=np|ψ{\bm{A}}_{pp}|\psi\rangle=n_{p}|\psi\rangle with npn_{p} the atomic population, while the operator 𝑨pq{\bm{A}}_{pq} (with pqp\neq q) promotes the transition of one atom from the level ωq\omega_{q} to the level ωp\omega_{p}; this is clear from (54) since, for the uncoupled state |ψ|\psi\rangle with atomic populations npn_{p} and nqn_{q} in the atomic levels ωp\omega_{p} and ωq\omega_{q} respectively (i.e., 𝑨pp|ψ=np|ψ{\bm{A}}_{pp}|\psi\rangle=n_{p}|\psi\rangle and 𝑨qq|ψ=nq|ψ{\bm{A}}_{qq}|\psi\rangle=n_{q}|\psi\rangle), after applying 𝑨pq|ψ=|ψ{\bm{A}}_{pq}|\psi\rangle=|\psi^{\prime}\rangle one has 𝑨pp|ψ=(np+1)|ψ{\bm{A}}_{pp}|\psi^{\prime}\rangle=(n_{p}+1)|\psi^{\prime}\rangle and 𝑨qq|ψ=(nq1)|ψ{\bm{A}}_{qq}|\psi^{\prime}\rangle=(n_{q}-1)|\psi^{\prime}\rangle, while the other atomic populations are preserved.

In similar fashion to equation (49), and using (48), one finds

𝑨pq𝑨rs=𝑨ps(𝑨rq+δrq)δrs𝑨pq+𝑶pqrs,{\bm{A}}_{pq}{\bm{A}}_{rs}={\bm{A}}_{ps}\left({\bm{A}}_{rq}+\delta_{rq}\right)-\delta_{rs}{\bm{A}}_{pq}+{\bm{O}}_{pqrs}\,, (55)

where

𝑶pqrs=γγNa(𝑨pq(γ)𝑨rs(γ)𝑨ps(γ)𝑨rq(γ)).{\bm{O}}_{pqrs}=\sum_{\gamma\neq\gamma^{\prime}}^{N_{a}}\left({\bm{A}}_{pq}^{(\gamma)}{\bm{A}}_{rs}^{(\gamma^{\prime})}-{\bm{A}}_{ps}^{(\gamma)}{\bm{A}}_{rq}^{(\gamma^{\prime})}\right)\,. (56)

It is straightforward to show the relationships 𝑶pqrs=𝑶psrq{\bm{O}}_{pqrs}=-{\bm{O}}_{psrq}, 𝑶pqrs=𝑶rspq{\bm{O}}_{pqrs}={\bm{O}}_{rspq} and 𝑶pqrq=0{\bm{O}}_{pqrq}=0. Also, for totally symmetric particles, where one may use the bosonic representation of the collective operators, one has the identity 𝑶pqrs=0{\bm{O}}_{pqrs}=0.

We define the oslash-product \oslash as the product of matter collective operators without self-interaction

𝑨pq𝑨rs:=γγNa𝑨pq(γ)𝑨rs(γ)=𝑨pq𝑨rsδqr𝑨ps.{\bm{A}}_{pq}\oslash{\bm{A}}_{rs}:=\sum_{\gamma\neq\gamma^{\prime}}^{N_{a}}{\bm{A}}_{pq}^{(\gamma)}\,{\bm{A}}_{rs}^{(\gamma^{\prime})}={\bm{A}}_{pq}\,{\bm{A}}_{rs}-\delta_{qr}{\bm{A}}_{ps}\,. (57)

Notice that 𝑨pq𝑨rs=𝑨rs𝑨pq{\bm{A}}_{pq}\oslash{\bm{A}}_{rs}={\bm{A}}_{rs}\oslash{\bm{A}}_{pq} and also 𝑶pqrs=𝑨pq𝑨rs𝑨ps𝑨rq{\bm{O}}_{pqrs}={\bm{A}}_{pq}\oslash{\bm{A}}_{rs}-{\bm{A}}_{ps}\oslash{\bm{A}}_{rq}, so that by replacing (57) into equation (55) the latter is satisfied trivially.

Appendix B Dipole-Dipole Operator

The atomic dipole-dipole interaction is written as in Eq. (10)

𝑯dd=12(Na1)jknlmngjklm𝑨jk𝑨lm.{\bm{H}}_{dd}=\frac{1}{2(N_{a}-1)}\sum_{j\neq k}^{n}\sum_{l\neq m}^{n}g_{jklm}{\bm{A}}_{jk}\oslash{\bm{A}}_{lm}\,. (58)

Taking into account the symmetries between the indices of gjklmg_{jklm}, and the possible transitions shown in table 1, we need only to replace the oslash product in (57) for the dipole-dipole operator (58) to read

𝑯dd\displaystyle{\bm{H}}_{dd} =\displaystyle= 12(Na1)jkn[gjkjk𝑨jk𝑨jk+gjkkj(𝑨jk𝑨kj𝑨jj)]+12(Na1)jkln[gjkjl𝑨jk𝑨jl+gjklj𝑨jk𝑨lj\displaystyle\frac{1}{2(N_{a}-1)}\sum_{j\neq k}^{n}\left[g_{jkjk}{\bm{A}}_{jk}{\bm{A}}_{jk}+g_{jkkj}({\bm{A}}_{jk}{\bm{A}}_{kj}-{\bm{A}}_{jj})\right]+\frac{1}{2(N_{a}-1)}\sum_{j\neq k\neq l}^{n}\left[g_{jkjl}{\bm{A}}_{jk}{\bm{A}}_{jl}+g_{jklj}{\bm{A}}_{jk}{\bm{A}}_{lj}\right. (59)
+gjklk𝑨jk𝑨lk+gjkkl(𝑨jk𝑨kl𝑨jl)]+12(Na1)jklmngjklm𝑨jk𝑨lm,\displaystyle+\left.g_{jklk}{\bm{A}}_{jk}{\bm{A}}_{lk}+g_{jkkl}({\bm{A}}_{jk}{\bm{A}}_{kl}-{\bm{A}}_{jl})\right]+\frac{1}{2(N_{a}-1)}\sum_{j\neq k\neq l\neq m}^{n}g_{jklm}{\bm{A}}_{jk}{\bm{A}}_{lm}\,,

where the first line refers to single dipole-dipole interactions, the second line to the interaction between dipoles which share an atomic level, and the third line to separate dipoles not sharing atomic levels.

We may rewrite the atomic dipole-dipole operator as

𝑯dd\displaystyle{\bm{H}}_{dd} =\displaystyle= 12!jkn𝑾jk2levels+12!jkln𝑾jlk3levels+14!jklmn𝑾jklm4levels.\displaystyle\frac{1}{2!}\sum_{j\neq k}^{n}{\bm{W}}_{jk}^{2-{\rm levels}}+\frac{1}{2!}\sum_{j\neq k\neq l}^{n}{\bm{W}}_{jlk}^{3-{\rm levels}}+\frac{1}{4!}\sum_{j\neq k\neq l\neq m}^{n}{\bm{W}}_{jklm}^{4-{\rm levels}}\,. (60)

with

𝑾jk2levels\displaystyle{\bm{W}}_{jk}^{2-{\rm levels}} =\displaystyle= 12(Na1)(gjkjk𝑨jk𝑨jk+gkjkj𝑨kj𝑨kj)+1Na1gjkkj(𝑨jk𝑨kj𝑨jj),\displaystyle\frac{1}{2(N_{a}-1)}\left(g_{jkjk}{\bm{A}}_{jk}{\bm{A}}_{jk}+g_{kjkj}{\bm{A}}_{kj}{\bm{A}}_{kj}\right)+\frac{1}{N_{a}-1}g_{jkkj}({\bm{A}}_{jk}{\bm{A}}_{kj}-{\bm{A}}_{jj})\,, (61)
𝑾jkl3levels\displaystyle{\bm{W}}_{jkl}^{3-{\rm levels}} =\displaystyle= 12(Na1)(gjklk{𝑨jk,𝑨lk}+gkjkl{𝑨kj,𝑨kl})\displaystyle\frac{1}{2(N_{a}-1)}\left(g_{jklk}\{{\bm{A}}_{jk},{\bm{A}}_{lk}\}+g_{kjkl}\{{\bm{A}}_{kj},{\bm{A}}_{kl}\}\right) (62)
+1Na1[gjkkl(𝑨jk𝑨kl𝑨jl)+gkjlk(𝑨kj𝑨lk𝑨kk)],\displaystyle+\frac{1}{N_{a}-1}\left[g_{jkkl}({\bm{A}}_{jk}{\bm{A}}_{kl}-{\bm{A}}_{jl})+g_{kjlk}({\bm{A}}_{kj}{\bm{A}}_{lk}-{\bm{A}}_{kk})\right]\,,
𝑾jklm4levels\displaystyle{\bm{W}}_{jklm}^{4-{\rm levels}} =\displaystyle= 1Na1(gjklm𝑨jk𝑨lm+gjkml𝑨jk𝑨ml+gjlkm𝑨jl𝑨km+gjlmk𝑨jl𝑨mk\displaystyle\frac{1}{N_{a}-1}\left(g_{jklm}{\bm{A}}_{jk}{\bm{A}}_{lm}+g_{jkml}{\bm{A}}_{jk}{\bm{A}}_{ml}\right.+g_{jlkm}{\bm{A}}_{jl}{\bm{A}}_{km}+g_{jlmk}{\bm{A}}_{jl}{\bm{A}}_{mk} (63)
+gjmkl𝑨jm𝑨kl+gjmlk𝑨jm𝑨lk+gkjlm𝑨kj𝑨lm+gkjml𝑨kj𝑨ml\displaystyle+g_{jmkl}{\bm{A}}_{jm}{\bm{A}}_{kl}+g_{jmlk}{\bm{A}}_{jm}{\bm{A}}_{lk}+g_{kjlm}{\bm{A}}_{kj}{\bm{A}}_{lm}+g_{kjml}{\bm{A}}_{kj}{\bm{A}}_{ml}
+gklmj𝑨kl𝑨mj+gkmlj𝑨km𝑨lj+gljmk𝑨lj𝑨mk+glkmj𝑨lk𝑨mj),\displaystyle+g_{klmj}{\bm{A}}_{kl}{\bm{A}}_{mj}+g_{kmlj}{\bm{A}}_{km}{\bm{A}}_{lj}+\left.g_{ljmk}{\bm{A}}_{lj}{\bm{A}}_{mk}+g_{lkmj}{\bm{A}}_{lk}{\bm{A}}_{mj}\right)\,,\qquad

where {𝑨jk,𝑨lm}=𝑨jk𝑨lm+𝑨lm𝑨jk\{{\bm{A}}_{jk},{\bm{A}}_{lm}\}={\bm{A}}_{jk}{\bm{A}}_{lm}+{\bm{A}}_{lm}{\bm{A}}_{jk} is the anti-commutator of 𝑨jk{\bm{A}}_{jk} and 𝑨lm{\bm{A}}_{lm}. The factor 1/p!,(p=2,4)1/p!\,,\ (p=2,4) in Eq. (60) eliminates the double summation, because 𝑾jk2levels=𝑾kj2levels{\bm{W}}_{jk}^{2-{\rm levels}}={\bm{W}}_{kj}^{2-{\rm levels}}, 𝑾jkl3levels=𝑾lkj3levels{\bm{W}}_{jkl}^{3-{\rm levels}}={\bm{W}}_{lkj}^{3-{\rm levels}} and 𝑾jklm4levels=𝑾σ(jklm)4levels{\bm{W}}_{jklm}^{4-{\rm levels}}={\bm{W}}_{\sigma(jklm)}^{4-{\rm levels}}, with σ(jklm)\sigma(jklm) a permutation of the indices (jklm)(jklm).

The contribution to the atomic dipole-dipole interaction given in (61) corresponds to transitions ωjωk\omega_{j}\rightleftharpoons\omega_{k} similar to a 22-level atom, while the contribution in (62) promotes the atomic transitions ωjωl\omega_{j}\rightleftharpoons\omega_{l} via an intermediate atomic level ωk\omega_{k}; here, the direct dipolar transition ωjωl\omega_{j}\rightleftharpoons\omega_{l} is prohibited. This contribution 𝑾jkl3levels{\bm{W}}_{jkl}^{3-{\rm levels}} appears for nn-level atoms with n3n\geq 3. The last term in Eq. (60) promotes transitions between two unconnected permitted dipolar transitions ωjωk\omega_{j}\rightleftharpoons\omega_{k} and ωlωm\omega_{l}\rightleftharpoons\omega_{m}, and is present for nn-level atoms with n4n\geq 4.

As an example, for 22-level atoms the dipole-dipole interaction reads

𝑯dd\displaystyle{\bm{H}}_{dd} =\displaystyle= 𝑾122levels,\displaystyle{\bm{W}}_{12}^{2-{\rm levels}}\,, (64)

while for 33-level atoms one finds the following for each configuration:

  • Ξ\Xi-configuration with prohibited dipolar transition ω1ω3\omega_{1}\rightleftharpoons\omega_{3} (d13=0\vec{d}_{13}=\vec{0})

    𝑯dd(Ξ)=𝑾122levels+𝑾232levels+𝑾1233levels.{\bm{H}}_{dd}^{(\Xi)}={\bm{W}}_{12}^{2-{\rm levels}}+{\bm{W}}_{23}^{2-{\rm levels}}+{\bm{W}}_{123}^{3-{\rm levels}}\,. (65)
  • Λ\Lambda-configuration with prohibited dipolar transition ω1ω2\omega_{1}\rightleftharpoons\omega_{2} (d12=0\vec{d}_{12}=\vec{0})

    𝑯dd(Λ)=𝑾132levels+𝑾232levels+𝑾1323levels,{\bm{H}}_{dd}^{(\Lambda)}={\bm{W}}_{13}^{2-{\rm levels}}+{\bm{W}}_{23}^{2-{\rm levels}}+{\bm{W}}_{132}^{3-{\rm levels}}\,, (66)
  • VV-configuration with prohibited dipolar transition ω2ω3\omega_{2}\rightleftharpoons\omega_{3} (d23=0\vec{d}_{23}=\vec{0})

    𝑯dd(V)\displaystyle{\bm{H}}_{dd}^{(V)} =\displaystyle= 𝑾122levels+𝑾132levels+𝑾2133levels.\displaystyle{\bm{W}}_{12}^{2-{\rm levels}}+{\bm{W}}_{13}^{2-{\rm levels}}+{\bm{W}}_{213}^{3-{\rm levels}}\,. (67)

Finally, we evaluate the dipole-dipole operator for two 44-level atomic configurations. In the particular case of the λ\lambda-configuration, with prohibited transitions d12=d14=d24=0\vec{d}_{12}=\vec{d}_{14}=\vec{d}_{24}=\vec{0}, the dipole-dipole operator reduces to

𝑯dd(λ)\displaystyle{\bm{H}}_{dd}^{(\lambda)} =\displaystyle= 𝑾132levels+𝑾232levels+𝑾342levels\displaystyle{\bm{W}}_{13}^{2-{\rm levels}}+{\bm{W}}_{23}^{2-{\rm levels}}+{\bm{W}}_{34}^{2-{\rm levels}} (68)
+𝑾1343levels+𝑾2343levels+𝑾1323levels;\displaystyle+{\bm{W}}_{134}^{3-{\rm levels}}+{\bm{W}}_{234}^{3-{\rm levels}}+{\bm{W}}_{132}^{3-{\rm levels}}\,;

notice that in this case we have no contribution of the form 𝑾12344levels{\bm{W}}_{1234}^{4-{\rm levels}} because all atomic levels are connected via the atomic level ω3\omega_{3}.

On the other hand, for atoms in the \largelozenge\largelozenge-configuration the prohibited dipolar transitions are d14=d23=0\vec{d}_{14}=\vec{d}_{23}=\vec{0} and, since this atomic configuration has isolated dipoles, the total dipole-dipole operator has a non-zero contribution from 𝑾12344levels{\bm{W}}_{1234}^{4-{\rm levels}} :

𝑯dd(\medlozenge)\displaystyle{\bm{H}}_{dd}^{(\medlozenge)} =\displaystyle= 𝑾122levels+𝑾132levels+𝑾242levels+𝑾342levels\displaystyle{\bm{W}}_{12}^{2-{\rm levels}}+{\bm{W}}_{13}^{2-{\rm levels}}+{\bm{W}}_{24}^{2-{\rm levels}}+{\bm{W}}_{34}^{2-{\rm levels}} (69)
+𝑾1243levels+𝑾1343levels+𝑾2133levels+𝑾2433levels\displaystyle+{\bm{W}}_{124}^{3-{\rm levels}}+{\bm{W}}_{134}^{3-{\rm levels}}+{\bm{W}}_{213}^{3-{\rm levels}}+{\bm{W}}_{243}^{3-{\rm levels}}
+𝑾12344levels.\displaystyle+{\bm{W}}_{1234}^{4-{\rm levels}}\,.

References

  • Cordero et al. (2021) S. Cordero, E. Nahmad-Achar, R. López-Peña,  and O. Castaños, Phys. Scr 96, 035104 (2021).
  • López-Peña et al. (2021) R. López-Peña, S. Cordero, E. Nahmad-Achar,  and O. Castaños, Phys. Scr. 96, 035103 (2021).
  • Weiner et al. (1999) J. Weiner, V. S. Bagnato, S. Zilio,  and P. S. Julienne, Rev. Mod. Phys. 71, 1 (1999).
  • Jones et al. (2006) K. M. Jones, E. Tiesinga, P. D. Lett,  and P. S. Julienne, Rev. Mod. Phys. 78, 483 (2006).
  • Lahaye et al. (2009) T. Lahaye, C. Menotti, L. Santos, M. Lewenstein,  and T. Pfau, Rep. Prog. Phys. 72, 126401 (2009).
  • Astrakharchik and Lozovik (2008) G. E. Astrakharchik and Y. E. Lozovik, Phys. Rev. A 77, 013404 (2008).
  • Cordero et al. (2015) S. Cordero, E. Nahmad-Achar, R. López-Peña,  and O. Castaños, Phys. Rev. A 92, 053843 (2015).
  • Gerry and Knight (2005) C. Gerry and P. Knight, Introductory Quantum Optics (Cambridge University Press, 2005).
  • Doyle et al. (2004) J. Doyle, B. Friedrich, R. V. Krems,  and F. Masnou-Seeuws, Eur. Phys. J. D 31, 149 (2004).
  • Yu et al. (2019) S. Yu, M. Titze, Y. Zhu, X. Liu,  and H. Li, Opt. Express 27, 28891 (2019).
  • Cordero et al. (2013) S. Cordero, O. Castaños, R. López-Peña,  and E. Nahmad-Achar, J. Phys. A: Math. Theor. 46, 505302 (2013).
  • Jackson (1999) J. D. Jackson, “Classical electrodynamics,”  (Wiley, New York, NY, 1999) Chap. 9, 3rd ed.
  • Cohen-Tannoudji et al. (heim) C. Cohen-Tannoudji, B. Diu,  and F. Laloë, Quantum Mechanics, Vol. II (Wiley-VCH, 2020, Weinheim).
  • Cordero et al. (2016) S. Cordero, O. Castaños, R. López-Peña,  and E. Nahmad-Achar, Phys. Rev. A 94, 013802 (2016).
  • Civitarese et al. (2010) O. Civitarese, M. Reboiro, L. Rebón,  and D. Tielas, Phys. Lett. A 374, 2117 (2010).
  • Iachello and Arima (1987) F. Iachello and A. Arima, The Interacting Boson Model (Cambridge: Cambridge University Press, 1987).
  • Klauder and Skagerstam (1985) J. R. Klauder and B. S. Skagerstam, Coherent States (World Scientific, Singapore, 1985).
  • Ali et al. (2014) S. T. Ali, J. P. Antoine,  and J. P. Gazeau, Coherent states, Wavelets, and their generalizations (Springer-Verlag, New York, 2014).
  • Gilmore (1993) R. Gilmore, Catastrophe Theory for Scientists and Engineers (Dover, 1993).
  • Bures (1969) D. Bures, Trans. Amer. Math. Soc. 135, 199 (1969).
  • Uhlmann (1976) A. Uhlmann, Rep. Math. Phys. 9, 273 (1976).