Next Article in Journal
Ionization of Hydrogen Atom by Proton Impact—How Accurate Is the Ionization Cross Section?
Previous Article in Journal
Stark Broadening of Lyman-α in the Presence of a Strong Magnetic Field
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electronic Structure, Spectroscopy, Cold Ion–Atom Elastic Collision Properties, and Photoassociation Formation Prediction of the (MgCs)+ Molecular Ion

1
Laboratory of Interfaces and Advanced Materials, Faculty of Sciences of Monastir, University of Monastir, Monastir 5019, Tunisia
2
School of Physical Sciences, Indian Association for the Cultivation of Science (IACS), Jadavpur, Kolkata 700032, India
3
Department of Mathematics and Natural Sciences, School of Arts and Sciences, American University of Ras Al Khaimah, Ras Al-Khaimah P.O. Box 10021, United Arab Emirates
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Atoms 2023, 11(9), 121; https://doi.org/10.3390/atoms11090121
Submission received: 15 May 2023 / Revised: 7 August 2023 / Accepted: 11 August 2023 / Published: 15 September 2023
(This article belongs to the Special Issue Cold and Ultracold Molecular Collisions)

Abstract

:
In this paper, we extensively study the electronic structure, interactions, and dynamics of the (MgCs)+ molecular ion. The exchanges between the alkaline atom and the low-energy cationic alkaline earths, which are important in the field of cold and ultracold quantum chemistry, are studied. We use an ab initio approach based on the formalism of non-empirical pseudo-potential for Mg2+ and Cs+ cores, large Gaussian basis sets, and full-valence configuration interaction. In this context, the (MgCs)+ cation is treated as an effective two-electron system. Adiabatic potential energy curves and their spectroscopic constants for the ground and the first 20 excited states of 1,3Σ+ symmetries are determined. Furthermore, we identify the avoided crossings between the electronic states of 1,3Σ+ symmetries. These crossings are related to the charge transfer process between the two ionic limits, Mg/Cs+ and Mg+/Cs. Therefore, vibrational-level spacings and the transition and permanent dipole moments are presented and analyzed. Using the produced potential energy data, the ground-state scattering wave functions and elastic cross-sections are calculated for a wide range of energies. In addition, we predict the formation of a translationally and rotationally cold molecular ion (MgCs)+ in the ground-state electronic potential energy through a stimulated Raman-type process aided by ion–atom cold collision. In the low-energy limit (<1 mK), elastic scattering cross-sections exhibit Wigner law threshold behavior, while in the high-energy limit, the cross-sections act as a function of energy E go as E−1/3. A qualitative discussion about the possibilities of forming cold (MgCs)+ molecular ions by photoassociative spectroscopy is presented.

1. Introduction

In recent years, cold and ultracold molecules [1] have received considerable attention owing to successful demonstrations of cooling and trapping of molecules [2] and their importance in high-resolution spectroscopy [3]. Due to the characteristics of long-range ion–atom potential, ion–neutral collisions [4,5,6] are different from neutral–neutral and ion–ion collisions. The long-range ion–atom potential is described by −C4/r4, where C 4 = α q2 /(8πε0), α is the static electric polarizability of the atom, r is the ion–atom separation, q is the charge of the ion, and ε0 is the permittivity of free space. The scattering caused by this potential offers the main channel for the exchange of energy between the ions and the atoms. Accordingly, research on low-energy ion–neutral-atom collisions is very useful for the study of the sympathetic cooling of the translational motion of atomic ions [7,8] and for obtaining precise control of the internal degrees of freedom and external molecular ions [9]. In addition, interactions between the alkali atoms and alkaline earth ions (Be+, Mg+, Sr+, and Ca+) offer opportunities for new developments in the field of ultracold quantum matter, with the benefit of simpler and more reliable trapping compared to neutral molecules. Furthermore, a multitude of cold molecular ion species could be created, paving the way for rich chemistry at temperatures of a few mK or less [10]. The collision between cold atoms and ions is also relevant for important applications related to molecule formation in Bose–Einstein condensates [11] and quantum information. Photoassociation [12] and Feshbach resonance tuning [13] are two main experimental techniques for the coherent production of ultracold molecules from atoms.
Radiative emission during cold collisions between trapped laser-cooled Rb atoms and the alkaline earth ions Ca+, Sr+, and Ba+ has been studied theoretically by Aymar et al. [14] using an effective core potential based on quantum chemistry calculations of potential energy curves and transition dipole moments of the related molecular ions. Furthermore, many experimental groups have carried out experiments with various combinations of alkali atoms and alkaline earth atomic ions: Rb atoms with Ca+ [15,16] and Ba+ [17] ions. Moreover, recently, other groups have studied the energy scaling of the cold-atom–atom–ion three-body recombination Ba+ + Rb + Rb in the mK regime where a single 138Ba+ ion in a Paul trap is immersed in a cloud of ultracold 87Rb atoms [18], as well as the life and death of a cold BaRb+ molecule inside an ultracold cloud of Rb atoms [18]. In addition, an optical dipole trap of Rb atoms has also been merged in a Paul trap containing a few Ba+ atoms [18].
Concerning MgH+, the first evidence of the formation of molecular MgH+ ions in a laser-cooled ion trap was reported by Baba and Waki [19], who introduced air into an Mg+ trap. Drewsen and colleagues [20] carried out a more controlled experiment, introducing a thermal gas H2 or D2 into a laser-cooled Mg+ trap, thereby creating trapped MgH+ or MgD+ ions. Many other studies concerning magnesium hydride investigated its photodissociation [21], branching ratios [21], etc. More recently, Aymar et al. [22] carried out a study of the electron structure of the MgH+ ionic molecule containing potential energy curves, transition and permanent dipole moment spectra, and static polarizabilities. Using an ab initio approach, Khemeri et al. [23] performed a study of the electronic properties of this molecule in the adiabatic representation.
Concerning the MgLi+ ionic molecule, Boldyrev et al. [24] determined its ground-state spectroscopic constants using the split-valence basis MP2/6-311+G*. Pyykkö [25] employs two methods—Hartree–Fock (HF/6-31G*) and Møller–Plesset (MP)—to calculate the equilibrium distance, well depth, and vibration frequency for the same states. More recently, Yufeng Gao and Tao Gao [26] applied the ab initio program package MOLPRO to determine the potential energy curves and the permanent and transition dipole moments of the same molecules. They used the multireference configuration interaction and valence full-configuration interaction with large aug-cc-pCVQZ basis sets, taking into account the core–valence and scalar relativistic corrections.
A review of the scientific literature shows that experimental studies on the (MgCs)+ [27] cationic molecule are practically absent. We only noticed the existence of one recent theoretical study of (MgCs)+, in which the authors Smialkowski and Tomza [28] only reported spectroscopic constants for the ground state. Therefore, a more detailed and refined investigation of this molecular system is desirable. Methods fundamentally similar to those used for the treatment of the (MgLi)+ molecule [29,30] will be employed in the present study.
Several groups in the world have experimentally investigated atom–ion and cold molecular diatomic mixtures. The list of such mixtures includes Yb2+ [31], RbCa+ [16], CaBa+ [32], CaYb+ [33], RbYb+ [34], LiCa+ [35], RbCa+ [15], NaCa+ [36], RbSr+ [37], LiYb+ [38], KCa+ [39], RbBa+ [40], Rb2+ [41,42], and BaRb+ [43], as well as homonuclear mixtures [44]. Alkaline earth ions are a common choice as advanced methods for manipulation and detection of such ions have been developed over the years, creating prospects for their applications in quantum simulation and computations [45,46]. Atom–ion diatomic molecular systems can be produced via cold collision-induced charge-transfer radiative association, light-induced photoassociation [47,48,49,50,51], or photoionization of ultracold neutral molecules [42,44]. The effective core potential methods followed by CI and quantum chemistry calculations are employed to calculate the potential energy curves and dipole moments. In a very recent work [52] by our group, we investigated the electronic structure of the five diatomic molecular ions composed of a Ca+ ion interacting with an alkali metal atom: CaX+ (X =Li, Na, K, Rb, Cs). Also, the molecular ion structure of BeX+ (X = Na, K, Rb) systems has been theoretically studied [53], and the forming of a cold ion–atom mixture by stimulated Raman adiabatic process has been investigated [54]. In addition, an extensive study of the electronic structure, the cold ion–atom elastic collision properties, and the possibility of laser cooling BeCs+ molecular ions has recently been investigated [55].
Very recently, Smialkowski et al. theoretically investigated the ground-state electronic structure of single-charged molecular ions formed from two or three interacting alkali metal and alkaline earth metal atoms. They calculated the ground-state electronic properties of all diatomic AB+ and most of the triatomic A2B+ molecular ions consisting of Li, Na, K, Rb, Cs, Mg, Ca, Sr, Ba, and Yb atoms. Smialkowski and Tomza [28] carried out a systematic study of the electronic structure of MgX+ (X = Li, Na, K, Rb and Cs) ionic systems and presented the potential energy curves and spectroscopic constants for their ground states.
Here, we investigate the adiabatic potential energy curves for many electronic states of 1,3Σ+, 1,3Π, and 1,3Δ symmetries below the Mg2+Cs asymptotic limit. Then, we extract the spectroscopic constants from these curves (equilibrium distance Re, well depth De, electronic excitation energy Te, frequency ωe, harmonicity constant ωexe, and rotational constant Be). In addition, vibrational-level spacings and the dipole moment function are also presented and analyzed. By exploiting the ab- initio potential energy data, we study elastic collisions between an alkali ion Cs+ and a neutral alkaline earth atom Mg for the ground-state collisional threshold. Besides that, we predict and analyze the formation of translationally and rotationally cold molecular ions by a two-photon incoherent Raman process in the presence of two applied lasers.
This paper is organized as follows. In Section 2, the methodology used is briefly presented. We report our results for (MgCs)+ and discuss them in Section 3.

2. Theory and Computational Details

In this section, we give an overview of the technical details of the calculation of the (MgCs)+ molecular ion. Numerous studies on heteronuclear alkaline earth dimers such as (MgK)+ [49], (BeLi)+ [56], (BeH)+ [57], (CaRb)+, (SrRb)+, and (BaRb)+ [15] have been performed. We used the CIPSI package (Configuration Interaction by Perturbation of a Multi-configuration Wave Function Selected Iteratively) of the Laboratoire de Physique et Chimie Quantique of Toulouse in France [58]. The ionic system (MgCs)+ has 67 electrons. Therefore, we replaced the cores of the Mg2+ (10 electrons) and Cs+ (54 electrons) with non-empirical pseudo-potentials proposed by Barthelat and Durand [59,60,61]. Consequently, (MgCs)+ is treated as an effective two-electron system, where the two valence electrons are moving in the field of the two cores. The pseudo-potential is complemented by the corrections of core–core and core–valence correlations according to the operator formalism of Muller et al. [62]. In addition, the cut-off function reported by the formulation of Foucrault et al. [63] is taken to be a function of l, the orbital angular momentum, in order to treat separately the interaction of valence electrons of different spatial symmetry and the core electrons. The cut-off radii for the lowest valence s, p, d, and f one-electron states are reported in Table 1.
In the present work, the core polarizabilities are taken to be 0.46904 a03 [63] for Mg2+ and 15.117a03 [64,65], respectively. For magnesium, we used a large Gaussian basis set (9s, 7p, 5d, and 4f) composed of 83 functions. The diffuse orbital exponents have been optimized to reproduce (with good accuracy) all atomic states dissociating into the following: Mg+(3s, 3p, 4s, 3d, 4p, 5s, 4d, 4f, 5p, and 6s) and Mg (3s2 (1S), 3s3p (3P), 3s3p (1P), 3s4s (3S), 3s4s (1S), 3s3d (1D), 3s4p (3P), 3s3d (3D), 3s4p (1P), 3s5s (3S), 3s5s (1S), 3s4d (1D), 3s4d (3D), and 3s5p (3P)). After contraction, this basis was reduced to 7s/7p/4d/4f, and the function number decreased to only 76. Aymar et al. [22] used a (7s/5p/4d/2f) basis set contracted to (6s/5p/2d/2f); therefore, the function number decreases from 56 to 45, while for Cs, we used a (7s4p5d1f/6s4p4d) Gaussian basis set taken from [22,66]. Molecular energies for (MgCs)+, at the dissociation limits, are given in Table 2.
The quality of the used basis sets and cut-off radii is confirmed by the good agreement between our values and the experimental [67] and theoretical [68,69] atomic energy levels. The difference between our results and the experimental values for almost all energy levels is lower than 179 cm−1, which is found for the Mg (3s4s) atomic level. However, a difference of 1445 cm−1 is found for the Mg (3s3d) atomic level. Nevertheless, the agreement between the energy levels of Mg+ and the experimental ones is much better, seeing that we have reproduced Mg+ experimental binding energies exactly for 3s, 3p, 3d, and 4p. The difference with experimental values for the Mg+(4s) atomic level is equal to 1.53 cm−1, and for the other highly excited states 5s, 4d, 4f, 5p, and 6s, the discrepancy is between 22 and 105 cm−1.
Table 2. Asymptotic energy of the alkali earth (MgCs)+ electronic states (in cm−1): comparison between our calculated energy, Moore et al.’s [70] calculated energy, and the corresponding experimental dissociation.
Table 2. Asymptotic energy of the alkali earth (MgCs)+ electronic states (in cm−1): comparison between our calculated energy, Moore et al.’s [70] calculated energy, and the corresponding experimental dissociation.
StateAsymptotic Molecular StateThis WorkExperiment [68,70]
X (1) 1Σ+Mg (3s2) + Cs+00
A (2) 1Σ+Mg+ (3s) + Cs (6s)30,263.6230,398.60
C (3) 1Σ+Mg (3s3p) + Cs+35,050.5935,319.67
D (4) 1Σ+Mg+ (3s) + Cs (6p)41,811.3041,946.72
E (5) 1Σ+Mg (3s4s) + Cs+43,502.5743,472.29
F (6) 1Σ+Mg+ (3s) + Cs (5d)44,821.4044,956.60
G (7) 1Σ+Mg (3s3d) + Cs+46,402.5047,967.79
H (8) 1Σ+ Mg+ (3s) + Cs (7s)48,802.0248,934.58
I (9) 1Σ+Mg (3s4p) + Cs+49,346.1049,487.00
J (10) 1Σ+ Mg+ (3s) + Cs (7p)52,143.7452,284.65
a (1) 3Σ+ Mg (3s3p) + Cs+21,876.6021,994.46
c (2) 3Σ+ Mg+ (3s) + Cs (6s)30,263.6230,398.60
d (3) 3Σ+ Mg (3s4s) + Cs+41,196.7741,154.85
e (4) 3Σ+ Mg+ (3s) + Cs (6p)41,811.3041,946.72
f (5) 3Σ+ Mg+ (3s) + Cs (5d)44,821.4044,956.60
g (6) 3Σ+ Mg (3s4p) + Cs+47,857.8448,197.80
h (7) 3Σ+ Mg (3s3d) + Cs+47,956.3846,647.87
I (8) 3Σ+Mg+ (3s) + Cs (7s)48,802.0248,934.58
J (9) 3Σ+ Mg (3s5s) + Cs+51,871.8151,976.72
k (10) 3Σ+ Mg+ (3s) + Cs (7p)52,143.7452,284.65

2.1. Results and Discussions

2.1.1. Adiabatic Potential Energy and Spectroscopic Constants

Using the method reported in the previous section, a full and expanded calculation was performed for many electronic states of 1,3Σ+, symmetries for the (MgCs)+ cationic molecule. These states dissociate into Mg+ (3s, 3p, 4s, 3d, 4p, 5s, 4d, 4f, 5p, and 6s) + Cs (6s, 6p, 4d, 5d, 7s, and 7p) and Mg {(3s2 (1S), 3s3p (3P), 3s3p (1P), 3s4s (3S), 3s4s (1S), 3s3d (1D), 3s4p (3P), 3s3d (3D), (1P), 3s5s (3S), 3s5s (1S), 3s4d (1D), 3s4d (3D), and 3s5p (3P))} + Cs+. The adiabatic potential energy is performed for an interval of intermolecular distances from 3.50 to 200.00 a0 with a step of 0.1 a0 around the avoided crossings. The potential energy curves as functions of the internuclear distance R for the 1,3Σ+ electronic states are drawn, respectively, in Figure 1 and Figure 2.
In order to assess the precision of our calculated potential energy curves for the states previously studied, we have extracted the spectroscopic constants (equilibrium distance Re, well depth De, electronic excitation energy Te, frequency ωe, harmonicity constant ωexe, and rotational constant Be for the equilibrium separation Re). These constants were determined by least-squares interpolation of the vibrational energy levels. Unfortunately, no experimental spectroscopic information for (MgCs+) has been published yet. However, Smialkowski and Tomza [28] reported a theoretical calculation for the ground state. They found Re = 7.85 a0, De = 1861 cm−1, ωe = 73.2cm−1, and Be = 0.0481 cm−1, which are close to our values (Re = 7.70 a0, De = 2047 cm−1, ωe = 73.2cm−1, and Be = 0.049364 cm−1) (see Table 3). Therefore, the spectroscopic constants for the excited states are presented here for the first time. Consequently, we think that these potential energies and spectroscopic data can initiate future experimental and theoretical investigations on this molecular ionic system. In Table 3, we report the spectroscopic constants for the ground and the excited states for all symmetries.
The potential energy curves for the 20 first 1,3Σ+ states of (MgCs)+ molecular ion are plotted in Figure 1 and Figure 2. These states are labelled by increasing numbers from 1 to 10 for both singlet and triplet states
The ground (X1Σ+) and the first excited (21Σ+) states are well separated from the highest states of the same symmetry, and they maintain an energy gap of about 42,012 cm−1. Both are found with unique minimums at 7.7 and 11.45 a0 and well depths of 2047 and 2662 cm−1, respectively. The first three excited triplet states (1–3)3Σ also exhibit unique minimums at 8.58, 14.98, and 17.20 a0, with significant well depths of 4504, 813, and 2658 cm−1, respectively. An accurate analysis of these curves shows that the most strongly bound states are the first excited states of both singlet and triplet states with well depths of 2662 and 4504 cm−1 located at 11.45 and 8.58 a0, respectively.
An interesting feature of the higher excited states of all symmetries of 1,3Σ+ states has been noticed. The shape becomes more involved, with double wells and potential barriers. This feature is due to the presence of many avoided crossings at short and large values of internuclear distances between excited states of the same symmetries. The positions of such avoided crossings are presented in Figure 1 and Figure 2. These series of avoided crossings are related to underlying charge transfer between the ionic configurations Mg+/Cs and Mg/Cs+. Later, these avoided crossings will be discussed in light of the behavior of the permanent dipole moments.
The PECs of excited states starting from 41Σ+ present double wells and barriers, which are the consequence of avoiding crossings that happen at specific internuclear distances. These avoided crossings are related to underlying charge transfer, as can be seen from the change of location of the positive charge at dissociation limits between two neighbor states. The positive charge moves from one atom to the second one. The existence of these avoided crossings will generate large nondiabetic couplings. This same behavior has been seen for other symmetries, such as 3Σ+ states.

2.1.2. Vibrational Levels

For a better understanding of the potential energy curve structure, the vibrational levels for all attractive states are computed using the Numerov algorithm. The interpolation of the potential curves is carried out with 12,000 points and a grid ranging from 3.5 to 180.0 a0.
The vibrational spacings (Ev+1–Ev) as a function of the vibrational quantum number v are reported in Figure 3 for the 1,3Σ+ states. As usual, these spacings are not constant due to the important anharmonicity observed in the potential energy curves. The number of the vibration levels for the ground and the first excited states are 80 and 180, respectively. The vibrational energy spacing of these two potentials shows a linear behavior for low vibrational levels, and then it vanishes at the dissociation limits.
The spacing between vibrational levels decreases gradually as vibrational energy increases. This behavior is particularly prominent for vibrational levels of the PECs that reflect the strong anharmonicity. The overall pattern of energy spacings of different electronic states of the molecular ion is similar. However, for some states, irregularities related to the avoided crossings are visible, e.g., for the 1Σ+ states.

2.1.3. Permanent and Transition Dipole Moments

In addition to the potential energy curves, the practical implementation of cold molecule formation via photoassociation requires knowledge of their electronic properties, such as the radial variation of permanent or transition dipole moments. Furthermore, the knowledge of the dipole function of molecular systems can be considered a sensitive test for the accuracy of the calculated electronic wave functions and energies.
Keeping this broad perspective of the ion–atom systems in mind, we have determined the permanent and transition dipole moments for the same large and dense grid of internuclear distances used for the potential energy curves.

Permanent Dipole Moments

The permanent and transition dipole moments are determined by considering the origin at atom Mg, and where the distance R between Mg/Mg+ and Cs/Cs+ stands for the internuclear distance between them. We calculated the permanent dipole moments as a function of R; the intermolecular distance for all studied states (1−101Σ+, 1−103 Σ+) are displayed in Figure 4 and Figure 5.
At a short and intermediate internuclear distance, the permanent dipole moments present undulations with abrupt variation between neighbor electronic states. For example, there is an abrupt variation between the 81Σ+ and 91Σ+ states located at 33.5 a0. We remark that this particular distance represents approximately the position of the avoided crossing between the two potential energy curves that we noticed before. We can conclude that the significant changes in permanent dipole moment at a small internuclear distance are due to the change in the polarity of the molecule. In addition, the positions of the irregularities in the R-dependence of the permanent dipole are correlated to the avoided crossings between potential energy curves, which are both manifestations of abrupt changes in the character of the electronic wave functions.
For the large internuclear distance, the permanent dipole moment of the X1Σ+, 31Σ+, 51Σ+, 71Σ+, and 91Σ+ states dissociating into Mg (3s2, 3s3p, 3s3p, 3s4s, 3s4s, 3s3d, 3s4p, 3s3d, 3s5s, 3s5s, 3s4d, 3s4d, and 3s5p) + Cs+ are significant and yield a pure linear behavior function of R, especially at intermediate and large distances. For the remaining states 21Σ+, 41 Σ+, 6 1Σ+, and 10 1Σ+ dissociating into Mg+ (3s, 3p, 4s, 3d, 4p, 5s, 4d, 4f, 5p, and 6s) + Cs (6s, 6p, 4d, 5d,7s and 7p), we note significant permanent dipole moments in specific regions and then they vanish swiftly at large distances.
For the permanent dipole moments of the first ten states of 3Σ+ symmetry, similar observations were noticed as for the 1Σ+ states. They exhibit undulations with abrupt variations at a short and intermediate internuclear distance R, such as between the two states 83Σ+ and 93Σ around 27.7 a0. We remark that the permanent dipole moments of these states, one after another, behave in a straight line on the same curve and then drop to zero at particular distances corresponding to the avoided crossings between neighboring electronic states.

Transition Dipole Moments

In addition, we have determined the transition dipole moment functions for the (MgCs)+ ionic molecular system between states of similar symmetries. As the transition dipole moments between adjacent states are the most significant, here, we only present the transition between neighbor states. They are displayed in Figure 6.
Theoretically, Fedorov et al. [27] reported dipole moments for the molecular ions (MgLi)+ and (MgNa)+. They have calculated the transition dipole moments for the latter systems; however, no data for the (MgCs)+ ionic molecule are available, and these important physical quantities are presented here for the first time. We note a meaningful variation with numerous extremum, which can be assigned to avoid crossing positions between adjacent states. We also note a change in the sign of some dipole transitions, such as between 31Σ+->41Σ+ states. This change in sign corresponds to a sudden change in electronic wave functions, which is usually related to the avoided crossings between the potential energy curves.
For the triplet sigma states, the transition dipole moment curves between the neighbor states are displayed in Figure 6. We note similar behaviors observed previously for the singlet states. For both multiplicities,1,3Σ states, the weakly avoided crossing series already found between neighboring states lead to abrupt variations in the transition dipole moments.

2.1.4. Radiative Lifetime

In this section, we present the vibrational state lifetimes for A1 Σ+ (21 Σ+) excited state. Given a vibrational level of the first excited electronic state, two possible transitions can occur: bound–bound and bound-free transitions. The radiative lifetime of a vibrational level corresponding to only bound–bound transitions is given by
τ υ = 1 Γ υ   where   Γ υ = υ = 0 n v t A υ υ
Avv’ is the Einstein coefficient linking, for example, the A1Σ+(v′) and X1Σ+(v) levels. Zemke et al. [71] have previously shown that there is a missing contribution in the radiative lifetime. It corresponds to the bound-free transitions, which are more significant for the higher vibrational levels close to the dissociation limit of the excited electronic state, found in the form of the continuum radiation to states above the dissociation limit of the ground state. It corresponds to the contribution of the bound-free transition missed in Equation (1). It is related to the transition between the vibrational level v′, which belongs to the excited electronic state, to the continuum of the ground state or the lower state in general. Such contribution is not negligible, as was demonstrated by Partridge et al. [72], Zemke et al. [71], and Berriche and Gadea [73,74]. It is more important for the higher vibrational levels due to the difference in the location of the potential wells. We calculate this contribution using two different approximations, “Franck-Condon” [72,73] and the “sum rule method” [72,75] approximation. In the Franck–Condon (FC) approximation proposed by Zemke et al. [71], the bound-free contribution is given by
A υ ( b o u n d f r e e ) = 2.14198639 × 10 10 | μ ( R υ + ) | 2 F C υ , c o n t ( Δ E ) 3 υ , c o n t
where Δ E υ , c o n t = E υ E a s  is the energy difference between the vibrational level v and the energy of the asymptotic limit of the lower electronic state to which the continuum belongs. The quantity μ ( R υ + ) corresponds to the transition dipole moment at the right external turning point of the vibrational level v′.
F C υ , c o n t = | χ υ | χ E | 2 d E = 1 υ = 0 n v t | χ υ | χ υ | 2
The approximate sum rule approach, implemented by Pazyuk et al. [76], allows the production of the radiative lifetime components for diatomic vibronic states, including the bound-free contribution. In addition, this approximation has a high efficiency for non-diagonal systems, particularly for those with significant continuum contributions. The radiative lifetime using the sum rule method is given by
1 τ = R min R max ϕ ν ( R ) ( D ( R 2 ) ) ( Δ ( R 3 ) ) ϕ ν ( R ) d R
φ υ ( R ) is the wave function of the vibrational level belonging to the A1Σ+ excited electronic state. D(R) is the transition dipole moment between the ground X 1Σ+ and the first excited A1Σ+ states. ∆U(R) is the energy difference between the ground X1Σ+ and the first excited A1Σ+ states. The total radiative lifetime, taking into account the bound–bound and bound-free contributions, using the two approaches, FC approximation and the approximate sum approach, are presented in Table 4.
To the best of our knowledge, this is the first time that the radiative lifetimes of the vibrational levels of the A1Σ+ state of the (MgCs)+ molecules have been presented. These radiative lifetimes increase with the vibrational level v in a range of 10–43 ns.

2.1.5. Ion–Atom Elastic Collisions

In this section, we study cold elastic collisions between an alkaline earth atom and an alkali ion in the ground electronic state (X1Σ+) of the (MgCs)+ molecular system. The ground-state potential asymptotically correlates to the heavier alkali element in the ionic state (Cs+) and the alkaline element in its neutral ground state (Mg). The first singlet excited state 21Σ+ asymptotically corresponds to the charge-exchanged state of the X1Σ+ potential. We use Mg + Cs+ as a prototype system, and several results equally apply to other alkaline-earth–alkali atom–ion collisional systems. The elastic collisions between the alkaline earth atom, Mg, and alkali ion, Cs+, for the ground-state potential (X1Σ+) can be represented as
M g + C s + { M g C s + ( X Σ + ) } M g + C s +
The partial-wave Schrödinger equation for the ion–atom elastic collision can be expressed as
[ d 2 d r 2 + k 2 2 μ 2 V ( r ) l ( l + 1 ) r 2 ] ψ l ( k r ) = 0
where r is the internuclear separation between an atom and ion, μ is the reduced mass of an ion–atom colliding pair, and ψ l ( k r ) is the wave function for the lth partial wave. The wave function satisfies the asymptotic boundary condition ψ l ( k r )   ~   s i n [ k r l π 2 + η l ] . The above second-order differential Equation 6 can be solved by the Numerov–Cooley algorithm [77] using three-point recursion relation. Initially, we set the values of the wave function at two initial consecutive points at small internuclear separation r ~ 0 . Thereafter, the code calculates the wave function at the third point by utilizing the three-point recursion relation. Then, the second and third points are reset to the initial point, and the wave function is calculated for the fourth point. This process is continued until the asymptotic boundary is attained. The initial boundary conditions are included by expanding the interaction potential at small r and solving the Schrödinger equation at this limit. The boundary condition at the asymptotic limit is incorporated by setting the condition that the effective long-range part of potential V(r) is much less than the collisional energy E (at least one-tenth of E). We considered a constant step size of 0.01 during the propagation of the code.
The total elastic scattering cross-section is given by
σ t o t = 4 π k 2 l = 0 ( 2 l + 1 ) s i n 2 ( η l )
where k is the wave number related to the collisional energy E = ħ 2 k 2 2 μ and ηl being the phase shift for the l-th partial wave. As with an increase in energy, a large number of partial waves start to contribute to elastic scattering cross-section. Under this condition, the scattering cross-section is approximated as [4]
                  σ t o t π ( μ C 4 2 2 ) 1 3 ( 1 + π 2 16 ) E 1 3
The coefficient C4 is related to the static electric dipole polarizability of the concerned atom.
But, at a very low energy limit, i.e., E 0 , the phase shift follows the Wigner threshold laws. At this energy regime, the phase shift is related to the s-wave scattering length   a s = lim k 0 t a n η 0 k . The scattering length carries the information about the interaction potential of a system. A positive (negative) scattering length is associated with repulsive (attractive) interaction.
In Figure 7, we show the logarithm of total elastic scattering cross-sections (σtot) in a.u. as a function of the logarithm of energy E in K for ground-state collisions between neutral alkaline earth atom Mg and alkali ion Cs+. In order to calculate the convergence results in σtot, we need 59 partial waves.
In our calculations, we have included 70 partial waves, since, with the increase in energy, large numbers of partial waves start to contribute to the total elastic scattering cross-section. Considering the logarithm on both sides of Equation (8), a straight line is obtained with a slope equal to −1/3. From the linear fitting of log σtot vs. log E of Figure 7, we have verified the numerically calculated slope of the straight line is quite close to the theoretical value.

2.1.6. Two-Photon Photoassociation: Molecular Ion Formation

In this section, we explore the molecular ion formation in the ground-state electronic potential X1Σ+ by a two-photon incoherent Raman process in the presence of two applied lasers, L1 and L2, respectively. The formation of the molecular ion needs two steps. Initially, the atom–ion pair present in the ground-state scattering continuum (X1Σ+) is irradiated with laser L1 with a suitable frequency to form the molecular ion in one of the bound levels of the first excited potential 21Σ+. Once the molecular ion is formed, laser L2 is turned on to de-excite the molecular ion to the ground electronic potential followed by bound–bound transition.
Here, the rate coefficient of the first step concerned, which is known as single-photon photoassociation (PA) at a temperature T, is given by [78]
K P A = π v r e l k 2 l = 0 ( 2 l + 1 ) | S P A ( E , l , ω l ) | 2
where vrel is the relative velocity of the interacting particles, SPA is the scattering matrix element, and implies an averaging over thermal velocity distribution. At a very low-temperature limit, the energy of the interacting particle is such that their dynamics may be described by the s-wave scattering. If one considers a Maxwellian velocity distribution of an ultracold mixture of gas, Equation (9) can be represented as
K P A ( T , ω L ) = 1 h Q T l = 0 ( 2 l + 1 ) 0 | S P A ( E , l , ω L ) | 2 e β E d E          
Here, Q T = ( 2 π μ k B T / h 2 ) 3 / 2 is the translational partition function, β = 1/kbT, and kB is the Boltzmann constant. The scattering S-matrix element is given by
| S P A | 2 = γ s Γ l δ E 2 + ( γ / 2 ) 2
The scattering phase shift (η) is related to the scattering T-matrix as T = 1 e 2 i η 2 i = e i η sin η , while the quantity | T | 2 = sin 2 η . Again, the scattering S-matrix is related to the T-matrix as S = 1 2 i T and the quantity | S | 2 = 1 + 4 sin 2 η . Therefore, in terms of phase shift and T-matrix, Equation (11) becomes
| T | 2 = sin 2 η = 1 4 [ γ s Γ l δ E 2 + ( γ / 2 ) 2 1 ]
where detuning parameters are defined as δ E = E ħ + δ v , j , δ v , j = ω L ω v , j with E v , j = ħ ω v , j is the binding energy of the excited co-vibrational state, ω L is the frequency of the applied laser L1. Here, γs is the spontaneous line width, and Γl is the stimulated line width. We consider that other decay processes, such as molecular predissociation, are either negligible or not present in our system.
The PA rate is determined, basically, by stimulated line width and can be expressed by the Golden rule as
Γ l = π I ε 0 c | ϕ v , j | D ( r ) | ψ l ( k r ) | 2
We termed the quantity D v , j , l = ϕ v , j | D ( r ) | ψ l ( k r ) as the radial transition dipole matrix element between the unit normalized bound state wave function ϕv,j and energy normalized continuum wave function ψl (kr), and D(r) is the transition dipole moment. Here, I is the intensity of the laser L1, c is the velocity of light, and ε0 is the vacuum permittivity.
In order to calculate the quantity Dv,j,l for the optimal production of molecular ions by the PA process, we have to calculate the scattering state and suitable bound state wave functions. We solve the partial-wave Schrödinger Equation (6) by standard renormalized Numerov–Cooley method [78] for the determination of the wave functions that we discussed. In general, molecular dipole transitions between two bound states or between continuum and bound states are controlled by the FC principle. The free-bound FC value is highest for the bound state ϕ v , j (v = 17, j = 1) of the first excited potential 21 Σ + of (MgCs)+ system. The PA rate coefficient is calculated according to Equation (9) to form the molecular ion in the excited electronic state 21 Σ + . In Figure 8, we have shown the rate coefficient of PA as a function of energy E (in Kelvin), considering the atom–ion pairs Mg-Cs+ are initially lying in the ground-state scattering continuum. We considered the laser intensity L1 = 10 W/cm2 tuned at the PA resonance. We find that our calculated PA rate coefficient is consistent with the results of Li+-Be [79] and K+-Mg [49] systems. Finally, in Figure 9, we have shown the variation of the rate of ion–atom PA as a function of the intensity of the applied laser L1 considering the temperature at 10 μK. We have observed that a saturation effect on the ion–atom PA rate occurs near the laser intensity I = 35 kW/cm2.
Now we discuss the formation of a ground-state molecular ion by stimulated Raman-type process in the presence of a second applied laser L2 tuned near a bound–bound transition between the potentials 21 Σ + and X1 Σ + , respectively. To find the efficacy of the coherent laser coupling between the two bound states, we calculate the Rabi coupling Ω given by
ħ Ω = ( I 4 π c ε 0 ) 1 / 2 | ϕ v , j | D ( r ) . ϵ | ϕ v j |  
where I is the intensity of the second applied laser L2 tuned to the transition frequency, ϵ is the unit vector of laser polarization, and ϕ v , j and ϕ v j are the two bound states of excited (21 Σ + ) and ground (X1 Σ + ) state electronic potentials, respectively. We have calculated the Rabi frequencies between the selected bound state (v = 15, j = 1) of the first excited molecular potential and different bound states of the ground electronic potential, considering an intensity of laser L2 = 10 kW/cm−2. We observed that the bound–bound Rabi coupling is maximum (Ω = 150 MHz) for the bound state ϕ v j ( v =16, j = 0 ) of X1 Σ + potential. Comparing this value with the spontaneous line width (γ = 120 kHz) of the excited bound state calculated using the formula [80], we found that bound–bound Rabi coupling (Ω) is several orders of magnitude higher. Therefore, we may infer that the ground-state molecular ion formation is likely to be possible by a stimulated Raman-type process [49,80,81].

3. Conclusions

In this work, an extended and complete study devoted to the (MgCs)+ molecular ion was presented. We have performed precise ab initio calculations for the potential energy curves and their related spectroscopic constants for the ground and the 20 low-lying excited states of 1,3Σ+ symmetries below the Mg2+Cs asymptotic limit.
The computational method used is based on a non-empirical pseudo-potential approach for Mg2+ and Cs+ cores, which allows treating the ionic molecule as a two-electron system where full configuration interaction FCI calculations can be easily performed. An interesting series of avoided crossings have been observed between the same symmetry of the 1,3Σ+ states. These avoided crossings often lead to relatively irregular dipole moment curves, which are both manifestations of abrupt changes in the character of the electronic wave functions with underlying strong interactions between the electronic states. These interactions could give rise to important charge transfer for collisions between the atom–ion combinations. In addition, the vibrational levels and their spacings were investigated for all attractive states. Finally, electric permanent and transition dipole moments were calculated. As expected, the permanent dipole moments of the electronic states dissociating into Mg {(3s2 (1S), 3s3p (3P), 3s3p (1P), 3s4s (3S), 3s4s (1S), 3s3d (1D), 3s4p (3P), 3s3d (3D), (1P), 3s5s (3S), 3s5s (1S), 3s4d (1D), 3s4d (3D), and 3s5p (3P))}+Cs+ have shown an almost linear features function of R, especially for intermediate and large internuclear distances. Moreover, the abrupt changes in the permanent dipole moment are localized at particular distances corresponding to the avoided crossings between the two neighboring electronic states.
In this study, the ab initio calculations were carefully performed to produce accurate data, even though there is a lack of experimental studies for this molecular ion. However, theoretically, only the ground state of (MgCs)+ was studied in a recent work published by Smialkowski and Tomza [28]. We expect that our data can serve as a solid foundation for future and advanced dynamic studies for the (MgCs)+ molecular ion.
These precise data were then used for elastic diffusion calculations between Mg and Cs+ at low temperatures. In fact, a theoretical understanding of low-energy atom–ion scattering can help probe the dynamics of quantum gases. Furthermore, we theoretically predict the formation of a cold molecular ion in translation and rotation by an incoherent two-photon Raman process. The formation of such cold molecular ions is very important in the research domain of ultracold chemistry and precision spectroscopy. Our data can be used to understand several phenomena, like charge transfer, diffusion, and partial-wave phase locking, and is comparable with the recently studied Li-Ba+ system [81]. We believe that our analysis, including the ab initio data, may serve as a solid platform for future study on (MgCs)+ molecular ionic systems.

Author Contributions

Conceptualization, H.B., M.F., D.S. and B.D.; Methodology, H.B., M.F., D.S. and B.D.; Investigation, M.F. and D.S.; Writing-original draft preparation, M.F. and D.S.; Writing-review and editing, H.B., M.F. and D.S. All authors have read and agreed to the published version of the manuscript.

Funding

H.B. would like to acknowledge the financial support of the American University of Ras Al Khaimah under seed grant no. AAS/003/21.

Data Availability Statement

Not applicable.

Acknowledgments

H.B. would like to acknowledge the financial support of the American University of Ras Al Khaimah under seed grant no. AAS/003/21.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Carr, L.D.; DeMille, D.; Krems, R.V.; Ye, J. Cold and ultracold molecules: Science, technology and applications. New J. Phys. 2009, 11, 055049. [Google Scholar] [CrossRef]
  2. Hummon, M.T.; Yeo, M.; Stuhl, B.K.; Collopy, A.L.; Xia, Y.; Ye, J. 2D Magneto-Optical Trapping of Diatomic Molecules. Phys. Rev. Lett. 2013, 110, 143001. [Google Scholar] [CrossRef] [PubMed]
  3. Stwalleyand, W.C.; Wang, H. Photoassociation of Ultracold Atoms: A New Spectroscopic Technique. J. Mol. Spectrosc. 1999, 195, 194. [Google Scholar] [CrossRef] [PubMed]
  4. Côté, R.; Dalgarno, A. Ultracold atom-ion collisions. Phys. Rev. A 2000, 62, 012709. [Google Scholar] [CrossRef]
  5. Idziaszek, Z.; Calarco, T.; Julienne, P.S.; Simoni, A. Quantum theory of ultracold atom-ion collisions. Phys. Rev. A 2009, 79, 010702. [Google Scholar] [CrossRef]
  6. Krych, M.; Skomorowski, W.; Pawłowski, F.; Moszynski, R.; Idziaszek, Z. Sympathetic cooling of the Ba+ ion by collisions with ultracold Rb atoms: Theoretical prospects. Phys. Rev. A 2011, 83, 032723. [Google Scholar] [CrossRef]
  7. Ravi, K.; Lee, S.; Sharma, A.; Werth, G.; Rangwala, S. Cooling and stabilization by collisions in a mixed ion–atom system. Nat. Commun. 2012, 3, 1126. [Google Scholar] [CrossRef] [PubMed]
  8. Sivarajah, I.; Goodman, D.S.; Wells, J.E.; Narducci, F.A.; Smith, W.W. Evidence of sympathetic cooling of Na+ ions by a Na magneto-optical trap in a hybrid trap. Phys. Rev. A 2012, 86, 063419. [Google Scholar] [CrossRef]
  9. Boyarkin, O.V.; Mercier, S.R.; Kamariotis, A.; Rizzo, T.R. Electronic spectroscopy of cold, protonated tryptophan and tyrosine. J. Am. Chem. Soc. 2006, 128, 2816–2817. [Google Scholar] [CrossRef]
  10. Willitsch, S.; Bell, T.M.; Gingell, A.D.; Procter, S.R.; Softley, T.P. Cold Reactive Collisions between Laser-Cooled Ions and Velocity-Selected Neutral Molecules. Phys. Rev. Lett. 2008, 100, 043203. [Google Scholar] [CrossRef]
  11. Schmid, S.; Härter, A.; Denschlag, J.H. Dynamics of a Cold Trapped Ion in a Bose-Einstein Condensate. Phys. Rev. Lett. 2010, 105, 133202. [Google Scholar] [CrossRef] [PubMed]
  12. Jones, K.M.; Tiesinga, E.; Lett, P.D.; Julienne, P.S. Ultracold photoassociation spectroscopy: Long-range molecules and atomic scattering, Rev. Mod. Phys. 2006, 78, 483. [Google Scholar] [CrossRef]
  13. Chin, C.; Grimm, R.; Julienne, P.; Tiesinga, E. Feshbach resonances in ultracold gases. Rev. Mod. Phys. 2010, 82, 1225–1286. [Google Scholar] [CrossRef]
  14. Da Silva, H., Jr.; Raoult, M.; Aymar, M.; Dulieu, O. Formation of molecular ions by radiative association of cold trapped atoms and ions. New J. Phys. 2015, 17, 045015. [Google Scholar] [CrossRef]
  15. Hall, F.H.J.; Aymar, M.; Bouloufa-Maafa, N.; Dulieu, O.; Willitsch, S. Light-Assisted Ion-Neutral Reactive Processes in the Cold Regime: Radiative Molecule Formation versus Charge Exchange. Phys. Rev. Lett. 2011, 107, 243202. [Google Scholar] [CrossRef] [PubMed]
  16. Hall, F.H.; Eberle, P.; Hegi, G.; Raoult, M.; Aymar, M.; Dulie, O.; Willitsch, S. Formation of molecular ions by radiative association of cold trapped atoms and ions. Mol. Phys. 2013, 111, 2020. [Google Scholar] [CrossRef]
  17. Mohammadi, A.; Krükow, A.; Mahdian, A.; Deiß, M.; Pérez-Ríos, J., Jr.; da Silva, H. Life and death of a cold BaRb+ molecule inside an ultracold cloud of Rb atoms. Phys. Rev. Res. 2021, 3, 013196. [Google Scholar] [CrossRef]
  18. Schmid, S.; Härter, A.; Frisch, A.; Hoinka, S.; Denschlag, J.H. An apparatus for immersing trapped ions into an ultracold gas of neutral atoms. Rev. Sci. Instrum. 2012, 83, 053108. [Google Scholar] [CrossRef]
  19. Baba, T.B.T.; Waki, I.W.I. Cooling and Mass-Analysis of Molecules Using Laser-Cooled Atoms. Jpn. J. Appl. Phys. 1996, 35, L1134. [Google Scholar] [CrossRef]
  20. Bertelsen, A.; Vogelius, I.S.; Jørgensen, S.; Kosloff, R.; Drewsen, M. Photo-dissociation of Cold MgH+ ions. Eur. Phys. J. D 2004, 31, 403. [Google Scholar] [CrossRef]
  21. Mølhave, K.; Drewsen, M. Formation of translationally cold MgH+ and MgD+ molecules in an ion trap. Phys. Rev. A. 2000, 62, 011401. [Google Scholar] [CrossRef]
  22. Aymar, M.; Guérout, R.; Sahlaoui, M.; Dulieu, O. Electronic structure of the magnesium hydride molecular ion. J. Phys. B: At. Mol. Opt. Phys. 2009, 42, 154025. [Google Scholar] [CrossRef]
  23. Khemiri, N.; Dardouri, R.; Oujia, B.; Gadéa, F.X. Ab Initio Investigation of Electronic Properties of the Magnesium Hydride Molecular Ion. J. Phys. Chem. A 2013, 117, 8915–8924. [Google Scholar] [CrossRef] [PubMed]
  24. Boldyrev, A.I.; Simons, J.; von, R. Schleyer, P. Ab initio study of the electronic structures of lithium containing diatomic molecules and ions. J. Chem. Phys. 1993, 99, 8793–8804. [Google Scholar] [CrossRef]
  25. Pekka Pyykk, Ö. Ab initio study of bonding trends among the 14-electron diatomic systems: From B24- to F24+. Mol. Phys. 1989, 67, 871. [Google Scholar] [CrossRef]
  26. Gao, Y.; Gao, T. Ab initio study of ground and low-lying excited states of MgLi and MgLi+ molecules with valence full configuration interaction and MRCI method. Mol. Phys. 2014, 112, 3015. [Google Scholar] [CrossRef]
  27. Fedorov, D.A.; Barnes, D.K.; Varganova, S.A. Ab initio calculations of spectroscopic constants and vibrational state lifetimes of diatomic alkali-alkaline-earth cations. J. Chem. Phys. 2017, 147, 124304. [Google Scholar] [CrossRef] [PubMed]
  28. Smialkowski, M.; Tomza, M. Interactions and chemical reactions in ionic alkali-metal and alkaline-earth-metal diatomic AB+ and triatomic A2B+ systems. Phys. Rev. A. 2020, 101, 012501. [Google Scholar] [CrossRef]
  29. ElOualhazi, R.; Berriche, H. Electronic Structure and Spectra of the MgLi+ Ionic Molecule. J. Phys. Chem. A 2016, 120, 452. [Google Scholar] [CrossRef]
  30. Bala, R.; Nataraj, H.S.; Abe, M.; Kajita, M. Calculations of electronic properties and vibrational parameters of alkaline-earth lithides: MgLi+ and CaLi+. Mol. Phys. 2018, 117, 712–725. [Google Scholar] [CrossRef]
  31. Grier, A.T.; Cetina, M.; Oručević, F.; Vuletić, V. Observation of Cold Collisions between Trapped Ions and Trapped Atoms. Phys. Rev. Lett. 2009, 102, 223201. [Google Scholar] [CrossRef]
  32. Sullivan, S.T.; Rellergert, W.G.; Kotochigova, S.; Hudson, E.R. Role of Electronic Excitations in Ground-State-Forbidden Inelastic Collisions Between Ultracold Atoms and Ions. Phys. Rev. Lett. 2012, 109, 223002. [Google Scholar] [CrossRef]
  33. Rellergert, W.G.; Sullivan, S.T.; Kotochigova, A.; Petrov, K.S.; Chen, S.; Schowalter, J.; Hudson, E.R.; Millikelvin, E. Reactive Collisions between Sympathetically Cooled Molecular Ions and Laser-Cooled Atoms in an Ion-Atom Hybrid Trap. Phys. Rev. Lett. 2011, 107, 24320. [Google Scholar]
  34. Zipkes, C.; Palzer, S.; Sias, C.; Köhl, M. A trapped single ion inside a Bose–Einstein condensate. Nature 2010, 464, 388–391. [Google Scholar] [CrossRef]
  35. Haze, S.; Hata, S.; Fujinaga, M.; Mukaiyama, T. Observation of elastic collisions between lithium atoms and calcium ions. Phys. Rev. A 2013, 87, 052715. [Google Scholar] [CrossRef]
  36. Smith, W.; Goodman, D.; Sivarajah, I.; Wells, J.; Banerjee, S.; Côté, R.; Michels, H.; Mongtomery, J.A.; Narducci, F. Experiments with an ion-neutral hybrid trap: Cold charge-exchange collisions. Appl. Phys. B 2014, 114, 75. [Google Scholar] [CrossRef]
  37. Meir, Z.; Sikorsky, T.; Ben-Shlomi, R.; Akerman, N.; Dallal, Y.; Ozeri, R. Dynamics of a Ground-State Cooled Ion Colliding with Ultracold Atoms. Phys. Rev. Lett. 2016, 117, 243401. [Google Scholar] [CrossRef] [PubMed]
  38. Fürst, H.; Feldker, T.; Ewald, N.V.; Joger, J.; Tomza, M.; Gerritsma, R. Dynamics of a single ion-spin impurity in a spin-polarized atomic bath. Phys. Rev. A 2018, 98, 012713. [Google Scholar] [CrossRef]
  39. yothi, S.; Egodapitiya, K.N.; Bondurant, B.; Jia, Z.; Pretzsch, E.; Chiappina, P.; Shu, G.; Brown, K.R. A hybrid ion-atom trap with integrated high resolution mass spectrometer. Rev. Sci. Instruments 2019, 90, 103201. [Google Scholar]
  40. Schmidt, J.; Weckesser, P.; Thielemann, F.; Schaetz, T.; Karpa, L. Optical Traps for Sympathetic Cooling of Ions with Ultracold Neutral Atoms. Phys. Rev. Lett. 2020, 124, 053402. [Google Scholar] [CrossRef]
  41. Härter, A.; Krükow, A.; Brunner, A.; Schnitzler, W.; Schmid, S.; Denschlag, J.H. Single Ion as a Three-Body Reaction Center in an Ultracold Atomic Gas. Phys. Rev. Lett. 2012, 109, 123201. [Google Scholar] [CrossRef]
  42. Jyothi, S.; Ray, T.; Dutta, S.; Allouche, A.R.; Vexiau, R.; Dulieu, O.; Rangwala, S.A. Photodissociation of Trapped Rb+2: Implications for Simultaneous Trapping of Atoms and Molecular Ions. Phys. Rev. Lett. 2016, 117, 213002. [Google Scholar] [CrossRef] [PubMed]
  43. da Silva, H., Jr.; Raoult, M.; Aymar, M.; Dulieu, O. Formation of molecular ions by radiative association of cold trapped atoms and ions. Mol. Phys. 2013, 111, 1683. [Google Scholar] [CrossRef]
  44. Alharzali, N.; Sardar, D.; Mlika, R.; Deb, B.; Berriche, H. Spectroscopic properties and cold elastic collisions of alkaline-earth Mg + Mg+ system. J. Phys. B At. Mol. Opt. Phys. 2018, 51, 195201. [Google Scholar] [CrossRef]
  45. Singer, K.; Poschinger, U.; Murphy, M.; Ivanov, P.; Ziesel, F.; Calarco, T.; Schmidt-Kaler, F. Trapped ions as quantum bits: Essential numerical tools. Rev. Mod. Phys. 2010, 82, 2609–2632. [Google Scholar] [CrossRef]
  46. Schneider, C.; Porras, D.; Schaetz, T. Experimental quantum simulations of many-body physics with trapped ions. Rep. Prog. Phys. 2012, 75, 024401. [Google Scholar] [CrossRef] [PubMed]
  47. Idziaszek, Z.; Simoni, A.; Calarco, T.; Julienne, P.S. Multichannel quantum-defect theory for ultracold atom–ion collisions. New J. Phys. 2011, 13, 083005. [Google Scholar] [CrossRef]
  48. Tomza, M.; Koch, C.P.; Moszynski, R. Quantum gas microscopy with spin, atom-number, and multilayer readout. Phys. Rev. A 2015, 91, 042706. [Google Scholar] [CrossRef]
  49. Farjallah, M.; Sardar, D.; El-Kork, N.; Deb, B.; Berriche, H. Electronic structure and photoassociation scheme of ultracold (MgK+) molecular ions. J. Phys. B At. Mol. Opt. Phys. 2018, 52, 045201. [Google Scholar] [CrossRef]
  50. Gacesa, M.; Montgomery, J.A.; Michels, H.H.; Côté, R. Controlling charge transfer in ultracold gases via Feshbach resonances. Phys. Rev. A 2016, 94, 013407. [Google Scholar] [CrossRef]
  51. Petrov, A.; Makrides, C.; Kotochigova, S.J. A new approach to molecular dynamics with non-adiabatic and spin-orbit effects with applications to QM/MM simulations of thiophene and selenophene. Chem. Phys. 2017, 146, 08430. [Google Scholar]
  52. Zrafi, W.; Ladjimi, H.; Said, H.; Berriche, H.; Tomza, M. Ab initio electronic structure and prospects for the formation of ultracold calcium–alkali-metal-atom molecular ions. New J. Phys. 2020, 22, 073015. [Google Scholar] [CrossRef]
  53. Ladjimi, H.; Farjallah, M.; Berriche, H. Spectroscopic, structural and lifetime calculations of the ground and low-lying excited states of the BeNa+, BeK+, and BeRb+ molecular ions. Phys. Scrip. 2020, 95, 055404. [Google Scholar] [CrossRef]
  54. Ladjimi, H.; Sardar, D.; Farjallah, M.; Alharzali, N.; Naskar, S.; Mlika, R.; Berriche, H.; Deb, B. Spectroscopic properties of the molecular ions BeX+ (X = Na, K, Rb): Forming cold molecular ions from an ion–atom mixture by stimulated Raman adiabatic process. Mol. Phys. 2018, 116, 1812–1826. [Google Scholar] [CrossRef]
  55. Ladjimi, H.; Zrafi, W.; Farjallah, M.; Bejaoui, M.; Berriche, H. Electronic structure, cold ion–atom elastic collision properties and possibility of laser cooling of BeCs+ molecular ion. Phys. Chem. Chem. Phys. 2022, 24, 18511–18522. [Google Scholar] [CrossRef] [PubMed]
  56. Ghanmi, C.; Farjallah, M.; Berriche, H. Theoretical study of the alkaline-earth (LiBe)+ ion: Structure, spectroscopy and dipole moments. J. Phys. B 2017, 50, 055101. [Google Scholar] [CrossRef]
  57. Farjallah, M.; Ghanmi, C.; Berriche, H. Structure and spectroscopic properties of the beryllium hydride ion BeH+: Potential energy curves, spectroscopic constants, vibrational levels and permanent dipole moments. Eur. Phys. J. D 2013, 67, 245. [Google Scholar] [CrossRef]
  58. Evangelisti, S.; Daudey, J.-P.; Malrieu, J.-P. Convergence of an improved CIPSI algorithm. Chem. Phys. 1983, 75, 91–102. [Google Scholar] [CrossRef]
  59. Durand, P.; Barthelat, J.C. A theoretical method to determine atomic pseudopotentials for electronic structure calculations of molecules and solids. Theor. Chim. Acta 1975, 38, 283–302. [Google Scholar] [CrossRef]
  60. Durand, P.; Barthelat, J.C. Pseudopotentials in atomic and molecular physics. Chem. Gazz. Chim. Ital. 1978, 108, 225–236. [Google Scholar]
  61. Barthelat, J.C.; Durand, P.; Serafini, A. Non-empirical pseudopotentials for molecular calculations. Mol. Phys. 1977, 33, 159–180. [Google Scholar] [CrossRef]
  62. Müller, W.; Flesch, J.; Meyer, W. Treatment of intershell 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, 3297–3310. [Google Scholar] [CrossRef]
  63. Foucrault, M.; Millié, P.; Daudey, J.P. Non-perturbative method for core–valence correlation in pseudopotential calculations: Application to the Rb2 and Cs2 molecules. J. Chem. Phys. 1992, 96, 1257–1264. [Google Scholar] [CrossRef]
  64. Wilson, J.N.; Curtis, R.M. Dipole polarizabilities of ions in alkali halide crystals. J. Phys. Chem. 1970, 74, 187–196. [Google Scholar] [CrossRef]
  65. Pavolini, D.; Gustavsson, T.; Spiegelmann, F.; Daudey, J.P. Theoretical study of the excited states of the heavier alkali dimers. I. The RbCs molecule. J. Phys. B Mol. Opt. 1989, 77, 221721. [Google Scholar] [CrossRef]
  66. Ghanmi, C.; Berriche, H.; Ben Ouada, H. Evaluation of the Atomic Polarisabities of Li and K using the Long Range Lik+ Electronic States Behaviour. In Proceedings of the International Conference on Computational and Mathematical Methods in Science and Engineering, CMMSE, Alicante, Spain, 27–30 June 2005; pp. 166–174. [Google Scholar]
  67. Bouzouita, H.; Ghanmi, C.; Berriche, H. Ab initio study of the alkali-dimer cation Li2+. J. Mol. Struct. Theochem. 2006, 777, 75–80. [Google Scholar] [CrossRef]
  68. Moore, C.E.; Mack, J.E. Atomic energy levels. At. Energy Levels Deriv. Anal. Opt. Spectra. Phys. Today 1952, 5, 23. [Google Scholar] [CrossRef]
  69. Mitroy, J.; Zhang, J.Y. Long range interactions of the Mg+ and Ca+ ions. Eur. Phys. J. D 2007, 46, 415–424. [Google Scholar] [CrossRef]
  70. Moore, C.E. Atomic Energy Levels; NSRDS-NBS No. 467; US Government Printing Office: Washington, DC, USA, 1989. [Google Scholar]
  71. Zemke, W.T.; Crooks, J.B.; Stwalley, W.C. Radiative and nonradiative lifetimes for vibrational levels of the A 1Σ+ state of 7LiH. J. Chem. Phys. 1978, 68, 4628. [Google Scholar] [CrossRef]
  72. Partridge, H.; Langhoff, S.R. Theoretical treatment of the X 1Σ+, A 1Σ+, and B 1Π states of LiH. J. Chem. Phys. 1981, 74, 2361. [Google Scholar] [CrossRef]
  73. Berriche, H.; Gadéa, F.X. Transition dipole function and radiative lifetimes for the A and C 1Σ+ states of the LiH molecule. Eur. Phys. J. D 2016, 70, 1–9. [Google Scholar] [CrossRef]
