Next Article in Journal
Oral Nano-Delivery of Crotoxin Modulates Experimental Ulcerative Colitis in a Mouse Model of Maximum Acute Inflammatory Response
Previous Article in Journal
Unravelling the Molecular Responses of the Yeast Schwanniomyces etchellsii to Hyperosmotic Stress in Seawater Medium Using Omic Approaches
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fine Structure Investigation and Laser Cooling Study of the CdBr Molecule

1
Faculty of Science, Beirut Arab University, P.O. Box 11-5020 Riad El Solh, Beirut 1107 2809, Lebanon
2
Department of Physics, Khalifa University, Abu Dhabi P.O. Box 57, United Arab Emirates
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2026, 27(1), 184; https://doi.org/10.3390/ijms27010184 (registering DOI)
Submission received: 5 October 2025 / Revised: 5 December 2025 / Accepted: 11 December 2025 / Published: 23 December 2025

Abstract

The ab initio calculations of the electronic structure of the low-lying electronic states of the CdBr molecule are characterized in the 2S+1Λ(+/−) and Ω(+/−) representations using the complete active-space self-consistent field (CASSCF) method, followed by the multireference configuration interaction (MRCI) method with Davidson correction (+Q). The potential energy curves are investigated, and spectroscopic parameters (Te, Re, ωe, Be, α e , μ e , and De) of the bound states are determined and analyzed. In addition, the rovibrational constants (Ev, Bv, Dv, Rmin, and Rmax) are reported for the investigated states with and without spin–orbit coupling. The electronic transition dipole moment curve (TDMC) is obtained for the C2Π1/2 − X2Σ+1/2 transition. Based on these data, Franck–Condon factors (FCFs), Einstein coefficient of spontaneous emission Aν’ν, radiative lifetime τ, vibrational branching ratios, and the associated slowing distance are evaluated. The results indicated that CdBr is a promising candidate for direct laser cooling, and a feasible cooling scheme employing four pumping and repumping lasers in the ultraviolet region with suitable experimentally accessible parameters is presented. These findings provide practical guidance for experimental spectroscopists exploring ultracold diatomic molecules and their applications.

1. Introduction

Group IIB metal monohalides have attracted broad interest across various disciplines, including high-temperature chemistry [1] and studies of chemical reactions and environmental pollution [2,3]. Among these species, cadmium and zinc bromide radicals (CdBr and ZnBr) are particularly noteworthy due to their applications in metal–halide lasers [4]. Cadmium, a toxic 4d transition metal, exhibits long atmospheric residence times and undergoes extensive long-range transport from its emission sources, which are dominated by natural volcanic activity [5,6]. Human activities release various amounts of cadmium into the atmosphere that are similar to or exceed those from natural sources, primarily driven by the incineration of wastes and non-ferrous metals production [5]. A detailed understanding of cadmium’s gas phase chemistry is essential for clarifying the mechanisms of particulate cadmium formation and for providing a foundation for future investigations into its heterogeneous chemistry. Moreover, gas phase reactions between cadmium and halogen-containing species are significant in combustion processes, thereby contributing to understanding the primary pathways by which cadmium enters the atmosphere [2,3].
In recent years, the ability to cool molecular gases to ultralow temperatures has opened new frontiers in quantum dynamics, quantum chemistry, and many-body physics, enabling precise control over molecular motion and interactions [7,8,9,10]. Cold and ultracold molecules have emerged as versatile platforms for exploring fundamental physical phenomena [11,12] and for enabling practical applications, including precision measurements [13,14], quantum information processing [15], simulation of solid state models [16,17], and ultracold controlled chemistry [18], with additional emerging applications in nanolithography and Bose-stimulated chemistry [19]. Within this context, several metallic halide molecules [20,21,22,23,24] have been investigated both theoretically and experimentally for their potential in laser cooling. In particular, bromine-containing species often show well-resolved isotopic features and significant spin–orbit coupling in vibronic transitions [25]. Our group has previously explored the electronic structures of various group IIB monohalides, including ZnI [26], ZnX (X = F, Cl, Br, I) [27], ZnF [28], HgF [29], ZnBr [30], and CdX (X = F, Cl, Br, I) [31] without explicitly considering spin–orbit coupling effects or evaluating their feasibility for laser cooling applications. In the present work, we extend this line of research to CdBr, performing high-level ab initio calculations that explicitly include spin–orbit coupling to provide accurate potential energy curves and transition properties relevant for assessing its suitability as a candidate for laser cooling studies.
Despite its chemical and environmental importance, the CdBr molecule has received limited attention in the literature, and its potential as a candidate for laser cooling remains unexplored. Experimentally, Huber and Herzberg [32] summarized a comprehensive compilation of all experimental data on the ground and low-lying excited states of the CdBr molecule. The first emission spectra of CdBr in the 307.5–324.7 nm region were examined by Wieland (1929) [33], who identified a component of the 2Π–2Σ (C–X) system with its (0,0) band at 317.7 nm and reported three additional bands attributed to another electronic system. Later, Howell [34] indicated that the probable (0,0) bands in Wieland’s measurements lie at 317.7–329.6 nm. Then, Gosav et al. [35] measured the UV absorption spectrum of CdBr produced by co-photolysis and observed both 2Π3/2-2Σ and 2Π1/2-2Σ systems. Ediger et al. [4] achieved lasing on the B 2 Σ 1 / 2 + X 2 Σ 1 / 2 + bands of CdBr at 811 nm by using an ArF laser to photodissociate CdBr2; from these measurements, they determined that the radiative lifetime for the v = 0 level of B 2 Σ 1 / 2 + is 25 ± 4 ns.
On the theoretical side, early work by Liao et al. [36] provided spectroscopic constants for the X2Σ+ state of CdBr using Density Functional Theory (DFT). Subsequently, Shepler et al. [37] performed extensive computations of the ground state X2Σ+ equilibrium structures and spectroscopic constants at the coupled cluster singles and doubles with perturbative triples level (CCSD(T)) by treating the core-valence (CV) correlation effect, the spin–orbit coupling (SOC) effect, and the scalar relativistic effects. Later, in 2007, Cheng et al. [38] optimized the ground-state geometry of CdBr and reported its spectroscopic constants, dissociation energy, and dipole moment. Extending beyond the ground state, Badreddine et al. (2017) [31] performed an intensive study on the 10 low-lying Λ-S states at the MRCI+Q level and determined their spectroscopic parameters. More recently (2018), