Abstract

Using an ab initio approach based on pseudopotential technique, pair potential approach, core polarization potentials, and large Gaussian basis sets, we investigate interaction of heavy alkali–krypton diatomic M–Kr (M = Rb, Cs, and Fr) van der Waals dimers. In this context, the core–core interactions for M+–Kr (M = Rb, Cs, and Fr) are calculated at coupled-cluster single and double excitation (CCSD) level and included in the total potential energy. Therefore, the potential energy curves are performed for 14 electronic states: eight of 2Σ+ symmetry, four of 2Π symmetry, and two of 2Δ symmetry. Furthermore, for each M–Kr dimer, the spin–orbit coupling has been considered for the B2Σ+, A2Π, 32Σ+, 22Π, 52Σ+, 32Π, and 12Δ states. In addition, the transition dipole moment has been determined, including the spin–orbit effect using the rotational matrix issued from the spin–orbit potential energy calculations.
I. Introduction
Interactions of alkali metal atoms with rare gas atoms are of great importance in several areas of physics and chemistry. These interactions have been investigated experimentally1−15 and theoretically16−45 for many years and for many systems. In fact, they are considered as model systems for van der Waals molecules. To understand these weak interactions, it is crucial to test both theoretical methods and basis sets. The pair potentials themselves can be used by theoreticians to model systems using molecular dynamics and by spectroscopists to aid in the understanding and analysis of molecular spectra. Due to the closed shells of atoms and/or ions that form these systems, their ground state presents very weak interactions. In contrast, considerable bonding can exist in the excited states. The interactions here concern diatomic systems, one alkali or alkaline-earth atom or ion with a single rare gas atom. One might expect the behavior of these complexes to be quite uncomplicated, with bonding attributable to purely physical interactions and van der Waals forces.
In the earlier experimental studies,1−15 the interactions between alkali and rare gas atoms have been considered as model for investigating the collisional processes such as line broadening, quenching, and electronic energy transfer. In addition, alkali metal vapor lasers that are pumped by diode lasers have been the focus of several investigations.29−33 Readle et al. have demonstrated this approach in many published papers.29−31
Hedges et al.15 investigated the emission profiles of the cesium resonance lines broadened by collisions with inert gases, based to theoretical models without knowledge of the cesium density.
Theoretically, there are several calculations16−46 of the interactions between the alkali atoms with noble gas atoms. The M(2S) + Rg pairs (M = alkali; Rg = rare gas) were investigated by Baylis16 using pseudopotential method and spin–orbit coupling. They calculated the potential energy of the ground state (X2Σ+) and the first excited states (B2Σ+ and A2Π) and extracted their spectroscopic constants. The semiempirical potential model of Baylis16 was used with a few modifications by Pascale and Vandeplanque19 to study some excited states for M–Rg species. However, the determination of the potential energy curves of alkali atoms interacting with noble gas atoms was studied using the model potential (MP)16−28 where the total potential is considered as a system with a single electron.
Ehara and Nakatsuji35 calculated the Cs–Rg system (Rg = Xe, Kr, Ar, and Ne) at the SAC-CI level. Goll et al.36 studied the M–Rg (M = Li, Na, K, Rb, and Cs; Rg = Ar, Kr, and Xe) systems using the short-range gradient-corrected density functional method combined with long-range coupled-cluster methods (CCSD, CCSD(T)). More recently, the van der Waals systems involving alkali and alkaline-earth atoms interacting with rare gas atoms have been investigated by our group37−44 using a one-electron pseudopotential approach, large Gaussian basis sets, and full configuration interaction.
The precision of produced data for the ground and excited states for similar systems, such as Cs–Ar, Rb–Ar, Cs–Xe, Rb–Xe, and Rb–He complexes,37−39,42,45 has been demonstrated by the prediction of the B ← X absorption spectra of Cs–Ar, which was found to be in excellent agreement with the experimental result of Readle et al.29 Miller et al.46 used our data for RbHe42 to investigate the temperature-dependent rubidium–helium line shape. The prediction of the pressure effect on spectrum broadening by the Anderson–Talman unified pressure broadening theory and using our data has provided good agreement with experimental observations, which proves the high precision of our calculation and the used model.
In the present paper, we report a theoretical investigation of the M–Kr (M = Rb, Cs, and Fr) van der Waals dimers. The calculation is based on a pseudopotential technique reducing the M–Kr system to one electron, the valence electron. The all-electrons calculation is difficult especially for excited states. To reduce the M–Kr (M = Rb, Cs, and Fr) systems to a one-electron problem, we have treated the M+ and Kr as two closed-shell cores interacting with the alkali metal valence electron via a semilocal pseudopotential. This permits us to reduce the M–Kr systems to one electron–core and core–core interactions to be included using the CCSD(T) accurate potential.
In the present study, we aim to calculate accurate data for the ground and excited states of heavy alkali M(2S)–Kr pairs, which are crucial to evaluate the scaling possibilities for alkali metal–rare gas laser systems. In section II, we present the theoretical method of calculations. The results for Rb–Kr, Cs–Kr, and Fr–Kr are presented in section III. Finally we conclude in section IV.
II. Method of Calculation
The electronic structures of heavy alkali atoms Rb, Cs, and Fr interacting with the Kr atom are obtained by resolving the electronic Schrödinger equation. Therefore, the number of active electrons of M–Kr is reduced to only one electron using the l-dependent semilocal pseudopotential proposed initially by Durand and Barthelat.47 Furthermore, the core polarization pseudopotentials VCPP are incorporated using the l-dependent formulation of Müller et al.48 They account for the polarization of the alkali ionic cores as well as of the Kr atom considered as a whole. For each atom (λ = Rb, Fr, or Kr), the core polarization effects are described by the following effective potential:
where λ is for either the M+ (M = Rb, Cs, and Fr) core or the krypton atom, α is the dipole polarizability, and f is the electric field created by valence electrons and cores of each center. The used polarizabilities of the alkali ions were taken as 9.245a03, 15.116a03, and 20.38a03 for Rb, Cs, and Fr, respectively,37−43 whereas that of the krypton atom was taken as 17.0a03.49
Cutoff radius has been adjusted for the three alkali atoms (Rb, Cs, and Fr) to reproduce the experimental energy level spectrum taken from ref (50). The cutoff parameters used here are 2.5213, 2.279, and 2.511 au for rubidium, 2.69, 1.58, and 2.810 au for cesium, and 3.16372, 3.045, and 3.1343 au for francium, respectively, for the lowest valence s, p, and d one-electron states.
To study the M–Kr systems, we have used the same uncontracted basis sets as in refs (37−43).
In turn, the one-electron SCF atomic calculations for the alkali atoms have been performed to test the quality of the basis sets. For the lowest valence s, p, and d one-electron states, the ionization potentials are deduced from the atomic data table50 and presented in previous publications.37−43
In addition, the electron–Kr effects have been treated through local pseudopotential. We have fitted the numerical pseudopotential of Yuan and Zhang51 using the analytical form according to
For each symmetry (l = 0 (s) orbital, l = 1 (p) orbital, and l = 2 (d) orbital), the pseudopotential parameters α, Ci, and ni are extracted. They are given in Table 1.
Table 1. Values of Parameters in the Pseudopotential of e–Kr in Semilocal Potential Form Wl(r) = exp(αr2)∑i=0nCirni (All in Atomic Units).
| l | α | Ci | ni |
|---|---|---|---|
| 0 | 0.9251 | 20.0 | –1 |
| 1 | 0.282 | 2.08 | 1 |
| 2 | 0.508 | 21.21 | 1 |
We have also used a basis set on the krypton atom. The use of a basis set on the Kr atom is essential to treat the steric distortion of the alkali valence orbitals resulting from their orthogonality to the Kr closed shells, which are represented by the pseudopotential. Since there are no active electrons on the Kr atom, the exponents were determined in order to provide correct overlap with the 3s and 2p orbitals of Kr and to extend toward the diffuse range.
In addition, to make them easy to use in a computing code based on the model potential, the core–core interaction of M+–Kr (M = Rb, Cs, and Fr) was achieved by using the numerical coupled-cluster singles, doubles, and perturbative triples (CCSD(T)) potential of Hickling et al.52 The CCSD(T) numerical potential was interpolated with the analytical form of Tang and Toennies.53 It is considered as the sum of three terms: the first one is an exponential short-range repulsion (Aeff exp(−bR)), the second is the polarization contribution (D4=−1/2αKrR–4), and the third one is the long-range attractive term (D6R–6 – D8R–8 – D10R–10). The fitted parameters for Rb+–Kr, Cs+–Kr, and Fr+–Kr are presented in Table 2.
Table 2. Interpolation Parameters of the M+–Kr (M = Rb, Cs, and Fr) Core–Core Interaction (All in Atomic Units).
| parameter | Rb+–Kr | Cs+–Kr | Fr+–Kr |
|---|---|---|---|
| Aeff | 269.41 | 316.586 | 279.684 |
| b | 1.64249 | 1.59915 | 1.56109 |
| D4 | 6.30076 | 6.46908 | 7.26873 |
| D6 | 594.698 | 610.39 | 639.961 |
| D8 | 738.447 | 2629.35 | 738.741 |
| D10 | 11857.1 | 12995.9 | 18301.8 |
In fact, the precision of the M+–Kr (M = Rb, Cs, and Fr) interactions, VM+–Kr, is crucial to obtain correct potentials for the M–Kr (M = Rb, Cs, and Fr) Rydberg states.
The spin–orbit coupling is introduced for all M–Kr (M = Rb, Cs, and Fr) complexes using the semiempirical scheme of Cohen and Schneider.55 The spin–orbit coupling matrices are isomorphic to those given by Cohen and Schneider55 for the 2p5 and 2p53s configurations of Ne+ and Ne*, respectively. The matrices for the np and nd configurations are detailed in refs (37−43). The corresponding spin–orbit constants, ξ, are determined using the atomic spectra from the NIST database.50
III. Results and Discussion
III.1. Potential Energy Curves and Spectroscopic Constants of the M–Kr (M = Rb, Cs, and Fr) Dimers
a. M–Kr Ground States (X2Σ+)
The potential energy curves (PECs) for the ground states (X2Σ+) of the M–Kr systems have been determined for a large and dense grid of intermolecular distances. Figure 1 shows the repulsive character of the PECs of the X2Σ+ ground states correlating to Rb(5s) + Kr, Cs(6s) + Kr, and Fr(7s) + Kr, respectively. We remark that the equilibrium position for their wells increases as the mass of the alkali atom increases. However, the depth of these wells decreases as the mass increases. Similarly to the M–Rg (M = Rb, Cs, and Fr; Rg = He, Ar, and Xe) cases,37−43 these repulsive states exhibit weak van der Waals minima at large internuclear distances.
Figure 1.

Potential energy curves of the X2Σ+ ground electronic states of the Rb–Kr, Cs–Kr, and Fr–Kr molecules.
The X2Σ+ states of all M–Kr van der Waals complexes are presented in Table 3 and compared with available theoretical16,35,36,45,56 and experimental57 works. They were determined for 86Rb–87Kr, 133Cs–87Kr, and 223Fr–87Kr isotopic species.
Table 3. Spectroscopic Constants of the X2Σ+ Ground States of M–Kr (M = Rb, Cs, and Fr) Systems.
| system | state | Re (a0) | De (cm–1) | ωe (cm–1) | ωexe (cm–1) | Be (cm–1) | refa |
|---|---|---|---|---|---|---|---|
| Rb–Kr | X2Σ+ (5s) | 9.91 | 80 | 9.39 | 0.69 | 0.014542 | this work |
| 9.4 | 72.3 | (16) | |||||
| 9.52 | 78.7 | 9.6 | (35) | ||||
| 9.95 | 65.2 | 8.2 | PBE/CCSD (36) | ||||
| 9.84 | 71.3 | 8.5 | PBE/CCSD(T) (36) | ||||
| 9.89 | 69.3 | 8.4 | LDA/CCSD(T)36 | ||||
| 10.47 | 48.1 | 6.6 | CCSD(T)36 | ||||
| 9.84 | 71.3 | 8.5 | (55) | ||||
| 9.99 | 73 | (57) | |||||
| 10.21 | 60.74829 | (58) | |||||
| Cs–Kr | X2Σ+ (6s) | 10.17 | 76 | 6.50 | 0.11 | 0.011353 | this work |
| 9.6 | 70.1 | (16) | |||||
| 9.63 | 91.4 | 8.9 | PBE (36) | ||||
| 10.14 | 68.1 | 7.0 | PBE/CCSD36 | ||||
| 9.98 | 75.9 | 7.3 | PBE/CCSD(T)36 | ||||
| 10.04 | 73.6 | 7.2 | LDA/CCSD(T) (36) | ||||
| 10.70 | 50.5 | 6.2 | CCSD(T) (36) | ||||
| 9.97 | 75.9 | 7.3 | (55) | ||||
| 10.28 | 74 | (57) | |||||
| 10.47 | 65.98271 | (58) | |||||
| Fr–Kr | X2Σ+ (7s) | 10.35 | 66 | 6.18 | 0.12 | 0.009237 | this work |
Abbreviations: PBE, Perdew–Burke–Ernzerhof; PBE/CCSD, Perdew–Burke–Ernzerhof/coupled-cluster singles and doubles; PBE/CCSD(T), Perdew–Burke–Ernzerhof/coupled-cluster singles, doubles, and perturbative triples; LDA/CCSD, local spin density/coupled-cluster singles and doubles; CCSD(T): coupled-cluster singles, doubles, and perturbative triples.
Our values for the X2Σ+ states for Rb–Kr and Cs–Kr can be considered in excellent agreement with the experimental values.57 The calculated potential curves of the ground states of the Rb–Kr and Cs–Kr systems have a shallow minimum of 80 and 76 cm–1 situated at 9.91a0 and 10.17a0 respectively, whereas a well of 73 cm–1 is observed at 9.99a0 for Rb–Kr and 74 cm–1 at 10.28a0 for Cs–Kr dimer by the atomic scattering experiment of Buck and Pauly.57 Similarly, shallow wells were observed theoretically by Baylis16 with the semiempirical pseudopotential approach and by Goll et al.36 using several theoretical methods, but the most reliable results were obtained using the PBE/CCSD(T) combination of density functional and coupled-cluster theory.
For Fr–Kr, the theoretical spectroscopic information for the X2Σ+ ground state is still limited. It is important to note that the spectroscopic constants are determined here for the first time. The corresponding PECs are depicted in Figure 1, and the spectroscopic constants are presented in Table 3.
b. M–Kr Excited States (2Σ+, 2Π, and 2Δ)
The available data in the literature concern the two lowest states, A2Π and B2Σ+. The A2Π and B2Σ+ states are the first excited states of the M–Kr systems correlating to Rb(5p) + Kr, Cs(5p) + Kr, and Fr(7p) + Kr. Concerning the higher (3–7)2Σ+, (2–4)2Π, and (1–2)2Δ excited states, the literature is more scarce. These states make a relatively deep well with a short equilibrium distance. The calculated excited states for all M–Kr combinations are displayed in Figure 2 and grouped by molecular term symbol. The spectroscopic constants of states with 2Σ+, 2Π, and 2Δ symmetries are listed in Tables 4–6 and compared with theoretical15,16,35 works. Comparing the results of Rb–Kr and Cs–Kr dimers to the results of refs (15) and (16), we observe a general good agreement for the well depth De. The considered spectroscopic constants of the A2Π state are De = 428/406 cm–1 and Re = 6.97/7.36a0 for Rb–Kr/Cs–Kr. In fact, we note that for all alkali–krypton dimers, the A2Π state is seen to be the more attractive state and B2Σ+ state the more repulsive state.
Figure 2.
Potential energy curves of the lowest and higher excited states of M–Kr dimers (M = Rb, Cs, and Fr).
Table 4. Spectroscopic Constants of the Excited States without and with Spin–Orbit Effect of Rb–Kr System.
| state | Re (a0) | De (cm–1) | Te (cm–1) | ωe (cm–1) | ωexe (cm–1) | Be (cm–1) | ref |
|---|---|---|---|---|---|---|---|
| A2Π (5p) | 6.97 | 428 | 12389 | 37.10 | 1.38 | 0.02926 | this work |
| A2Π1/2 (5p) | 6.92 | 392 | 12266 | 38.61 | 1.37 | 0.029749 | this work |
| 6.8 | 474 | (16) | |||||
| A2Π3/2 (5p) | 6.97 | 428 | 12468 | 37.13 | 1.38 | 0.029262 | this work |
| 6.8 | 552 | (16) | |||||
| B2Σ+ (5p) | 13.47 | 53 | 12764 | 5.47 | 1.19 | 0.031932 | this work |
| 13.7 | 17.3 | (16) | |||||
| B2Σ1/2+ (5p) | 12.84 | 52 | 12844 | 5.65 | 1.18 | 0.031568 | this work |
| 32Σ+ (4d) | 6.62 | 652 | 18783 | 58.92 | 1.11 | 0.032491 | this work |
| 15.39 | 40 | 19395 | 3.91 | ||||
| 22Π (4d) | 12.66 | 44 | 19390 | 4.77 | 0.19 | 0.008886 | this work |
| 12Δ (4d) | 6.79 | 928 | 18506 | 51.83 | 1.02 | 0.03087 | this work |
| 42Σ+ (6s) | 6.95 | 413 | 19768 | 34.57 | 2.04 | 0.029443 | this work |
| 52Σ+ (6p) | 6.61 | 740 | 231395 | 9.31 | 1.16 | 0.032613 | this work |
| 11.93 | 175 | 23704 | 6.69 | ||||
| 32Π (6p) | 6.75 | 943 | 22936 | 49.80 | 1.00 | 0.031263 | this work |
| 62Σ+ (5d) | 6.57 | 765 | 25024 | 61.30 | 1.10 | 0.032916 | this work |
| 14.16 | 125 | 25664 | 3.08 | ||||
| 42Π (5d) | 7.29 | 676 | 25113 | 35.62 | 0.52 | 0.026753 | this work |
| 22Δ (5d) | 6.63 | 1155 | 24634 | 58.12 | 0.91 | 0.032353 | this work |
| 72Σ+ (7s) | 6.72 | 821 | 25560 | 64 | 1.45 | 0.031473 | this work |
Table 6. Spectroscopic Constants of the Excited States without and with Spin–Orbit Effect of Fr–Kr System.
| state | Re (a0) | De (cm–1) | Te (cm–1) | ωe (cm–1) | ωexe (cm–1) | Be (cm–1) | ref |
|---|---|---|---|---|---|---|---|
| A2Π (7p) | 7.53 | 323 | 13140 | 34.13 | 1.25 | 0.025097 | this work |
| A2Π1/2 (7p) | 7.44 | 310 | 12706 | 37.30 | 1.13 | 0.025690 | this work |
| A2Π3/2 (7p) | 7.53 | 323 | 14380 | 34.12 | 1.28 | 0.025105 | this work |
| B2Σ+ (7p) | 14.03 | 50 | 13377 | 4.18 | 0.83 | 0.018554 | this work |
| B2Σ1/2+ (7p) | 13.46 | 46 | 14657 | 4.54 | 1.24 | 0.026249 | this work |
| 32Σ+ (6d) | 14.81 | 37 | 16407 | 3.54 | 0.07 | 0.004507 | this work |
| 22Π (6d) | 11.90 | 56 | 16423 | 5.41 | 0.15 | 0.010043 | this work |
| 12Δ (6d) | 7.48 | 602 | 15877 | 40.71 | 0.94 | 0.025447 | this work |
| 42Σ+ (8s) | 7.75 | 408 | 19346 | 22.84 | 0.67 | 0.016465 | this work |
| 52Σ1/2+ (8p) | 7.28 | 440 | 23027 | 38.14 | 0.82 | 0.018667 | this work |
| 13.43 | 148 | 23318 | 4.65 | ||||
| 32Π (8p) | 7.41 | 617 | 22885 | 40.00 | 0.92 | 0.025882 | this work |
| 62Σ+ (7d) | 7.47 | 210 | 24310 | 27.76 | 1.21 | 0.017710 | this work |
| 11.32 | 150 | 24369 | 4.18 | ||||
| 42Π (7d) | 7.73 | 562 | 23994 | 36.11 | 0.75 | 0.023790 | this work |
| 22Δ (7d) | 7.28 | 748 | 23807 | 47.01 | 1.03 | 0.026863 | this work |
| 72Σ+ (9s) | 7.41 | 621 | 25131 | 33.56 | 0.68 | 0.018000 | this work |
In the X2Σ+ state, the alkali electron is mainly in a spherical sσ orbital; however, in the B2Σ+ state, the alkali electron is mainly in a pσ orbital which overlaps the krypton atom at a larger internuclear separation than the sσ orbital. In the A2Π state, the alkali electronic wave function has pπ character with a node along the internuclear axis, and the krypton atom can approach the alkali rather closely before the repulsive interactions dominate.
For all M–Kr (M = Rb, Cs, and Fr) complexes, the np (n = 5, 6, 7) 2Σ+ and nd (n = 4, 5, 6) 2Σ+ states are due to the excitation from the 6s nonbonding molecular orbital (MO) to the feebly antibonding npσ (n = 5, 6, 7) and ndσ (n = 4, 5, 6) MOs. Therefore, the potential energy curves are repulsive, and the systems become unstable. The repulsive interaction in the np (n = 5, 6, 7) 2Σ+ states seems to be smaller than that in the nd (n = 4, 5, 6) 2Σ+ states and starts from a smaller internuclear distance. For example, Rb–Kr shows that the B2Σ+ and 52Σ+ states have a shoulder humps located around 9.0a0, while the 32Σ+ and 62Σ+ states have a characteristic hump situated at 9.89a0 and 10.5a0, respectively. The humps of the 32Σ+ and 62Σ+ states are caused by an avoided crossing between the 62Σ+ and 72Σ+ states. The same effects are also observed for the both Cs–Kr and Fr–Kr, respectively, for np (n = 7, 8) 2Σ+ and nd (n = 6, 7) 2Σ+ states. Recently, these features were also observed in our group for the alkali and alkaline-earth atoms interacting with He, Ar, and Xe atoms.37−44 They were also noticed by Ehara and Nakatsuji35 in their theoretical calculation for the Cs–Rg systems.
For the higher excited states, the potential energy curves of 2Σ+, 2Π, and 2Δ symmetries have shallow minima. The Rydberg orbitals of these states of the alkali atoms are not directed toward the rare gas atom, but the ion core of alkali atoms polarizes the electron density of the rare gas atom, which is responsible for the attractive character.
For all M–Kr complexes, the spectroscopic constants presented in Tables 4–6 show that the well depths De of the A2Π and 12Δ states are found to be superior to those of the 22Π state and that the equilibrium position of the former is larger. The potential energy of the 72Σ+ state has a similar behavior to the ionic system, which is due to the avoided crossing between the 62Σ+ and 72Σ+ states. For Rb–Kr, Cs–Kr, and Fr–Kr, the dissociation energy and equilibrium distance of 12Δ states are calculated to be 928 cm–1 at 6.79a0, 700 cm–1 at 7.31a0, and 602 cm–1 at 7.48a0, respectively, which is explained by the polarizability and the radius of the alkali atom.
The difference potentials for the transitions from the ground electronic states ns 2Σ+ [n(Rb) = 5, n(Cs) = 6, and n(Fr) = 7] to the first and higher excited states ns 2Σ+ [n(Rb) = 6, 7, n(Cs) = 7, 8, and n(Fr) = 8, 9], np 2Σ+ [n(Rb) = 5, 6, n(Cs) = 6, 7, and n(Fr) = 7, 8], and nd 2Σ+ [n(Rb) = 4, 5, n(Cs) = 5, 6, and n(Fr) = 6, 7] of alkali-rare gas dimers M–Kr are investigated for M = Rb, Cs, and Fr. They are presented in Figure 3. The difference potentials for the transitions from ns 2Σ+ [n(Rb) = 5, n(Cs) = 6, and n(Fr) = 7] to nd 2Σ+ [n(Rb) = 4, 5, n(Cs) = 5, 6, and n(Fr) = 6, 7] have extrema at 5.89a0, 6.14a0, and 6.23a0 for Rb–Kr, Cs–Kr, and Fr–Kr, respectively. These values mirror the position of the hump in the nd 2Σ+ states provoked by the avoided crossing between the nd 2Σ+ state [n(Rb) = 4, n(Cs) = 5, and n(Fr) = 6] and the 7s 2Σ+ state. On the other hand, the excitation energies of the ns 2Σ+ state and the nd 2Π state [n(Rb) = 4, n(Cs) = 5, and n(Fr) = 6)] decrease with internuclear distance, and there exists a flat region at small distances.
Figure 3.
Excitation energies from the ground state to the excited states of the M–Kr dimers (M = Rb, Cs, and Fr) as functions of the internuclear distance.
III.2. M–Kr (2Σ+, 2Π, and 2Δ) Excited States Introducing the Spin–Orbit Coupling
We have introduced the spin–orbit coupling for the first and higher excited states of the M–Kr systems correlated to the nΠ states for Rb, Cs, and Fr atoms, where n(Rb) = 5, 6, n(Cs) = 6, 7, and n(Fr) = 7, 8, by using the semiempirical scheme of Cohen and Schneider55 for the 2p53s and 2p5 configurations of Ne* and Ne+, respectively. The spin–orbit coupling constants ξ used in this calculation are as follows: ξ5p/6p(Rb) = 158.396/77.51 cm–1, ξ6p/7p(Cs) = 369.36/181.046 cm–1, and ξ7p/8p(Fr) = 1124.392/545.346 cm–1.
Figure 4 left shows the calculated PECs of M–Kr states dissociating into np + Kr and n2P1/2,3/2 + Kr, where n = 5, 6, and 7 for Rb, Cs, and Fr, respectively. The n2P1/2 and n2P3/2 states are obtained by diagonalization of the spin–orbit coupling matrix. The calculated spectroscopic constants for all M–Kr exciplexes are summarized in Tables 4–6.
Figure 4.
M–Kr (M = Rb, Cs, and Fr) potential energy curves including spin–orbit coupling dissociating into n2P1/2,3/2 + Kr for (left) n = 5–7 and (right) n = 6–8. Red lines are for states including spin–orbit coupling; black lines are for states without spin–orbit coupling.
These new states that include spin–orbit effect demonstrate considerable quantitative variation and are qualitatively similar. The excitations (ΔTe) for n2P1/2 and n2P3/2 states are 202, 529, and 1574 cm–1 for Rb–Kr, Cs–Kr, and Fr–Kr, respectively. The larger splitting value for Fr–Kr states could be explained by the large spin–orbit constant, ξ7p(Fr) = 1124.392 cm–1, in comparison with ξ6p(Cs) = 369.36 cm–1 and ξ5p(Rb) = 158.396 cm–1.
For the Rb–Kr and Cs–Kr systems, including the spin–orbit effects, we show that our potential energy curves of the B2Σ1/2+, A2Π1/2, and A2Π3/2 states corresponding to the n2P1/2,3/2 + Kr (n = 5, 6) asymptote and those obtained experimentally by Hedges et al.15 are in good agreement. The same accord is observed also with theoretical results of Baylis16 and Ehara and Nakatsuji.35 For Cs–Kr, the dissociation energy and the equilibrium position are found to be De = 375/400 cm–1 and Re = 7.29/7.37a0 for the A2Π1/2 and A2Π3/2 states, respectively, to be compared with the experimental values15 of De = 300/350 cm–1, respectively.
By introducing the spin–orbit coupling, the first A2Π1/2 and A2Π3/2 excited states were studied by Baylis,16 who found two different values for De (112 and 439 cm–1 for the A2Π1/2 and A2Π3/2 states, respectively) and the same Re (7.18a0). Ehara and Nakatsuji35 found shallow potential wells of De = 163 and 283 cm–1, respectively. Our results for the A2Π3/2 state and those obtained by Baylis16 are in good agreement, while for the A2Π1/2 state, the well depth found by Baylis16 (112 cm–1) seems to be underestimated compared to the present work as well as the experimental work.15
Using the atomic spin–orbit constants ξ6p(Rb) = 77.51 cm–1, ξ7p(Cs) = 181.046 cm–1, and ξ8p (Fr) = 545.346 cm–1, the spin–orbit effect is also investigated for states correlating to Rb(6p), Cs(7p), and Fr(8p). Figure 4 right depicts the energies of the five states produced (52Σ+, 32Π, 52Σ1/2+, 32Π1/2, and 32Π3/2) without and with spin–orbit perturbation. We show that the PECs including the spin–orbit coupling are considerably similar to those of B2Σ+ and A2Π states in all of the M–Kr complexes from Rb to Fr.
We noticed that the molecular splitting energies at equilibrium position for the 32Π1/2 and 32Π3/2 states in M–Kr (M = Rb, Cs, and Fr) are smaller than the splittings found for the A2Π1/2 and A2Π3/2 states, which is due to the small spin–orbit constant for the np atomic limit, where n = 6, 7, and 8, for Rb, Cs, and Fr, respectively. The calculated data, before inclusion of the spin–orbit coupling, are presented in Tables 4–6. The present calculations illustrate a slight change for the equilibrium distance and the vibrational constant by introducing spin–orbit perturbation, but a larger change in vertical transition energy Te.
The literature about the higher excited states (52Σ1/2+, 32Π1/2, and 32Π3/2) with spin–orbit coupling is limited. To our best knowledge, no experimental or theoretical results are available for the M–Kr molecular systems.
For the 32Σ+, 22Π, and 12Δ excited states correlating to the Cs(5d) + Kr limit, the spin–orbit effect is introduced by using the following value: ξ5d(Cs) = 39.03356 cm–1. For Rb–Kr, it should be noticed that the spin–orbit coupling for the states dissociating into the Rb(4d) + Kr dissociation limit is not considered due to the small spin–orbit constant.
The new Cs–Kr molecular states, dissociating into 52D1/2 + Kr (32Σ1/2+ and 22Π1/2), n2D3/2 + Kr (22Π3/2 and 12Δ3/2), and n2D5/2 + Kr (12Δ5/2), are presented in Figure 5. The corresponding spectroscopic constants are presented here for the first time in Table 5, since there is no comparison of these values with other results.
Figure 5.
Cs–Kr potential energy curves including spin–orbit coupling dissociating into 52D1/2,3/2,5/2 + Kr. Solid lines show states including spin–orbit coupling and dashed lines states without spin–orbit coupling.
Table 5. Spectroscopic Constants of the Excited States without and with Spin–Orbit Effect of Cs–Kr System.
| state | Re (a0) | De (cm–1) | Te (cm–1) | ωe (cm–1) | ωexe (cm–1) | Be (cm–1) | ref |
|---|---|---|---|---|---|---|---|
| A2Π (6p) | 7.36 | 406 | 11219 | 32.92 | 1.06 | 0.021609 | this work |
| A2Π1/2 (6p) | 7.31 | 375 | 10881 | 34.80 | 1.07 | 0.021894 | this work |
| 300 | (15) | ||||||
| 7.46 | 163 | 35.1 | (35) | ||||
| 7.20 | 112 | (16) | |||||
| A2Π3/2 (6p) | 7.38 | 406 | 11410 | 32.27 | 0.92 | 0.021494 | this work |
| 350 | (15) | ||||||
| 7.52 | 283 | 38.3 | (35) | ||||
| 7.20 | 439 | (16) | |||||
| B2Σ+ (6p) | 13.45 | 56 | 11569 | 4.74 | 0.91 | 0.022881 | this work |
| B2Σ1/2+ (6p) | 12.84 | 52 | 12844 | 5.65 | 1.18 | 0.031568 | this work |
| 16.2 | 17.1 | (16) | |||||
| 32Σ+ (5d) | 14.60 | 35 | 14592 | 3.91 | 0.11 | 0.005495 | this work |
| 32Σ1/2+ (5d) | 13.65 | 41 | 14633 | 4.88 | 0.23 | 0.006282 | this work |
| 22Π (5d) | 11.16 | 74 | 6.84 | 0.37 | 0.009402 | this work | |
| 22Π1/2 (5d) | 11.08 | 70 | 14534 | 7.21 | 0.83 | 0.009516 | this work |
| 300 | (15) | ||||||
| 7.46 | 163 | 35.1 | (35) | ||||
| 22Π3/2 (5d) | 11.14 | 74 | 14573 | 9.02 | 0.77 | 0.009437 | this work |
| 350 | (15) | ||||||
| 12Δ (5d) | 7.31 | 700 | 13927 | 39.02 | 0.79 | 0.021895 | this work |
| 350 | (15) | ||||||
| 7.37 | 441 | 46.2 | (35) | ||||
| 12Δ3/2 (5d) | 7.32 | 657 | 13887 | 39.02 | 0.81 | 0.021890 | this work |
| 12Δ5/2 (5d) | 7.31 | 700 | 13966 | 39.01 | 0.81 | 0.021902 | this work |
| 42Σ+ (7s) | 7.47 | 461 | 18154 | 27.98 | 0.91 | 0.020985 | this work |
| 8.22 | 322 | 22.3 | (35) | ||||
| 52Σ+ (7p) | 7.11 | 599 | 21433 | 43.71 | 0.92 | 0.023160 | this work |
| 52Σ1/2+ (7p) | 7.15 | 624 | 21499 | 42.44 | 0.83 | 0.022927 | this work |
| 32Π (7p) | 7.24 | 741 | 21292 | 38.79 | 0.69 | 0.022331 | this work |
| 32Π1/2 (7p) | 7.24 | 699 | 21153 | 39.02 | 0.73 | 0.022375 | this work |
| 32Π3/2 (7p) | 7.26 | 734 | 21390 | 38.63 | 0.69 | 0.022231 | this work |
| 62Σ+ (6d) | 7.41 | 220 | 22662 | 26.23 | 2.12 | 0.021341 | this work |
| 42Π (6d) | 7.56 | 670 | 22212 | 37.02 | 0.68 | 0.020485 | this work |
| 22Δ (6d) | 7.11 | 859 | 22023 | 44.78 | 0.81 | 0.023164 | this work |
| 72Σ+ (8s) | 7.27 | 658 | 23775 | 36.34 | 0.84 | 0.022181 | this work |
In addition, we note that the spin–orbit effect for the Cs n2P1/2,3/2 (n = 6, 7) states is much larger than that observed for the Cs(52D1/2,3/2,5/2) ones, which is due to the large spin–orbit constant for the np atomic limit.
III.3. Qualitative Prediction of the Broadening Spectrum of Alkali (Rb, Cs, and Fr) D2 Line
Using our potential energy curves for the lower (X2Σ+) and upper (B2Σ+ and A2Π) electronic states, the transition wavelengths were generated for all M–Kr systems using the following expression:43
where V′(R) and V''(R) are the potential energy curves for the upper and lower electronic states respectively. In fact, the interaction potentials of the X2Σ+ state and the B2Σ+ and A2Π states and their differences are relevant to the broadening of the D2 and D1 lines45 of alkali atoms. The latter quantity provides information about contributions from dimers or collision pairs. Figure 6 shows the derived transition wavelengths λ(R) for all M–Kr (M = Rb, Cs, and Fr) dimers calculated from the difference potentials B2Σ+ – X2Σ+ and A2Π – X2Σ+ (Figure 6). As can be seen from Figure 6, two different internuclear separations could contribute to the same transition and its wavelength. From this figure we predict that the D2 broadening spectra are confined to the regions 770–865, 849–925, and 715–825 nm for Rb–Kr, Cs–Kr, and Fr–Kr complexes, respectively. Heaven and Stolyarov58 determined the maximum for the blue wing of the D2 line from Cs in Kr to be 835 nm using the QS approximation,45 which is in a general good agreement with the observed value of 841 nm. It is expected that our maximum will be at 849 nm, which is overestimated compared to the theoretical and experimental values.45 The same thing is observed for the Rb–Kr pair, as we found a value of 770 nm to be compared with the calculated and observed values58 of 753 and 759 nm.45 Once again, our wavelength is overestimated compared to the calculated and observed ones.45 This is due to the original used potential interactions for the X2Σ+ and B2Σ+ states. It seems that the potential used by Heaven and Stolyarov58 for the B2Σ+ state including spin–orbit coupling is more repulsive than ours, which makes their wavelength shorter. It is clear that the difference between our work and that of Heaven and Stolyarov58 is primarily related to the difference in the potential energy interaction calculation approaches. We used a simple model that reduced the M–Kr pair to only one valence electron interacting with the M+–Kr ionic system. However, Heaven and Stolyarov58 used more sophisticated ab initio approaches such as multireference configuration interaction (MRCI) and multireference averaged quadratic coupled cluster (MRAQCC) to produce the potential interactions used in the spectrum broadening simulations.
Figure 6.
Transition wavelengths calculated from the difference potentials B2Σ+ – X2Σ+ and A2Π – X2Σ+ for the M–Kr dimers (M = Rb, Cs, and Fr).
III.4. Vibrational Analysis and Transition Dipole Moments of M–Kr (M = Rb, Cs, and Fr) Dimers Introducing the Spin–Orbit Coupling
By solving the nuclear radial Schrödinger equation, the vibrational level spacing of alkali–rare gas diatomic molecules M–Kr are investigated for the molecular states correlated asymptotically to the degenerate 2P (np) states, where n(Rb) =5, 6, n(Cs) = 6, 7, and n(Fr) = 7, 8, introducing the spin–orbit effect.
They were determined for 86Rb–87Kr, 133Cs–87Kr, and 223Fr–87Kr isotopic species using the Numerov algorithm with 30 000 points ranging from Rmin = 3.0a0 to Rmax = 200a0 and reduced masses μRb–Kr = 42.3123 a.u, μCs–Kr = 51.3939 a.u, and μFr–Kr = 60.9096 a.u.
The vibrational level spacings (Ev – Ev–1) in terms of the vibrational number v of the M–Kr dimers are depicted in Figures 7 and 8. As can be seen, our calculation for the 32Π and 52Σ+ higher excited states correlating asymptotically with 2P (np), where n(Rb) = 6, n(Cs) = 7, and n(Fr) = 8, indicates more vibrational levels than for the A2Π and B2Σ+ states correlating with 2P (np), where n(Rb) = 5, n(Cs) = 6, and n(Fr) = 7, which present, respectively, 40, 56, and 58 levels for Rb–Kr, Cs–Kr, and Fr–Kr dimers. This dissimilarity is obviously related to the difference in the well depths.
Figure 7.

Vibrational level spacings of M–Kr (M = Rb, Cs, and Fr) diatomic molecules between the molecular states correlating asymptotically to the degenerate 2P (np) states, where n(Rb) = 5, n(Cs) = 6, and n(Fr) = 7.
Figure 8.

Vibrational level spacing of M–Kr (M = Rb, Cs, and Fr) diatomic molecules between the molecular states correlating asymptotically to the degenerate 2P (np) states, where n(Rb) = 6, n(Cs) = 7, and n(Fr) = 8.
We have also determined vibrational levels for the B2Σ+ and A2Π states and 52Σ+ and 32Π states by introducing the spin–orbit coupling. For all systems where the spin–orbit coupling plays a more significant role, the variation of (Ev – Ev–1) with v is split into two curves. We are now interested in Rb interacting with the 87Kr isotope. We found 32 and 40 vibrational levels for Ω = 1/2 and Ω = 3/2, respectively. Moreover, when approaching the dissociation limit, the energy level spacing vanishes and the common behavior is a decrease as v rises. The same behavior has been observed also for all other dimers. The vibrational levels for heavy alkali–krypton diatomic molecules are more important than for the light alkali–rare gas molecules.59 To our knowledge, the vibrational levels for heavy alkali–krypton diatomic molecules have been calculated here for the first time.
The calculated transition dipole moment is an essential factor in determining the absorption spectrum. To investigate the nS → nP [n(Rb) = 5, n(Cs) = 6, n(Fr) = 7] transition dipole moments of the alkali atoms by the interaction with Kr atom, the spin–orbit coupling is included by using the energy matrix discussed previously issued from the diagonalization of the rotational matrix. The spin–orbit constants are estimated to be ξ5p(Rb) = 158.396 cm–1, ξ6p(Cs) = 369.36 cm–1, and ξ7p(Fr) = 1124.392 cm–1 from the atomic spectral data.50
The A ← X and B ← X transition dipole moments with and without spin–orbit effect are shown in Figures 9–11 as functions of the internuclear distance. These figures show that the X2Σ+ → B2Σ+ transition moments for Rb–Kr, Cs–Kr, and Fr–Kr have a maximum around 3.0a0. The X2Σ+ → A2Π transition dipole is split into two curves when the spin–orbit coupling is included. They are labeled as X2Σ+ → A2Π1/2 and X2Σ+ → A2Π3/2. The difference between the X2Σ+ → A2Π1/2 and X2Σ+ → A2Π3/2 transition dipole moments is significant at intermediate distances. The transition moments should converge to the value of the pure atomic 2P ← 2S transition for large internuclear separations.
Figure 9.

Transition dipole moments for the X2Σ+ → A2Π, X2Σ+ → B2Σ+, X2Σ+ → A2Π1/2, X2Σ+ → A2Π3/2, and X2Σ+ → BX2Σ1/2+ transitions as functions of the internuclear distance for Rb–Kr dimer.
Figure 11.

Transition dipole moments for the X2Σ+ → A2Π, X2Σ+ → B2Σ+, X2Σ+ → A2Π1/2, X2Σ+ → A2Π3/2, and X2Σ+ → B2Σ1/2+ transitions as functions of the internuclear distance for Fr–Kr dimer.
Figure 10.

Transition dipole moments for the X2Σ+ → A2Π, X2Σ+ → B2Σ+, X2Σ+ → A2Π1/2, X2Σ+ → A2Π3/2, and X2Σ+ → B2Σ1/2+ transitions as functions of the internuclear distance for Cs–Kr dimer.
Reliable and accurate transition dipole moments at long-range internuclear distances in addition to precise potential energies are fundamental for prediction and design of the scaling characteristics of lasers driven by pumping of alkali–rare gas collision pairs.60 In fact, highly accurate potential interaction data are needed to predict the absorption spectra and broadening of alkali atomic lines. More precisely, the accurate transition dipole functions and interaction energies are used to simulate the absorption spectrum and broadening in the D2 line for Rb–Kr, Cs–Kr, and Fr–Kr dimers.
The long-range behavior of the permitted transition dipole moment was extensively explored in the papers of Chu and Dalgarno61 and Pazyuk et al.62 The latter demonstrated that the asymptotic behavior of the Rb2 and Cs262 transition dipole functions at R = ∞ tends to the limiting atomic value Bn/Rn with n = 3, 4, where the coefficients Bn can be calculated using the atomic wave functions and energies. A quantitative fitting of some transition dipole moments showed this behavior for Cs–Kr and Rb–Kr dimers. A detailed study of the transition dipole functions at long-range internuclear distances is in progress for a better understanding of the long-range behavior. This will permit the probing of their quality. This quality is related to that of the electronic wave functions, which were also used to evaluate the molecular energies.
IV. Conclusion
We have performed an accurate theoretical study of alkali metal (Rb, Cs, and Fr) atoms interacting with a rare gas (Kr) atom. The M+ (M = Rb, Cs, and Fr) cores and the electron–Kr interactions were replaced by a semilocal pseudopotential, and the M–Kr systems were reduced to one-electron SCF calculations. The potential energy curves and their corresponding spectroscopic constants have been determined. The spectroscopic parameters of the X2Σ+ ground states show a good agreement compared to the available theoretical15,16,35,36 and experimental57 works.
Moreover, the semiempirical scheme of Cohen and Schneider55 was included to study the spin–orbit effect. One can notice that the largest energy splitting is observed for the first excited states B2Σ+ and A2Π. Our spectroscopic constants and the earlier studies for the B2Σ1/2+, A2Π1/2, and A2Π3/2 excited states are compared with experimental57 and theoretical15,16,35,36 results. Good agreement was observed for the equilibrium distances and dissociation energies. Because of the spin–orbit effect, a splitting was also observed in the case of the vibrational level spacing and the transition dipole moment.
In addition, a qualitative prediction of the D2 broadening spectra was performed in this study. We found that broadening D2 spectra are confined to the regions 770–865, 849–925, and 715–825 nm for Rb–Kr, Cs–Kr, and Fr–Kr, respectively. Expected maxima for the blue-wing D2 lines are predicted to appear at 770, 849, and 715 for Rb–Kr, Cs–Kr, and Fr–Kr, respectively. Heaven and Stolyarov58 determined the maxima for the blue wing of the D2 line from Rb and Cs in Kr to be 753 and 835 nm using the QS approximation, in a good agreement with the observed values of 759 and 841 nm. Our predictions are 770 and 849 nm, which are overestimated compared to the theoretical and experimental values.48−58 This could be associated to the difference in the potential interactions for the X2Σ+ and B2Σ1/2+ states, especially at short and intermediate repulsive parts in both states’ potentials, making the difference in potential energy smaller and therefore giving longer wavelengths. It is clear that the difference between our work and that of Heaven and Stolyarov58 is primarily related to the difference in the potential energy interaction calculation approaches. The potential used by Heaven and Stolyarov58 for the B2Σ1/2 state including spin–orbit coupling is more repulsive than ours. Considering the simplicity of our used model, reducing the M–Kr pair to only one valence electron interacting with the M+–Kr ionic system, and the more sophisticated ab initio approaches used by Heaven and Stolyarov,58 our results are in general good agreement with theirs.
In addition, to study the structure and dynamics of large M+–Krn clusters, the one-electron pseudopotential approach can be used. Furthermore, to investigate the expansion effect of the alkali atom by collision with the Kr atom and also to evaluate transport coefficients for alkali atoms moving through a bath of Kr atoms, the accurate results of the M–Kr systems will be used.
The authors declare no competing financial interest.
References
- Zhdanov B. V.; Ehrenreich T.; Knize R. Highly efficient optically pumped cesium vapor laser. J. Opt. Commun. 2006, 260, 696. 10.1016/j.optcom.2005.11.042. [DOI] [Google Scholar]
- Krupke W.; Beach R.; Kanz V.; Payne S. Resonance transition 795-nm rubidium laser. Opt. Lett. 2003, 28, 2336. 10.1364/OL.28.002336. [DOI] [PubMed] [Google Scholar]
- Beach R. J.; Krupke W. F.; Kanz V. K.; Payne S. A.; Dubinskii M. A.; Merkle L. D. End-pumped continuous-wave alkali vapor lasers: experiment, model, and power scaling. J. Opt. Soc. Am. B 2004, 21, 2151. 10.1364/JOSAB.21.002151. [DOI] [Google Scholar]
- Page R. H.; Beach R. J.; Kanz V. K.; Krupke W. F. Multimode-diode-pumped gas (alkali-vapor) laser. Opt. Lett. 2006, 31, 353. 10.1364/OL.31.000353. [DOI] [PubMed] [Google Scholar]
- Lyyra M.; Bunker P.R. The potential energy curve of the B 1Πu state of Na2. Chem. Phys. Lett. 1979, 61, 67–68. 10.1016/0009-2614(79)85086-1. [DOI] [Google Scholar]
- Tellinghuisen J.; Ragone A.; Kim M. S.; Auerbach D. J.; Smalley R. E.; Wharton L.; Levy D. H. The dispersed Fluorescence Spectrum of NaAr: Ground and Excited State Potential Curves. J. Chem. Phys. 1979, 71, 1283. 10.1063/1.438428. [DOI] [Google Scholar]
- Goble J. H.; Winn J. S. Analytic potential functions for weakly bound molecules: The X and A states of NaAr and the A state of NaNe. J. Chem. Phys. 1979, 70, 2051. 10.1063/1.437762. [DOI] [Google Scholar]
- Brühl R.; Zimmermann D. High-resolution Laser Spectroscopy of the A ← X Transition of LiAr. Chem. Phys. Lett. 1995, 233, 455–460. 10.1016/0009-2614(94)01476-C. [DOI] [Google Scholar]
- Brühl R.; Zimmermann D. High-resolution Laser Spectroscopy of LiAr: Spectroscopic Parameters and Interaction Potentials of the A2Π and the B2Σ States. J. Chem. Phys. 2001, 114, 3035. 10.1063/1.1340565. [DOI] [Google Scholar]
- Aepfelbach G.; Nunnemann A.; Zimmermann D. Laser Spectroscopic Investigation of the X 2Σ1/2 State of the van der Waals Molecule NaAr. Chem. Phys. Lett. 1983, 96, 311–315. 10.1016/0009-2614(83)80679-4. [DOI] [Google Scholar]
- Bokelmann F.; Zimmermann D. Determination of the K–Ar Interaction Potential in the XΣ and AΠ State from Laser Spectroscopic Data. J. Chem. Phys. 1996, 104, 923. 10.1063/1.470816. [DOI] [Google Scholar]
- Michalak R.; Zimmermann D. Laser-Spectroscopic Investigation of Higher Excited Electronic States of the KAr Molecule. J. Mol. Spectrosc. 1999, 193, 260–272. 10.1006/jmsp.1998.7727. [DOI] [PubMed] [Google Scholar]
- Schwarzhans D.; Zimmermann D. High Resolution Laser Spectroscopy of NaAr: Improved Interaction Potential for the X2Σ+ Ground State. Eur. Phys. J. D 2003, 22, 193–198. 10.1140/epjd/e2002-00242-8. [DOI] [Google Scholar]
- Braune M.; Valipour H.; Zimmermann D. High-resolution Laser Spectroscopy of the Potassium–Argon Molecule: The B2Σ+ State. J. Mol. Spectrosc. 2006, 235, 84–92. 10.1016/j.jms.2005.10.006. [DOI] [Google Scholar]
- Hedges R. E. M.; Drummond D. L.; Gallagher A. Extreme-Wing Line Broadening and Cs-Inert-Gas Potentials. Phys. Rev. A 1972, 6, 1519. 10.1103/PhysRevA.6.1519. [DOI] [Google Scholar]
- Baylis W. Semi empirical, Pseudopotential Calculation of Alkali–Noble-Gas Interatomic Potentials. J. Chem. Phys. 1969, 51, 2665. 10.1063/1.1672393. [DOI] [Google Scholar]
- Dubourg I.; Ferray M.; Visticot J. P.; Sayer B. Experimental investigation of the Rb (6S or 4D)-rare gas interaction: determination of interaction potentials and oscillator strengths. J. Phys. B 1986, 19, 1165. 10.1088/0022-3700/19/8/012. [DOI] [Google Scholar]
- Drummond D. L.; Gallagher A. Potentials and continuum spectra of Rb-noble gas molecules. J. Chem. Phys. 1974, 60, 3426. 10.1063/1.1681555. [DOI] [Google Scholar]
- Pascale J.; Vandeplanque J. Excited Molecular Terms of The Alkali-Rare Gas Atom Pairs. J. Chem. Phys. 1974, 60, 2278. 10.1063/1.1681360. [DOI] [Google Scholar]
- Saxon R. P.; Olson R. E.; Liu B. Ab initio calculations for the X2Σ+, A2Π, and B2Σ+ states of NaAr: Emission spectra and cross sections for fine-structure transitions in Na–Ar collisions. J. Chem. Phys. 1977, 67, 2692. 10.1063/1.435183. [DOI] [Google Scholar]
- Czuchaj E.; Sienkiewicz J. Adiabatic potentials of the alkali-rare gas atom pairs. Z. Naturforsch., A 1979, 34, 694. 10.1515/zna-1979-0604. [DOI] [Google Scholar]
- Laskowski B. C.; Langhoff S. R.; Stallcop J. R. Theoretical calculation of low-lying states of NaAr and NaXe. J. Chem. Phys. 1981, 75, 815. 10.1063/1.442125. [DOI] [Google Scholar]
- Sadlej J.; Edwards W. D. Ab initio study of the ground and first excited state of LiAr. Int. J. Quantum Chem. 1995, 53, 607–615. 10.1002/qua.560530604. [DOI] [Google Scholar]
- Gu J.-p.; Hirsch G. R.; Buenker J.; Petsalakis I.; Theodorakopoulos G.; Huang M.-b. Electronic states and radiative transitions in LiAr. Chem. Phys. Lett. 1994, 230, 473–479. 10.1016/0009-2614(94)01198-2. [DOI] [Google Scholar]
- Park S. J.; Lee Y. S.; Jeung G.-H. Double-well of the C2Σ+ state of LiAr. Chem. Phys. Lett. 1997, 277, 208–214. 10.1016/S0009-2614(97)00887-7. [DOI] [Google Scholar]
- Ahokas J.; Kiljunen T.; Eloranta J.; Kunttu H. Theoretical analysis of alkali metal trapping sites in rare gas matrices. J. Chem. Phys. 2000, 112, 2420. 10.1063/1.480825. [DOI] [Google Scholar]
- Rhouma M. B. E. H.; Berriche H.; Lakhdar Z. B.; Spiegelman F. One-Electron Pseudopotential Calculations of Excited States of LiAr, NaAr, and KAr. J. Chem. Phys. 2002, 116, 1839. 10.1063/1.1429247. [DOI] [Google Scholar]
- El Hadj Rhouma M. B.; Ben Lakhdar Z.; Berriche H.; Spiegelman F. Rydberg States of Small NaArn* Clusters. J. Chem. Phys. 2006, 125, 084315. 10.1063/1.2229210. [DOI] [PubMed] [Google Scholar]
- Readle J. D.; Eden J. G.; Verdeyen J. T.; Carroll D. L. Four level, atomic Cs laser at 852.1 nm with a quantum efficiency above 98%: Observation of three-body photoassociation. Appl. Phys. Lett. 2010, 97, 021104. 10.1063/1.3452338. [DOI] [Google Scholar]
- Readle J. D.; Verdeyen J. T.; Eden J. G.; Davis S. J.; Galbally-Kinney K. L.; Rawlins W. T.; Kessler W. J. Cs 894.3 nm laser pumped by photoassociation of Cs–Kr pairs: excitation of the Cs D2 blue and red satellites. Opt. Lett. 2009, 34, 3638–3640. 10.1364/OL.34.003638. [DOI] [PubMed] [Google Scholar]
- Readle J. D.; Wagner C. J.; Verdeyen J. T.; Spinka T. M.; Carroll D. L.; Eden J. G. Pumping of atomic alkali lasers by photo excitation of a resonance line blue satellite and alkali-rare gas excimer dissociation. Appl. Phys. Lett. 2009, 94, 251112. 10.1063/1.3151854. [DOI] [Google Scholar]
- Blank L.; Weeks D. E.; Kedziora G. S. M+Ng potential energy curves including spin-orbit coupling for M = K, Rb, Cs and Ng = He, Ne, Ar. J. Chem. Phys. 2012, 136, 124315. 10.1063/1.3696377. [DOI] [PubMed] [Google Scholar]
- Zbiri M.; Daul C. Investigating the exciplexes M*He M = {Li, Na, K, Rb, Cs,Fr}: Density functional approach. J. Chem. Phys. 2004, 121, 11625. 10.1063/1.1810133. [DOI] [PubMed] [Google Scholar]
- Galbis E.; Douady J.; Jacquet E.; Giglio E.; Gervais B. Potential energy curves and spin-orbit coupling of light alkali-heavy rare gas molecules. J. Chem. Phys. 2013, 138, 014314. 10.1063/1.4773019. [DOI] [PubMed] [Google Scholar]
- Ehara M.; Nakatsuji H. Collision induced absorption spectra and line broadening of CsRg system (Rg = Xe, Kr, Ar, Ne) studied by the symmetry adapted cluster-configuration interaction (SAC-CI) method. J. Chem. Phys. 1995, 102, 6822. 10.1063/1.469118. [DOI] [Google Scholar]
- Goll E.; Werner H.-J.; Stoll H.; Leininger T.; Gori-Giorgi P.; Savin A. A Short-Range Gradient-Corrected Spin Density Functional in Combination with Long-Range Coupled-Cluster Methods: Application to Alkali-Metal Rare-Gas Dimers. Chem. Phys. 2006, 329, 276–282. 10.1016/j.chemphys.2006.05.020. [DOI] [Google Scholar]
- Dhiflaoui J.; Berriche H. One-Electron Pseudopotential Investigation Of CsAr Van der Waals System Including The Spin–Orbit Interaction. J. Phys. Chem. A 2010, 114, 7139–7145. 10.1021/jp101957s. [DOI] [PubMed] [Google Scholar]
- Dhiflaoui J.; Berriche H.; Herbane M.; Al Sehimi M. H.; Heaven M. Electronic Structure and Spectra of the RbAr van der Waals System Including Spin–Orbit Interaction. J. Phys. Chem. A 2012, 116, 10589–10596. 10.1021/jp301486t. [DOI] [PubMed] [Google Scholar]
- Dhiflaoui J.; Berriche H. One-electron Pseudopotential Investigation of the RbAr and FrAr van der Waals Systems. AIP Conf. Proc. 2012, 1504, 664–669. 10.1063/1.4771783. [DOI] [Google Scholar]
- Berriche H. Electronic, Spectra, and Spin Orbit Interaction for FrAr van der Waals System. Int. J. Quantum Chem. 2013, 113, 1349–1357. 10.1002/qua.24302. [DOI] [Google Scholar]
- Bejaoui M.; Dhiflaoui J.; Mabrouk N.; El Ouelhazi R.; Berriche H. Theoretical Investigation of the Electronic Structure and Spectra of Mg2+He and Mg+He. J. Phys. Chem. A 2016, 120, 747–753. 10.1021/acs.jpca.5b10089. [DOI] [PubMed] [Google Scholar]
- Dhiflaoui J.; Bejaoui M.; Berriche H. Electronic structure and spectra of the RbHe van der Waals system including spin orbit interaction. Eur. Phys. J. D 2017, 71, 70576. 10.1140/epjd/e2017-70576-1. [DOI] [PubMed] [Google Scholar]
- Dhiflaoui J.; Berriche H.; Heaven M. Theoretical investigation of RbXe and CsXe excimers including the spin–orbit interaction. J. Phys. B 2016, 49, 205101. 10.1088/0953-4075/49/20/205101. [DOI] [Google Scholar]
- Zrafi W.; Bejaoui M.; Dhiflaoui J.; Farjallah M.; Berriche H. Non-relativistic and relativistic investigation of the low lying electronic states of Sr2+Xe, Sr+Xe and SrXe systems. Eur. Phys. J. D 2019, 73, 90564. 10.1140/epjd/e2019-90564-7. [DOI] [Google Scholar]
- Heaven M. C.; Stolyarov A. V.. Potential Energy Curves for Alkali Metal-Rare Gas Exciplex Lasers. In 41st Plasmadynamics and Lasers Conference, Chicago, IL, June 28–July 1, 2010; American Institute of Aeronautics and Astronautics, 2010. 10.2514/6.2010-4877. [DOI]
- Miller W.; Rice C.; Hager G.; Rotondaro M.; Berriche H.; Perram G. High Pressure Line shapes of the Rb D1 and D2 lines for 4He and 3He collisions. J. Quant. Spectrosc. Radiat. Transfer 2016, 184, 118. 10.1016/j.jqsrt.2016.06.027. [DOI] [Google Scholar]
- Durand P.; Barthelat J. Theoretical Method to Determine Atomic Pseudopotentials for Electronic Structure Calculations of Molecules and Solids. Theor. Chem. Acta. 1975, 38, 283–302. 10.1007/BF00963468. [DOI] [Google Scholar]
- Müller W.; Flesch J.; Meyer W. Treatment of Inter Shell Correlation Effects in Ab Initio Calculations by Use of Core Polarization Potentials. Method and Application to Alkali and Alkaline Earth Atoms. J. Chem. Phys. 1984, 80, 3297. 10.1063/1.447083. [DOI] [Google Scholar]
- Hohm U.; Kerl K. Interferometric measurements of the dipole polarizability α of molecules between 300 and 1100 K. Mol. Phys. 1990, 69, 803–817. 10.1080/00268979000100611. [DOI] [Google Scholar]
- National Institute of Standards and Technology , August 2009. http://physics.nist.gov.
- Yuan J.; Zhang Z. Low-energy electron scattering from Kr and Xe atoms. J. Phys. B 1989, 22, 2581. 10.1088/0953-4075/22/16/015. [DOI] [Google Scholar]
- Hickling H. L.; Viehland L. A.; Shepherd D. T.; Soldán P.; Lee E. P. F.; Wright T. G. Spectroscopy of M+·Rg and transport coefficients of M+ in Rg (M = Rb–Fr; Rg = He–Rn). Phys. Chem. Chem. Phys. 2004, 6, 4233–4239. 10.1039/B405221H. [DOI] [Google Scholar]
- Tang K. T.; Toennies J. P. An improved simple model for the van der Waals potential based on universal damping functions for the dispersion coefficients. J. Chem. Phys. 1984, 80, 3726–3741. 10.1063/1.447150. [DOI] [Google Scholar]
- Hinde R. J. Mg–He and Ca–He van der Waals interactions: approaching the Born–Oppenheimer limit. J. Phys. B: At. Mol. Opt. Phys. 2003, 36, 3119. 10.1088/0953-4075/36/14/313. [DOI] [Google Scholar]
- Cohen J. S.; Schneider B. Ground and Excited States of Ne2 and Ne2+. Potential Curves with and without Spin-Orbit Coupling. J. Chem. Phys. 1974, 61, 3230. 10.1063/1.1682481. [DOI] [Google Scholar]
- Merritt J. M.; Han J.; Chang T.; Heaven M. C. Theoretical investigations of alkali metal: rare gas interaction potentials. Proc. SPIE 2009, 7196, 71960H. 10.1117/12.815155. [DOI] [Google Scholar]
- Buck U.; Pauly H. Interferenzen bei atomaren Stoßprozessen und ihre Interpretation durch ein modifiziertes Lennard-Jones-Potential. Z. Phys. 1968, 208, 390–417. 10.1007/BF01382701. [DOI] [Google Scholar]
- Medvedev A.; Meshkov V.; Stolyarov A.; Heaven M. C. Ab initio interatomic potentials and transport properties of alkali metal (M = Rb, Cs) - rare gas (Rg = He, Ne, Ar, Kr, Xe) media. Phys. Chem. Chem. Phys. 2018, 20, 25974–25982. 10.1039/C8CP04397C. [DOI] [PubMed] [Google Scholar]
- Zanuttini D.; Jacquet E.; Giglio E.; Douady J.; Gervais B. An accurate model potential for alkali neon systems. J. Chem. Phys. 2009, 131, 214104. 10.1063/1.3269801. [DOI] [PubMed] [Google Scholar]
- Palla A. D.; Carroll D. L.; Verdeyen J. T.; Readle J. D.; Spinka T. M.; Wagner C. J.; Eden J. G.; Heaven M. C. Multi-dimensional modeling of the XPAL system. Proc. SPIE 2010, 7581, 75810L. 10.1117/12.845917. [DOI] [Google Scholar]
- Chu X.; Dalgarno A. Molecular transition moment at large internuclear distances. Phys. Rev. A 2002, 66, 024701. 10.1103/PhysRevA.66.024701. [DOI] [Google Scholar]
- Pazyuk E. A.; Revina E. I.; Stolyarov A. V. Ab initio and long-range studies of the electronic transition dipole moments among the low-lying states of Rb2 and Cs2 molecules. J. Quant. Spectrosc. Radiat. Transfer 2016, 177, 283–290. 10.1016/j.jqsrt.2016.01.004. [DOI] [Google Scholar]





