Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Fine structure investigation and laser cooling study of the LaH molecule

Nariman Abu El Khera, Nayla El-Kork*ab, Nissrin Alharzali a, Joumana Assafcd and Mahmoud Koreke
aDepartment of Physics, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates. E-mail: nayla.elkork@ku.ac.ae
bSpace and Planetary Science Center, Khalifa University, Abu Dhabi, United Arab Emirates
cDoctoral School of Sciences and Technology, Lebanese University, Hadath, Lebanon
dCenter for Educational Research and Development, CERD, Sin El Fil, Lebanon
eFaculty of Science, Beirut Arab University, Riad El Solh, Beirut 1107 2809, Lebanon

Received 17th April 2025 , Accepted 8th August 2025

First published on 29th October 2025


Abstract

A theoretical feasibility study of the spin–orbit laser cooling of the molecule LaH has been performed based on a complete active space self-consistent field (CASSCF)/MRCI ab initio calculation with Davidson correction in the Λ(±) and Ω(±) representations. The adiabatic potential energy curves and spectroscopic constants have been investigated for the considered electronic states. The small value of the equilibrium positions difference ΔRe between the ground and the electronic states X3Σ0+, (1)3Π0+, and (1)3Δ1 predicts the candidacy of the molecule LaH for direct laser cooling between the first two states with the intermediate state (1)3Δ1. The calculation of the diagonal Franck–Condon factors, the short radiative lifetime, and the experimental parameters (slowing distance, Doppler and recoil temperature, …) suggest that the molecule LaH is a good candidate for Doppler laser cooling, and a corresponding laser cooling scheme is presented.


1. Introduction

The development of techniques for producing, trapping, and controlling ultracold molecules in the gas phase has marked a major milestone in physics, attracting significant research interest over the years. Ultracold molecules offer a wide range of potential applications, including precision measurements,1,2 simulation of solid-state systems,3,4 quantum information processing,5 and the control of chemical reactions at ultracold temperatures.6 Numerous studies have explored a range of diatomic polar molecules and ions, including hydrogen, as potential candidates for laser cooling. These include alkaline-earth-metal monohydrides7 (such as BeH, MgH, CaH, SrH, and BaH), transition metal hydrides (such as AgH,8 CuH,9 and AuH10), and other molecular systems (like CH,11 AlH and AlF,12 BH+ and AlH+13). Our recent work reported HfH14 as a highly promising candidate for experimental laser cooling.

In general, the study of the electronic structure of molecules with open d and f orbitals (partial occupation), such as transition metals hydrides and Lanthanides hydrides, presents a major challenge for both theorists and experimentalists due to the significant electron–degeneracy correlations15–17 involved. The formation of chemical bonds arising from d-electrons poses a significant difficulty for theorists, as accurately modeling these molecules requires considering relativistic effects and spin–orbit coupling.18 Besides their importance in theoretical chemistry, the group of lanthanides hydrides such as LaH plays a critical role in various fields, such as astrophysics (since hydrogen is the most abundant element in the universe, LaH is found in the spectra of sunspots and cool stars),19 catalysis,20 organometallic chemistry,21 and electron's electric dipole moment (EDM) measurements.22 The electronic structure of the LaH molecule has been experimentally examined in the literature,23–28 with previous theoretical studies provided in ref. 29–31. Recently, Assaf et al.32 conducted a comprehensive theoretical study of the LaH molecule in both Λ(±) and Ω(±) representations. Nevertheless, the study of the laser-cooling candidacy of the LaH molecule has never been explored.

This paper presents a spin–orbit coupling theoretical calculation and laser cooling investigation of the LaH molecule based on Assaf et al.'s previous work.32 Additionally, we investigated the diagonal Frank-Condon factors (FCFs) and the radiative lifetimes (τ) for the two transitions X3Σ0+–(1)3Π0+ and (1)3Δ1–(1)3Π0+, where (1)3Δ1 is an intermediate state between X3Σ0+ and (1)3Π0+ states. The branching ratios of the vibrational transitions Rνν have been calculated along with the number of cycles (N) for photon absorption/emission and the slowing distance L, which falls within the practical experimental limits. A laser cooling scheme with an intermediate state is presented.

2. Computational approach

In this paper, we have employed the state-averaged Complete Active Space Self-Consistent Field (CASSCF) method, followed by Multi-Reference Configuration Interaction (MRCI) calculations with Davidson correction (+Q) to study the electronic structure of the LaH molecule. The high-level ab initio computations were conducted with the MOLPRO33 software, utilizing the GABEDIT34 graphical user interface. Due to the large number of electrons in the lanthanum atom (57La), selecting appropriate basis sets that accurately describe its electronic structure is challenging. Given that this study focuses on the laser cooling of the LaH molecule, we are particularly interested in its ground state and the first two excited states (1)3Δ and (1)3Π. To ensure high-precision calculations while accounting for spin–orbit coupling effects, we examined various basis sets based on previously published data.

In 2014, Mahmoud and Korek31 reported theoretical calculations on the low-lying electronic states of the LaH molecule, both with and without spin–orbit coupling, using the CASSCF/MRCI method. They employed the SBKJC-VDZ (ECP46MHF) valence double-zeta35 basis set for lanthanum, which incorporates a relativistic effective core potential, and the augmented correlation-consistent polarized valence quadruple-zeta (aug-cc-pVQZ) basis set for hydrogen.36 With the 12 electrons explicitly considered for the LaH molecule in the C2v symmetry, the authors performed the calculations with different valence electrons (2, 6, 8, and 10 valence electrons) to check their influence on the values of the transition energy with respect to the ground state minimum (Te). Their findings indicated that when two or six valence electrons were included, the (1)3Δ state was lower in energy than the (1)3Π state, consistent with the experimentally observed order of states.25 However, for a higher number of valence electrons, the (1)3Π state became lower in energy than the (1)3Δ state. Recently, Assaf et al.32 determined the spectroscopic constants of the LaH molecule using the quasi-relativistic effective core potential (ECP28-MWB) basis set37,38 for lanthanum. In this approach, lanthanum is described by replacing its 28 inner-core electrons with the effective core potential, while the remaining 29 electrons are explicitly represented by the ANO Gaussian basis set with a contraction scheme of (14s, 13p, 10d, 8f, 6g)/[6s, 6p, 5d, 4f, 3g].37,38 The hydrogen single electron is treated with the augmented correlation-consistent polarized valence quadruple-zeta (aug-cc-pVQZ) basis set,36 contracted as (7s, 4p, 3d, 2f)/(5s, 4p, 3d, 2f). The authors calculated the low-lying excited states in both Λ(±) and Ω(±) representations and determined the corresponding spectroscopic constants. Their results showed strong agreement with experimental data. Motivated by this, we employ the same basis sets (ECP28-MWB in conjunction with ANO Gaussian basis set37,38 for La and aug-cc-pVQZ36 for H) as well as the all-electron approach described by Assaf et al.32 to carry out our own calculations, both with and without spin–orbit coupling (S.O.C) effects, aiming for a more accurate analysis relevant to laser cooling of the LaH molecule.

To assess the reliability of the employed pseudopotentials and basis sets, we have performed the calculations of our study using a benchmark of basis sets B1–B5, along with literature values, all compared with experimental data as shown in Table 1. The B1 set employs the pseudopotential ECP28MWB37,38 for lanthanum with a (s, p, d, f, g) ANO Gaussian basis set,37,38 combined with the aug-cc-pVQZ36 (s, p, d, f) basis for hydrogen. The B2 set retains the same basis for La as in B1 but uses a higher-level aug-cc-pV5Z39 (s, p, d, f) basis for H. The B3 set again uses ECP28MWB37,38 for La but combines it with the smaller aug-cc-pVTZ39 (s, p, d) basis for H. In contrast, B4 and B5 utilize the ECP46MWB35,40 pseudopotential for La, where Lanthanum is described as a system of 46 inner electrons, and the remaining 11 electrons are represented by the corresponding basis set ECP46MWB-II ((6s6p5d)/[4s4p4d] + 2s1p1d).35,41,42 The active space of C2v point group symmetry contains 5σ (La: 5d+2, 6p0, 5d0; H: 1 s, 2 s), 2π (La: 5d+1, 6p+1; H: 0), 1δ (La: 5d−2; H: 0) molecular orbitals and are distributed into the irreducible representation as 5A1, 2B1, 2B2, and 1A2, denoted by [5, 2, 2, 1]. A CASSCF calculation was performed with two valence electrons from LaH distributed over the ten active orbitals. B4 combines ECP46MWB35,40 with aug-cc-pV5Z39 (s, p, d, f) for H, while B5 pairs it with aug-cc-pVQZ36 (s, p, d, f) for H. Across all considered electronic states in the Λ(±) representation X1Σ+, (1)3Δ, and (1)3Π, this benchmarking enables a detailed assessment of the sensitivity of spectroscopic constants: the equilibrium bond length Re, the transition energy with respect to the ground state minimum Te, the harmonic frequency ωe, and the anharmonicity constant ωexe to the choice of basis sets and pseudopotentials, as presented in Table 1.

Table 1 The spectroscopic constants of X1Σ+, (1)3Δ, and (1)3П states of LaH molecule, without spin–orbit coupling using a benchmark of basis sets B1–B5, along with literature values, all compared with experimental data
States Ref. Te (cm−1) Re (Å) Re| (Å) ωe (cm−1) ωe| (cm−1) ωexe (cm−1) ωexe| (cm−1)
a Energy corresponding to ν00.b Estimated energy of 3Λ(±) state determined by calculating the average of the spin–orbit components' energy. B1 This work using the ECP28MWB (s, p, d, f, g) basis set37,38 for lanthanum and aug-cc-pVQZ (s, p, d, f)36 for hydrogen. B2 This work using the ECP28MWB (s, p, d, f, g) basis set37,38 for lanthanum and aug-cc-pV5Z (s, p, d, f)39 for hydrogen. B3 This work using the ECP28MWB (s, p, d, f, g) basis set37,38 for lanthanum and aug-cc-pVTZ (s, p, d)39 for hydrogen. B4 This work using the ECP46MWB (s, p, d, f, g) basis set35,40 for lanthanum and aug-cc-pV5Z (s, p, d, f)39 for hydrogen. B5 This work using the ECP46MWB (s, p, d, f, g) basis set35,40 for lanthanum and aug-cc-pVQZ (s, p, d, f)36 for hydrogen. Theoretical work31 used the ECP46MHF (s, p, d) basis set35 for lanthanum and aug-cc-pVQZ (s, p, d, f)36 for hydrogen.
X1Σ+ Exp.24 0.0 2.032      
Exp.25 0.0   1418   15.6  
This workB1 0.0 2.024 0.008 1447.28 29.28 15.43 0.17
This workB2 0.0 2.024 0.008 1447.49 29.29 13.81 1.79
This workB3 0.0 2.026 0.006 1468.69 50.69 28.79 13.19
This workB4 0.0 2.078 0.046 1433.43 15.43 15.74 0.14
This workB5 0.0 2.078 0.046 1437.38 19.38 15.45 0.15
Theo.29 0.0 2.08 0.048 1433 15
Theo.30 0.0 2.060 0.028 1429 11 20.93 5.33
Theo.31 0.0 2.235 0.203 1353.26 64.74
Theo.32 0.0 2.025 0.007 1439.77 21.77 15.242 0.358
(1)3Δ Exp.25   1355   14.4  
This workB1 2179.37 2.081   1347.34 7.66 15.22 0.82
This workB2 1962.36 2.081   1371.59 16.59 15.89 1.49
This workB3 2052.86 2.086   1357.44 2.44 11.02 3.38
This workB4 2043.29 2.143   1338.96 16.04 14.49 0.09
This workB5 2022.45 2.142   1335.94 19.06 11.03 3.37
Theo.29 2805 2.13   1352 3.00
Theo.31 3916 2.272   1314.98 40.02
Theo.32 2232 2.082   1371.65 16.65 15.16 0.76
(1)3П Exp.25 3732a,b      
This workB1 4222.7 2.065   1358.32   18.39  
This workB2 4222.74 2.064   1354.91   17.86  
This workB3 4261.98 2.074   1343.06   18.05  
This workB4 4764.40 2.145   1337.96   16.97  
This workB5 4753.36 2.145   1314.65   17.75  
Theo.29 5147 2.12   1341    
Theo.31 3880 2.235   1341.37    
Theo.32 4263 2.066   1359.20   18.65  


In terms of effective core potential, and for the ground state X1Σ+, B1 shows excellent agreement in the equilibrium bond length (Re = 2.024 Å), closely matching the experimental value of 2.032 Å with only a 0.008 Å deviation. Case B5, however, displays a much more important difference of 0.046 Å. At the same time, the vibrational constant ωe from B5 (1437.38 cm−1) is slightly closer to the experimental value (1418 cm−1) than that from B1 (1447.28 cm−1). For the excited state (1)3Δ, the ωe and ωexe values from B1 align well with experimental values, whereas B5 significantly underestimates this anharmonicity constant. Finally, for the (1)3Π state, B1 yields excitation with an energy Te, which is more consistent with available measurements.

These comparisons demonstrate that pseudopotential B1 is more accurate and reliable for describing the spectroscopic properties of LaH, and it is therefore preferred in our study.

In terms of basis sets, B3 yields a slightly improved vibrational constant ωe (1468.69 cm−1) compared to experiment (1418 cm−1); however, it also leads to significantly higher deviations in ωexe. For the (1)3Δ state, B1 provides an excitation energy Te = 2179.37 cm−1 and Re = 2.081 Å, which are closer to experiment than those from B4 or B5, which overestimate Re and show larger deviations in vibrational constants. Concerning the (1)3Π state, B1 again delivers consistent performance, with Te = 4222.7 cm−1 and ωe = 1358.32 cm−1, reasonably close to the experimental values, while B4 and B5 significantly overestimate the excitation energy and distort vibrational constants.

Overall, B1 exhibits the most balanced and accurate agreement across all three electronic states compared to experimental data. Therefore, the B1 basis set (ECP28-MWB for La and aug-cc-pVQZ for H) is validated as the most suitable choice for describing the spectroscopic properties of LaH molecule.

3. Ab initio results

Our primary focus in this work is the study of the laser cooling of LaH, by considering transitions between the ground and low-lying excited states. The investigated potential energy curves for the three low-lying Λ(±) and Ω(±) states: X1Σ+ (X1Σ0+), (1)3Δ {(1)3Δ1, (1)3Δ2, (1)3Δ3}, and (1)3Π {(1)3Π0−, (1)3Π0+, (1)3Π1, (1)3Π2} are displayed as a function of internuclear separation in Fig. 1 and Fig. 2. The spin–orbit coupling splitting of the electronic states (1)3Δ and (1)3Π of the molecule LaH is illustrated in Fig. 3. Due to the dominance of fine structure in the spectra of heavy molecules such as LaH, it is crucial to account for the spin–orbit interaction of lanthanum. This requirement is validated by the notably large splitting energies observed for the 3Δ state (approximately 448 cm−1 for (1)3Δ2 – (1)3Δ1) and 3Π state (around 320 cm−1 for (1)3Π2 – (1)3Π1) electronic states, as illustrated in Fig. 2 and 3. These results highlight the significant influence of spin–orbit coupling on the electronic states of the LaH molecule. In addition, the transition dipole moment curves (TDMCs) of the allowed transitions X1Σ0+–(1)3Π0+ and (1)3Π0+–(1)3Δ1 of the molecule LaH have been computed as a function of internuclear distance and displayed in Fig. 4.
image file: d5ra02708j-f1.tif
Fig. 1 The potential energy curves of the ground and first two low-lying excited states of the LaH molecule, in the Λ(±) representation.

image file: d5ra02708j-f2.tif
Fig. 2 The potential energy curves of the ground and first two low-lying excited states of the LaH molecule, in the Ω(±) representation.

image file: d5ra02708j-f3.tif
Fig. 3 Spin–orbit coupling splitting of the electronic states (1)3Δ and (1)3Π of the molecule LaH.

image file: d5ra02708j-f4.tif
Fig. 4 Transition dipole moments of the transitions (a) X1Σ0+–(1)3Π0+ and (b) (1)3Δ1–(1)3Π0+ of the molecule LaH.

The spectroscopic constants such as the equilibrium bond length Re, the transition energy with respect to the ground state minimum Te, the harmonic frequency ωe, and the anharmonicity constant ωexe of all the calculated Λ(±) (listed as (B1) in Table 1) and Ω(±) states are determined and listed in Table 2. As previously mentioned, the ground state has 1Σ+ symmetry, and the first two low-lying excited states are (1)3Δ and (1)3Π. Their spectroscopic constants strongly agree with experimental data.24,25 In the Λ(±) representation, the equilibrium internuclear distance Re of the ground state exhibits a relative error of only 0.4% compared to the experimental value (Re = 2.032 Å).24 Similarly, the vibrational constants ωe and ωexe for (1)3Δ state show relative errors of 0.6% and 5.7%, respectively, compared to experimental data.25 The ab initio investigation of LaH conducted in 2014[thin space (1/6-em)]31 showed that the use of the large-core pseudopotential ECP46MHF35 for lanthanum (La) with 10 valence electrons introduced considerable inaccuracies in the computed equilibrium bond lengths (Re) of various electronic states, and the transition energies (Te) for several predicted states were significantly overestimated. For instance, the ground state Re deviated by approximately 9.4% from the experimental value.24

Table 2 The spectroscopic constants of X1Σ+, (1)3Δ, and (1)3П states with and without spin–orbit coupling of LaH
Spectroscopic constants in the Λ(±) representation
States Ref. Te (cm−1) ΔTe/Te% Re (Å) ΔRe/Re% ωe (cm−1) Δωe/ωe% ωexe (cm−1) Δωexe/ωexe%
a Energy corresponding to ν00.b Estimated energy of 3Λ(±) state determined by calculating the average of the spin–orbit components' energy.
X1Σ+ This work 0.0   2.024 0.4 1447.28 15.43
Exp.24 0.0   2.032 2.1 1.1
Exp.25 0.0   2.7 1418 1.0 15.6
Theo.29 0.0   2.08 1.7 1433 1.3 26.3
Theo.30 0.0   2.060 9.4 1429 6.9 20.93
Theo.31 0.0   2.235 0.0 1353.26 0.5 1.2
Theo.32 0.0   2.025   1439.77   15.242  
(1)3Δ This work 2179.37 2.081 1347.34 0.6 15.22 5.7
Exp.25 22.3 2.3 1355 0.3 14.4
Theo.29 2805 44.3 2.13 8.4 1352 2.5
Theo.31 3916 2.4 2.272 0.0 1314.98 1.8 0.4
Theo.32 2232   2.082   1371.65   15.16  
(1)3П This work 4222.7 2.065 1358.32 18.39
Exp.25 3732a,b 18.0 2.6 1.3
Theo.29 5147 8.8 2.12 7.6 1341 1.3
Theo.31 3880 0.9 2.235 0.0 1341.37 0.1 1.4
Theo.32 4263   2.066   1359.20   18.65  

Spectroscopic constants in the Ω(±) representation
States Ref. Te (cm−1) ΔTe/Te% Re (Å) ΔRe/Re% ωe (cm−1) Δωe/ωe% ωexe (cm−1) Δωexe/ωexe%
X1Σ+0+ This work 0.0   2.027 1443.63 1.8 14.56 6.7
Exp.25 0.0   0.2 1418 15.6
Theo.29 0.0   2.0319 0.0 0.1 8.5
Theo.32 0.0   2.027   1444.66   15.904  
(1)3Δ1 This work 1664.4 2.080 1.4 1359.66 13.67
Exp.23 2.1102 0.3 5.1
Exp.25 1259.5a 0.9 1355 14.4
Theo.29 0.0 2.099 0.2 0.3 4.6
Theo.32 1665   2.085   1363.659   14.331  
(1)3Δ2 This work 2112.1 2.078 0.8 1363.87 14.79
Exp.23 2.0938
Exp.25 1646a 0.2
Theo.29 0.1 2.083 0.3 0.2 2.4
Theo.32 2111   2.084   1366.582   15.15  
(1)3Δ3 This work 2682.8 2.085   1371.49   17.81
Exp.23 2.0925 0.4
Theo.29 0.0 2.081 0.2 0.1 0.8
Theo.32 2684   2.082 0.1 1373.275   17.960  
(1)3П0− This work 4011.9 2.0637 1356.63 15.656
Exp.25 3542a 0.1 0.2 0.3 2.0
Theo.32 4014   2.068   1352.946   15.977  
(1)3П0+ This work 4043.9 2.0619 1361.47 17.899
Exp.25 3586a 0.1 0.2 0.2 2.4
Theo.32 4047   2.067   1358.629   18.345  
(1)3П1 This work 4239.7 2.0646 1352.34 16.478
Exp.25 3754a 0.0 0.1 0.1 1.1
Theo.32 4241   2.067   1353.606   16.669  
(1)3П2 This work 4559.7 2.0639  1369.78 15.975
Exp.27 4048a 0.1 0.1 0.1 3.6
Theo.32 4557   2.065   1367.963   16.579  


The transition energies (Te) associated with the spin–orbit components of the (1)3Δ and (1)3Π electronic states were not reported in the published experimental studies.25 Instead, it only provided the energies corresponding to ν00, i.e., the T0 values for these spin–orbit components. As a result, a direct comparison between our calculated Te values and the experimental T0 data is not feasible. For a meaningful comparison, we instead consider the spin–orbit splitting energies. However, the calculated splitting energy of the (1)3Δ state is ΔΩ1–2 (448 cm−1), which is closer to the experimental value25Ω1–2 = 387 cm−1). The calculated spin–orbit splitting for the (1)3Π state demonstrates a remarkably strong agreement with experimental observations reported in references.25,27 The total splitting energy obtained from our calculations, ΔETotal = 548 cm−1, aligns closely with the experimental value25,27 of 506 cm−1, differing by only ∼7%. This level of agreement underscores the accuracy and reliability of the theoretical methods employed, particularly in capturing the fine-structure effects arising from spin–orbit coupling. Such consistency between theory and experiment validates the computational treatment of the (1)3Π state, which plays a key role in our analysis of electronic transitions and laser cooling feasibility.

4. Laser cooling study of the LaH molecule

Laser cooling is a technique that reduces the motion of atoms or molecules by repeatedly scattering photons through fast and controlled optical transitions.43 Each photon scatter imparts a small, directional momentum change, effectively reducing the system's kinetic energy and entropy. Although both direct44–47 methods (e.g., buffer gas cooling, Stark deceleration) and indirect48,49 methods (e.g., photoassociation of cold atoms) can be used to achieve molecular cooling and trapping. Laser cooling has uniquely succeeded in reaching the sub-millikelvin range for various diatomic50–53 and linear triatomic54–56 species. Several criteria are to be followed when considering cooling transitions among vibronic levels in a diatomic molecule.

(i) A highly diagonal Franck–Condon array for the considered band system, which would ensure a low number of lasers that would be used to retain the closed-loop cycle of the molecule. One could usually recognize band systems with high Franck–Condon arrays when the equilibrium internuclear distance (Re) among the considered electronic states is minimal.57

(ii) No intervening electronic state that would disturb the laser cooling cycle. One should make a distinction in this case between intervening and non-intervening electronic states. An intervening electronic state is usually an intermediate state situated between the excited state and the ground state, forming the cycling loop, and that intervenes with the transition band. This usually takes place if there is a high probability of transition between the upper-level electronic state and this intermediate state. A lower transition probability would render the intermediate state as non-intervening. Recent studies, however, have shown the possible involvement of intervening intermediate states in the cooling process.58–60

(iii) The transition radiative lifetime among the considered vibrational levels should be very short to ensure high photon scattering rates. Usually, the considered radiative lifetimes are in the range of ns-ms.14,61,62

The equilibrium positions difference ΔRe between the ground state X1Σ0+ and the excited state (1)3Π0+ (about 0.0349 Å)[thin space (1/6-em)]of the molecule LaH is minimal. This encouraged the authors to consider a closed cycle formed from bands within these states. Fig. (5a) shows highly diagonal Franck–Condon arrays for the transitions X1Σ0+–(1)3Π0+ for the vibrational levels 0 ≤ v ≤ 5, obtained using the Level 11 program,63 thus fulfilling criteria (i).


image file: d5ra02708j-f5.tif
Fig. 5 The Frank–Condon factor of the transitions X1Σ0+–(1)3Π0 and (1)3Δ1–(1)3Π0 of the molecule LaH.

Fig. 2 shows four intermediate states to the considered cycle: (1)3Π0−, (1)3Δ1, (1)3Δ2, and (1)3Δ3. Transition among states (1)3Π0+ and (1)3Π0− are not allowed due to the rule 0+ ↛ 0[thin space (1/6-em)]64 in Hund's case-c. The transitions (1)3Π0+ – (1)3Δ2 and (1)3Π0+ – (1)3Δ2 are not allowed either, since transitions with ΔΩ > 1 are also forbidden64 in Hund's case c. As a consequence, the laser cooling of the LaH molecule through the cycle made of the transitions X1Σ0+–(1)3Π0+ will necessitate investigating the intervening degree of the intermediate state (1)3Δ1 only, as required through criteria (ii). The diagonality of the FCF for (1)3Δ1–(1)3Π0+ is shown in Fig. (5b), showing a high transition probability to the first vibrational levels of the intermediate state.

Criteria (iiii) can be evaluated by considering the Transition Dipole Moment curves (TDMCs), i.e., the transition dipole moment variation μ(R) in terms of the internuclear distance R among the involved electronic states. The TDMC for X1Σ0+–(1)3Π0+ and (1)3Δ1–(1)3Π0+ transitions are given in Fig. 5, as obtained with the Molpro program.33 The vibrational radiative lifetime τvv can be calculated as the inverse of the Einstein coefficient Avv image file: d5ra02708j-t1.tif.65,66 The vibrational Einstein Coefficient among the transition (1)3Δ1–(1)3Π0+ is calculated using the LEVEL 11 program according to the following formula:

 
Avv = (3.1361891)(10−7)(ΔE)3(〈ψν′|μ(r)|ψν〉)2 (1)
where Aνν has as units s−1, ΔE is the emission frequency (in cm−1) and μ(r) is the electronic transition dipole moment between the two considered electronic states (in Debye).67 For transitions of the nature ΣΠ, such as X1Σ0+–(1)3Π0+, the transition dipole moment (TDM) is vertical, as calculated in terms of μx, μy, μz. In this case, the Einstein coefficient expressed in (1) has to be divided by two.68

The vibrational branching loss ratio measures how much an intermediate state affects a laser cooling cycle between two other states. In our study, we have to examine how the intermediate state (1)3Δ1 influences the cycle between the X1Σ0+ and (1)3Π0+ states. The vibrational branching loss ratio to this state is approximately equal to:

image file: d5ra02708j-t2.tif

Given the high degree of interference of the intermediate state with the cycling loop, the laser cooling analysis will be one to include the intermediate state, as done previously in the literature.69,70 The vibrational branching ratio, which represents the percentage of transition probability between two vibrational levels, is obtained by using the formula:71

 
image file: d5ra02708j-t3.tif(2)

Since we are studying the laser cooling between the two electronic states (1)3Π0+ and X1Σ0+ with the intermediate state (1)3Δ1, the values of the vibrational branching ratio Rv′′v and Rv′′v for the first five vibrational levels are given by:72

 
image file: d5ra02708j-t4.tif(3.1)
 
image file: d5ra02708j-t5.tif(3.2)
where Av′′v and Av′′v are the Einstein coefficients for the transitions (1)3Π0+X1Σ0+ and (1)3Π0+–(1)3Δ1, respectively. The corresponding calculated values of these Einstein coefficients and the vibrational branching ratio are given in Table 3, along with the value of the radiative lifetime, which is within the experimental conditions to realize the laser cooling of the molecule LaH. The experimental parameters needed to realize laser cooling are73
 
image file: d5ra02708j-t6.tif(4.1)
 
image file: d5ra02708j-t7.tif(4.2)
 
image file: d5ra02708j-t8.tif(4.3)
 
image file: d5ra02708j-t9.tif(4.4)
where h and kB are, respectively, the Planck and Boltzmann constants, m is the mass of the molecule, and V, amax, and L are the speed, the maximum acceleration, and the slowing distance, respectively. In the main cycling transition, Ne is the number of the excited states, while Ntot is the number of the excited states connected to the ground state plus Ne.

Table 3 The radiative lifetimes τ, and the vibrational branching ratio of the vibrational transitions. Between the electronic states (1)3Π0+X1Σ0+ and (1)3Π0+ – (1)3Δ1 of the molecule LaH
    ν′′ ((1)3Π0+) = 0 1 2 3 4 5
ν ((1X0+)) = 0 Avv′′ 2458.937631 130.9866596 2.52485 × 10−5 1.017740631 0.185340271 0.02356491
Rvv′′ 0.123427154 0.01235997 2.52701 × 10−9 2.52701 × 10−9 2.1245 × 10−5 3.29397 × 10−5
1 Av′′ 54.13476602 1988.819523 269.2849008 0.732617602 3.268505736 0.990[thin space (1/6-em)]576[thin space (1/6-em)]903
Rvv′′ 0.002717312 0.187666051 0.026951574 8.13052 × 10−5 0.000374658 0.001384656
2 Avv′′ 0.212860911 84.56872455 1574.649772 392.4625508 5.437394721 4.106859785
Rvv′′ 1.06846 × 10−5 0.007979949 0.15759996 0.043555105 0.000623271 0.005740684
3 Avv′′ 0.009726067 0.804036255 99.6551288 1230.988993 502.8191021 14.80275134
Rvv′′ 4.88203 × 10−7 7.58693 × 10−5 0.009974056 0.136613938 0.057636543 0.020691701
4 Avv′′ 0.000172085 0.019056646 1.710011321 101.2045812 927.9466442 601.3369858
Rvv′′ 8.63788 × 10−9 1.7982 × 10−6 0.000171148 0.011231584 0.106367552 0.840565695
ν′ ((1)3Δ1) = 0 Av'v′′ 17353.11041 56.71484895 0.36948472 6.17829E-10 6.17829E-10 0.000120469
Rv'v′′ 0.871044876 0.005351643 3.69801 × 10−5 6.8566 × 10−14 7.08197 × 10−14 1.68395 × 10−7
1 Av'v′′ 43.77651799 8225.202174 52.13234864 0.000400248 0.000400248 2.85186 × 10−6
Rv'v′′ 0.002197376 0.776134381 0.005217704 4.44191 × 10−8 4.58792 × 10−8 3.98641 × 10−9
2 Av'v′′ 10.64881134 78.79191542 7837.876719 1.211771658 1.211771658 0.001[thin space (1/6-em)]576[thin space (1/6-em)]422
Rv'v′′ 0.00053452 0.007434846 0.784459554 0.000134481 0.000138902 2.20357 × 10−6
3 Av'v′′ 1.304985207 26.12163032 111.8377067 82.58291984 82.58291984 1.955219988
Rv'v′′ 6.55041 × 10−5 0.002464851 0.011193358 0.009164971 0.009466216 0.002733061
4 Av'v′′ 0.04137685 5.623563619 43.91900286 7200.511483 7200.511483 92.1779253
Rv'v′′ 2.07692 × 10−6 0.000530642 0.004395665 0.799105624 0.825371568 0.128848888
ΣAvv′′   19922.17726 10597.65213 9991.4351 9010.713057 8723.963561 715.3955837
τ = 1/ΣAvv   5.01953 × 10−5 9.43605 × 10−5 0.000100086 0.000110979 0.000114627 0.001397828
τ (μs)   50.2 94.4 100.1 111.0 114.6 1397.8


The laser cooling scheme for the molecule LaH for the main transition (1)3Π0+X1Σ0+ with the intermediate state (1)3Δ1 is given in Fig. 6. The driving and the repumping lasers (of wavelengths λ0′′0 = 1258.8 nm, λ0′′2 = 1927.5 nm) are given in solid red lines for the transition (1)3Π0+X1Σ0+ and in solid green lines (of wavelengths λ0′′0′ = 2147.3 nm, λ0′′1′ = 1676.1 nm) for the transition (1)3Π0+–(1)3Δ1. These four suggested lasers are in the near-infrared region, a region of the spectrum for which commercial lasers are already available in the market. The spontaneous decays are represented in blue dotted lines for the transition (1)3Π0+X1Σ0+ and in purple dotted lines for the transition (1)3Π0+–(1)3Δ1. The values of the FCF (fν′′ν and fv′′v) and the vibrational branching ratios Rv′′v and Rv′′ are specified for the vibrational levels in the laser cooling scheme. The loss to the vibrational level v = 2 is negligible (R0′′2 = 1.06846 × 10−5), so that the corresponding vibrational level is not considered in the laser cooling scheme.


image file: d5ra02708j-f6.tif
Fig. 6 Laser cooling scheme of transitions X1Σ0+–(1)3Π0 and (1)3Δ1–(1)3Π0 of the molecule LaH.

The number of cycles (N) for photon absorption/emission for the vibrational levels is reciprocal to the total loss:

 
image file: d5ra02708j-t10.tif(5)

The values of the corresponding experimental parameters are L = 6.05 m, V = 3.73 m s−1, Tini = 0.117 K, Ne/Ntot = 1/5, and amax = 1.15 m s−2. For this cooling scheme, the temperature that can be reached during the process is given by the Doppler limit temperature TD and the recoil temperature Tr:71

 
TD = h/(4 × π × τ × kB) = 9.6 nK and Tr = h2/(m × λ200 × kB) = 88.1 nK, (6)

5. Conclusion

A theoretical ab initio calculation has been done based on a complete active space self-consistent field (CASSCF)/(MRCI + Q) with Davidson correction in the Λ(±) and Ω(±) representations. The calculation of the adiabatic potential energy curves, the spectroscopic constants, the FCFs, and the radiative lifetime for the transition X1Σ0+–(1)3Π0+ shows the candidacy of the LaH molecule for direct laser cooling. The calculation of the FCF and the radiative lifetime for the transition (1)3Δ1–(1)3Π0+ shows that the presence of the intermediate state (1)3Δ between X1Σ0+ and (1)3Π0+ cannot be ignored. Therefore, a total vibrational branching ratio is calculated, leading to a total radiative lifetime (τ = 50.2 μs). Experimental parameters such as the Doppler and recoil temperatures, the slowing distance, and the number of cycles (N) for photon absorption/emission are proposed. A laser cooling scheme is presented for the transition X1Σ0+–(1)3Π0+ with the intermediate state (1)3Δ1. These results open the way for direct laser cooling of the molecule LaH.

Conflicts of interest

The authors declare no competing interests.

Data availability

All data generated or analysed during this study are included in this published article.

Acknowledgements

This publication is based upon work supported by the Khalifa University of Science and Technology under Award No. RIG-2024-053 and the Abu-Dhabi Department of Education and Knowledge, under award number AARE20-031. Al MESBAR High Power Computer was used for the completion of this work.

References

  1. R. Blatt, P. Gill and R. C. Thompson, Current Perspectives on the Physics of Trapped Ions, J. Mod. Opt., 1992, 39(2), 193–220,  DOI:10.1080/09500349214550221.
  2. S. A. Diddams, Th. Udem, J. C. Bergquist, E. A. Curtis, R. E. Drullinger, L. Hollberg, W. M. Itano, W. D. Lee, C. W. Oates, K. R. Vogel and D. J. Wineland, An Optical Clock Based on a Single Trapped 199Hg+ Ion, Science, 2001, 293(5531), 825–828,  DOI:10.1126/science.1061171.
  3. K. Góral, L. Santos and M. Lewenstein, Quantum Phases of Dipolar Bosons in Optical Lattices, Phys. Rev. Lett., 2002, 88(17), 170406,  DOI:10.1103/PhysRevLett.88.170406.
  4. R. Barnett, D. Petrov, M. Lukin and E. Demler, Quantum Magnetism with Multicomponent Dipolar Molecules in an Optical Lattice, Phys. Rev. Lett., 2006, 96(19), 190401,  DOI:10.1103/PhysRevLett.96.190401.
  5. A. Micheli, G. K. Brennen and P. Zoller, A toolbox for lattice-spin models with polar molecules, Nat. Phys., 2006, 2(5), 341–347,  DOI:10.1038/nphys287.
  6. R. V. Krems, Cold controlled chemistry, Phys. Chem. Chem. Phys., 2008, 10(28), 4079–4092,  10.1039/B802322K.
  7. Y. Gao and T. Gao, Laser cooling of the alkaline-earth-metal monohydrides: Insights from an ab initio theory study, Phys. Rev. A, 2014, 90(5), 052506,  DOI:10.1103/PhysRevA.90.052506.
  8. C. Li, Y. Li, Z. Ji, X. Qiu, Y. Lai, J. Wei, Y. Zhao, L. Deng, Y. Chen and J. Liu, Candidates for direct laser cooling of diatomic molecules with the simplest 1Σ – 1Σ electronic system, Phys. Rev. A, 2018, 97(6), 062501,  DOI:10.1103/PhysRevA.97.062501.
  9. M.-S. Xu, C.-L. Yang, M.-S. Wang and X.-G. Ma, Theoretical study on the low-lying excited electronic states and laser cooling feasibility of CuH molecule, Spectrochim. Acta. A. Mol. Biomol. Spectrosc., 2019, 212, 55–60,  DOI:10.1016/j.saa.2018.12.038.
  10. M.-S. Xu, C.-L. Yang, M.-S. Wang, X.-G. Ma and W.-W. Liu, Theoretical study on the low-lying electronic excited states and laser cooling feasibility of AuH molecule, J. Quant. Spectrosc. Radiat. Transf., 2020, 242, 106770,  DOI:10.1016/j.jqsrt.2019.106770.
  11. N. Wells and I. C. Lane, Prospects for ultracold carbonvia charge exchange reactions and laser cooled carbides, Phys. Chem. Chem. Phys., 2011, 13(42), 19036–19051,  10.1039/C1CP21304K.
  12. N. Wells and I. C. Lane, Electronic states and spin-forbidden cooling transitions of AlH and AlF, Phys. Chem. Chem. Phys., 2011, 13(42), 19018–19025,  10.1039/C1CP21313J.
  13. J. H. V. Nguyen, C. R. Viteri, E. G. Hohenstein, C. D. Sherrill, K. R. Brown and B. Odom, Challenges of laser-cooling molecular ions, New J. Phys., 2011, 13(6), 063023,  DOI:10.1088/1367-2630/13/6/063023.
  14. A. El K. Nariman, et al., Electronic Structure with Spin-Orbit Coupling Effect of HfH Molecule for Laser Cooling Investigations, Spectrochim. Acta, Part A, 2024, 314, 124106,  DOI:10.1016/j.saa.2024.124106.
  15. F. Furche and J. P. Perdew, The performance of semilocal and hybrid density functionals in 3d transition-metal chemistry, J. Chem. Phys., 2006, 124(4), 044103,  DOI:10.1063/1.2162161.
  16. J. G. Harrison, Density functional calculations for atoms in the first transition series, J. Chem. Phys., 1983, 79(5), 2265–2269,  DOI:10.1063/1.446076.
  17. K. P. Jensen, B. O. Roos and U. Ryde, Performance of density functionals for first row transition metal systems, J. Chem. Phys., 2007, 126(1), 014103,  DOI:10.1063/1.2406071.
  18. J. F. Harrison, Electronic Structure of Diatomic Molecules Composed of a First-Row Transition Metal and Main-Group Element (H−F), Chem. Rev., 2000, 100(2), 679–716,  DOI:10.1021/cr980411m.
  19. A. J. Sauval, Predicted presence and tentative identification of new molecules in the pure S star R CYG, Astron. Astrophys., 1978, 62, 295–298 CAS.
  20. J. A. M. Simoes and J. L. Beauchamp, Transition metal-hydrogen and metal-carbon bond strengths: the keys to catalysis, ACS Publications, 2023,  DOI:10.1021/cr00102a004.
  21. F. A. Cotton and G. Wilkinson, Advanced Inorganic Chemistry, A Comprehensive Text, 5th edn., Wiley, New York, 1988 Search PubMed.
  22. E. R. Meyer, J. L. Bohn and M. P. Deskevich, Candidate molecular ions for an electron electric dipole moment experiment, Phys. Rev. A, 2006, 73, 062108,  DOI:10.1103/PhysRevA.73.062108.
  23. R. B. Bernard, Can. J. Phys., 1976, 54(14), 1509 CrossRef.
  24. R. S. Ram and P. F. Bernath, J. Chem. Phys., 1996, 104(17), 6444 CrossRef CAS.
  25. S. Mukund and S. Yarlagadda, Energy linkage between the singlet and triplet manifolds in LaH, and observation of new low-energy states, J. Chem. Phys., 2012, 137, 234309 CrossRef PubMed.
  26. S. Yarlagadda, S. Mukund and S. G. Nakhate, Jet-cooled laser-induced fluorescence spectroscopy of LaH: Observation of new excited electronic states, Chem. Phys. Lett., 2012, 537, 1 CrossRef CAS.
  27. S. Yarlagadda, S. Mukund, S. Bhattacharyya and S. G. Nakhate, J. Mol. Spectrosc., 2013, 289, 1 CrossRef CAS.
  28. S. Yarlagadda, S. Mukund, S. Bhattacharyya and S. G. Nakhate, J. Quant. Spectrosc. Radiat. Transfer, 2014, 145, 17 CrossRef CAS.
  29. K. D. Das and K. Balasubramanian, Chem. Phys. Lett., 1990, 172, 372 CrossRef CAS.
  30. S. Koseki, Y. Ishihara, D. G. Fedorov and H. Umeda, et al., J. Phys. Chem. A, 2004, 108(21), 4707 CrossRef CAS.
  31. S. Mahmoud and M. Korek, Theoretical calculation of the low-lying electronic states of the LaH molecule, Can. J. Chem., 2014, 92(9), 855–861,  DOI:10.1139/cjc-2014-0057.
  32. J. Assaf, R. Assaf and E. C. M. Nascimento, New insight into the spectroscopy of LaH by ab initio methods, Comput. Theor. Chem., 2021, 1203, 113363,  DOI:10.1016/j.comptc.2021.113363.
  33. H.-J. Werner, et al., The molpro quantum chemistry package, J. Chem. Phys., 2020, 152(14), 144107,  DOI:10.1063/5.0005081.
  34. A.-R. Allouche, Gabedit—a graphical user interface for computational chemistry softwares, J. Comput. Chem., 2011, 32(1), 174–182,  DOI:10.1002/jcc.21600.
  35. M. Dolg, H. Stoll, A. Savin and H. Preuss, Theor. Chim. Acta, 1989, 75, 173 CrossRef CAS.
  36. B. P. Pritchard, D. Altarawy, B. Didier, T. D. Gibson and T. L. Windus, A New Basis Set Exchange: An Open, Up-to-date Resource for the Molecular Sciences Community, J. Chem. Inf. Model., 2019, 59(11), 4814–4820,  DOI:10.1021/acs.jcim.9b00725.
  37. X. Cao and M. Dolg, Valence basis sets for relativistic energy-consistent small-core lanthanide pseudopotentials, J. Chem. Phys., 2001, 115, 7348,  DOI:10.1063/1.1406535.
  38. X. Cao and M. Dolg, Segmented contraction scheme for small-core lanthanide pseudopotential basis sets, J. Molec. Struct. (Theochem), 2002, 581, 139,  DOI:10.1016/S0166-1280(01)00751-5.
  39. R. A. Kendall, T. H. Dunning and R. J. Harrison, Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions, J. Chem. Phys., 1992, 96, 6796–6806 Search PubMed.
  40. M. Dolg, H. Stoll and H. Preuss, Theor. Chim. Acta, 1993, 85, 441 CrossRef CAS.
  41. J. Yang and M. Dolg, Theor. Chem. Acc., 2005, 113, 212 Search PubMed.
  42. A. Weigand, X. Cao, J. Yang and M. Dolg, Theor. Chem. Acc., 2009, 126, 117 Search PubMed.
  43. D. Mitra, N. B. Vilas, C. Hallas, L. Anderegg, B. L. Augenbraun, L. Baum, C. Miller, S. Raval and J. M. Doyle, Direct laser cooling of a symmetric top molecule, Science, 2020, 369, 1366–1369 CrossRef CAS PubMed.
  44. M. Lemeshko, R. V. Krems, J. M. Doyle and S. Kais, Manipulation of molecules with electromagnetic fields, Mol. Phys., 2013, 111, 1648–1682 CrossRef CAS.
  45. A. Prehn, M. Ibrügger, R. Glöckner, G. Rempe and M. Zeppenfeld, Optoelectrical Cooling of Polar Molecules to Submillikelvin Temperatures, Phys. Rev. Lett., 2016, 116, 063005 CrossRef PubMed.
  46. X. Wu, T. Gantner, M. Koller, M. Zeppenfeld, S. Chervenkov and G. Rempe, A cryofuge for cold-collision experiments with slow polar molecules, Science, 2017, 358, 645–648 CrossRef CAS PubMed.
  47. Y. Liu, M. Vashishta, P. Djuricanin, S. Zhou, W. Zhong, T. Mittertreiner, D. Carty and T. Momose, Magnetic Trapping of Cold Methyl Radicals, Phys. Rev. Lett., 2017, 118, 093201 CrossRef PubMed.
  48. K.-K. Ni, S. Ospelkaus, M. H. G. de Miranda, A. Pe’er, B. Neyenhuis, J. J. Zirbel, S. Kotochigova, P. S. Julienne, D. S. Jin and J. Ye, A high phase-space-density gas of polar molecules, Science, 2008, 322, 231–235 CrossRef CAS PubMed.
  49. L. De Marco, G. Valtolina, K. Matsuda, W. G. Tobias, J. P. Covey and J. Ye, A degenerate Fermi gas of polar molecules, Science, 2019, 363, 853–856 CrossRef CAS PubMed.
  50. E. S. Shuman, J. F. Barry and D. Demille, Laser cooling of a diatomic molecule, Nature, 2010, 467, 820–823 CrossRef CAS PubMed.
  51. S. Truppe, H. J. Williams, M. Hambach, L. Caldwell, N. J. Fitch, E. A. Hinds, B. E. Sauer and M. R. Tarbutt, Molecules cooled below the Doppler limit, Nat. Phys., 2017, 13, 1173–1176 Search PubMed.
  52. L. Anderegg, B. L. Augenbraun, E. Chae, B. Hemmerling, N. R. Hutzler, A. Ravi, A. Collopy, J. Ye, W. Ketterle and J. M. Doyle, Radio Frequency Magneto-Optical Trapping of CaF with High Density, Phys. Rev. Lett., 2017, 119, 103201 CrossRef PubMed.
  53. A. L. Collopy, S. Ding, Y. Wu, I. A. Finneran, L. Anderegg, B. L. Augenbraun, J. M. Doyle and J. Ye, 3D Magneto-Optical Trap of Yttrium Monoxide, Phys. Rev. Lett., 2018, 121, 213201 CrossRef CAS PubMed.
  54. I. Kozyryev, L. Baum, K. Matsuda, B. L. Augenbraun, L. Anderegg, A. P. Sedlack and J. M. Doyle, Sisyphus Laser Cooling of a Polyatomic Molecule, Phys. Rev. Lett., 2017, 118, 173201 CrossRef PubMed.
  55. B. L. Augenbraun, Z. D. Lasner, A. Frenett, H. Sawaoka, C. Miller, T. C. Steimle and J. M. Doyle, Laser-cooled polyatomic molecules for improved electron electric dipole moment searches, New J. Phys., 2020, 22, 022003 CrossRef CAS.
  56. L. Baum, N. B. Vilas, C. Hallas, B. L. Augenbraun, S. Raval, D. Mitra and J. M. Doyle, 1D Magneto-Optical Trap of Polyatomic Molecules, Phys. Rev. Lett., 2020, 124, 133201 CrossRef CAS PubMed.
  57. M. D. Di Rosa, Laser-Cooling Molecules: Concept, Candidates, and Supporting Hyperfine-Resolved Measurements of Rotational Lines in the AX (0, 0) Band of CaH, Eur. Phys. J. D, 2004, 31, 395–402 CrossRef.
  58. N. El-Kork, A. AlMasri Alwan, N. Abu El Kher, J. Assaf, T. Ayari, E. Alhseinat and M. Korek, Laser cooling with intermediate state of spin-orbit coupling of LuF molecule, Sci. Rep., 2023, 13(1), 7087,  DOI:10.1038/s41598-023-32439-1.
  59. A. Moussa, et al., Laser Cooling with an Intermediate State and Electronic Structure Studies of the Molecules CaCs and CaNa, ACS Omega, 2022, 7(22), 18577–18596,  DOI:10.1021/acsomega.2c01224.
  60. Y. Xiang et al., Laser-Cooling with an Intermediate Electronic State: Theoretical Prediction on Bismuth Hydride, J. Chem. Phys. 150, no. 22 ( 2019), https://pubs.aip.org/aip/jcp/article/150/22/224305/197640 Search PubMed.
  61. A. Madi, et al., Theoretical Electronic Structure with Spin–Orbit Coupling Effect of the Molecules SrAt and BaAt for Laser Cooling Studies, Sci. Rep., 2024, 14(1), 6289 CrossRef CAS PubMed.
  62. F. Rabah et al., Theoretical Spin–Orbit Laser Cooling for AlZn Molecule, J. Chem. Phys. 161, no. 15 ( 2024), https://pubs.aip.org/aip/jcp/article/161/15/154305/3317272 Search PubMed.
  63. R. Le Roy, LEVEL: A Computer Program for Solving the Radial Schrödinger Equation for Bound and Quasibound Levels, JQSRT, 2017, 186, 167–178 CrossRef CAS.
  64. G. Herzberg, Molecular spectra and molecular structure. I Diatomic molecules, ACS Publications. 2023,  DOI:10.1021/j150403a025.
  65. Y. Gao and T. Gao, Laser Cooling of BH and GaF: Insights from an Ab Initio Study, Phys. Chem. Chem. Phys., 2015, 17(16), 10830–10837 RSC.
  66. I. Zeid, et al., A Theoretical Electronic Structure with a Feasibility Study of Laser Cooling of LaNa Molecules with a Spin Orbit Effect, Phys. Chem. Chem. Phys., 2022, 24(13), 7862–7873 RSC.
  67. S.-Y. Kang, et al., Ab Initio Study of Laser Cooling of AlF+ and AlCl+ Molecular Ions, J. Phys. B: At., Mol. Opt. Phys., 2017, 50(10), 105103 CrossRef.
  68. P. F. Bernath, Spectra of Atoms and Molecules, third edn, Oxford University press 2016 Search PubMed.
  69. A. Moussa, et al., Laser Cooling with an Intermediate State and Electronic Structure Studies of the Molecules CaCs and CaNa, ACS Omega, 2022, 7(22), 18577–18596,  DOI:10.1021/acsomega.2c01224.
  70. Y. Xiang et al., Laser-Cooling with an Intermediate Electronic State: Theoretical Prediction on Bismuth Hydride, J. Chem. Phys. 150, no. 22 ( 2019), https://pubs.aip.org/aip/jcp/article/150/22/224305/197640 Search PubMed.
  71. R. Li, X. Yuan, G. Liang, Y. Wu, J. Wang and B. Yan, Laser Cooling of the SiO+ Molecular Ion: A Theoretical Contribution, Chem. Phys., 2019, 525, 110412 CrossRef CAS.
  72. Y. Xiang, H.-J. Guo, Yu-M. Wang, J.-L. Xue, H.-F. Xu and B. Yan, J. Chem. Phys., 2019, 150, 224305,  DOI:10.1063/1.5094367.
  73. J. F. Barry, Laser Cooling and Slowing of a Diatomic Molecule, Ph.D. Thesis, Yale University ( 2013 Search PubMed.

Footnote

Current affiliation: Department of Physical and Theoretical Chemistry at Comenius University in Bratislava- Slovakia.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.