  74. Berriche, H. Ab Initio Adiabatic and Diabatic Permanent Dipoles for the Low-Lying States of the LiH Molecule. Ph.D. Thesis, Paul Sabatier University, Toulouse, France, 1995. [Google Scholar]
  75. Stolyarov, A.V.; Pupyshev, V.I. Approximate sum rule for diatomic vibronic states. Phys. Rev. A 1994, 49, 1693. [Google Scholar] [CrossRef]
  76. Pazyuk, E.; Stolyarov, A.; Pupyshev, V. Approximate sum rule for diatomic vibronic states as a tool for the evaluation of molecular properties. Chem. Phys. Lett. 1994, 228, 219–224. [Google Scholar] [CrossRef]
  77. Johnson, R.B. New numerical methods applied to solving the one-dimensional eigenvalue problem. J. Chem. Phys. 1977, 67, 4086. [Google Scholar] [CrossRef]
  78. Napolitano, R.; Weiner, J.; Williams, C.J.; Julienne, P.S. Line Shapes of High Resolution Photoassociation Spectra of Optically Cooled Atoms. Phys. Rev. Lett. 1994, 73, 1352–1355. [Google Scholar] [CrossRef] [PubMed]
  79. Rakshit, A.; Deb, B. Formation of cold molecular ions by radiative processes in cold ion-atom collisions. Phys. Rev. A 2011, 83, 022703. [Google Scholar] [CrossRef]
  80. Sardar, D.; Naskar, S.; Pal, A.; Berriche, H.; Deb, B. Formation of a molecular ion by photoassociative Raman processes. J. Phys. B At. Mol. Opt. Phys. 2016, 49, 245202. [Google Scholar] [CrossRef]
  81. Sardar, D.; Naskar, S. Cold collisions between alkali metals and alkaline-earth metals in the heteronuclear atom-ion system Li + Ba+. Phys. Rev. A 2023, 107, 043323. [Google Scholar] [CrossRef]
Figure 1. Adiabatic potential energy curves of the first 101Σ+ electronic states of (MgCs)+.
Figure 1. Adiabatic potential energy curves of the first 101Σ+ electronic states of (MgCs)+.
Atoms 11 00121 g001
Figure 2. Adiabatic potential energy curves of the first 103Σ+ electronic states of (MgCs)+.
Figure 2. Adiabatic potential energy curves of the first 103Σ+ electronic states of (MgCs)+.
Atoms 11 00121 g002
Figure 3. Vibrational energy level spacing (E(υ+1)–E(υ)) (in cm−1) for the (1,3Σ+) electronic states of the (MgCs)+ molecule.
Figure 3. Vibrational energy level spacing (E(υ+1)–E(υ)) (in cm−1) for the (1,3Σ+) electronic states of the (MgCs)+ molecule.
Atoms 11 00121 g003
Figure 4. (a) Permanent dipole moment for the 1Σ+ states of the (MgCs)+ molecular ion. (b) Zoomed area at distances between 0 and 15 a.u.
Figure 4. (a) Permanent dipole moment for the 1Σ+ states of the (MgCs)+ molecular ion. (b) Zoomed area at distances between 0 and 15 a.u.
Atoms 11 00121 g004
Figure 5. (a) Permanent dipole moment for the 3Σ+ states of the (MgCs)+ molecular ion. (b) Zoomed area at distance between 0 and 15 a.u.
Figure 5. (a) Permanent dipole moment for the 3Σ+ states of the (MgCs)+ molecular ion. (b) Zoomed area at distance between 0 and 15 a.u.
Atoms 11 00121 g005
Figure 6. Transition dipole moments for selected (1–>2, 2–>3, and 3–>4) 1,3Σ+ states as a function of the internuclear distance R.
Figure 6. Transition dipole moments for selected (1–>2, 2–>3, and 3–>4) 1,3Σ+ states as a function of the internuclear distance R.
Atoms 11 00121 g006
Figure 7. Logarithm of total elastic scattering cross-sections (σtot) in a.u. as a function of logarithm energy E of the (MgCs)+ molecular ion.
Figure 7. Logarithm of total elastic scattering cross-sections (σtot) in a.u. as a function of logarithm energy E of the (MgCs)+ molecular ion.
Atoms 11 00121 g007
Figure 8. The rate constant KPA (cm3s-1) of PA as a function of energy E (in Kelvin) for laser frequency 10 W/cm2 tuned at PA resonance.
Figure 8. The rate constant KPA (cm3s-1) of PA as a function of energy E (in Kelvin) for laser frequency 10 W/cm2 tuned at PA resonance.
Atoms 11 00121 g008
Figure 9. The variation of the rate constant of photoassociation as a function of the intensity of the applied laser.
Figure 9. The variation of the rate constant of photoassociation as a function of the intensity of the applied laser.
Atoms 11 00121 g009
Table 1. Polarizabilities, cut-off radii, and pseudo-potential parameter (in a.u.) atoms.
Table 1. Polarizabilities, cut-off radii, and pseudo-potential parameter (in a.u.) atoms.
Atomαρsρpρd
Mg(Z = 12)0.469040.91.254991.500
Cs(Z = 55)15.1172.691.852.810
Table 3. Spectroscopic constants of the ground and excited singlet and triplet electronic states of (MgCs)+ molecule.
Table 3. Spectroscopic constants of the ground and excited singlet and triplet electronic states of (MgCs)+ molecule.
StateRe (a.u.)De (cm−1)Te (cm−1)ωe (cm−1)ωexe (cm−1)Be (cm−1)
X1Σ+7.702047073.20.630.049364
7.85 [28]1861 [28] 73.2 [28]0.0481 [28]
21Σ+11.45266229,78244.640.250.022354
31Σ+16.88105436,39121.920.160.010280
41Σ+9.33−2750
19.46189142,10121.730.060.007700
51Σ+28.2426345,2569.120.070.003674
61Σ+58.89446,990−0.4814.610.001160
71Σ+8.68−3301
27.8350948,5078.9226.510.038919
81Σ+8.68−8647
31.9992350,05910.1329.950.038864
91Σ+8.69−10,460
42.7530051,2405.4430.840.038839
101Σ+8.71−10,297
42.27110553,2898.0740.100.038614
13Σ+8.58450419,55674.780.520.039826
23Σ+14.9881331,63025.560.170.013050
33Σ+17.20265840,54226.850.050.009905
43Σ+28.1550643,48610.120.040.003698
53Σ+58.89446,990−0.4814.610.001160
63Σ+28.13160048,41313.190.020.003704
73Σ+38.203150,6676.810.030.002008
83Σ+49.3213150,8504.590.020.001205
93Σ+40.09105552,9678.790.010.001823
103Σ+56.7639553,9993.950.010.000910
Table 4. Radiative lifetimes and vibrational-level spacing E ν E v 1 of the 21Σ+ state of the (MgCs)+ ion.
Table 4. Radiative lifetimes and vibrational-level spacing E ν E v 1 of the 21Σ+ state of the (MgCs)+ ion.
Vibrational LevelEv − Ev−1 (cm−1)Franck–Condon (ns)Sum Rule (ns)
0 10.57910.582
144.11010.65810.661
243.87810.73910.742
343.59910.82510.828
443.36710.91210.915
543.08011.00311.006
642.84811.09611.098
742.55011.19111.194
842.27811.29011.294
942.01611.39411.399
1041.74411.50211.511
1141.47711.63411.666
1241.21311.77011.824
1340.94112.05512.253
1440.66612.29412.589
1540.38912.98913.787
1640.10913.54914.649
1739.83214.48616.450
1839.53415.30117.708
1939.23715.68918.283
2038.92615.96718.465
2138.60416.25918.689
2238.29317.04219.857
2337.99217.53220.454
2437.66818.32121.644
2537.35218.66021.795
2637.04119.14222.251
2736.70719.72822.916
2836.38120.50423.969
2936.04421.15824.555
3035.69921.57424.792
3135.36122.44625.859
3235.01223.07526.408
3334.66623.81627.190
3434.31224.27727.467
3533.95925.21728.468
3633.59926.40029.687
3733.23526.97630.140
3832.86827.66830.647
3932.49328.38631.252
4032.11429.68932.700
4131.72630.55433.375
4231.33331.44634.124
4330.93232.93935.617
4430.52633.86436.390
4530.11834.80737.156
4629.70535.79637.999
4729.28737.17539.363
4828.86338.66440.724
4928.43539.78841.692
5028.00041.62543.490
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Farjallah, M.; Sardar, D.; Deb, B.; Berriche, H. Electronic Structure, Spectroscopy, Cold Ion–Atom Elastic Collision Properties, and Photoassociation Formation Prediction of the (MgCs)+ Molecular Ion. Atoms 2023, 11, 121. https://doi.org/10.3390/atoms11090121

AMA Style

Farjallah M, Sardar D, Deb B, Berriche H. Electronic Structure, Spectroscopy, Cold Ion–Atom Elastic Collision Properties, and Photoassociation Formation Prediction of the (MgCs)+ Molecular Ion. Atoms. 2023; 11(9):121. https://doi.org/10.3390/atoms11090121

Chicago/Turabian Style

Farjallah, Mohamed, Dibyendu Sardar, Bimalendu Deb, and Hamid Berriche. 2023. "Electronic Structure, Spectroscopy, Cold Ion–Atom Elastic Collision Properties, and Photoassociation Formation Prediction of the (MgCs)+ Molecular Ion" Atoms 11, no. 9: 121. https://doi.org/10.3390/atoms11090121

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop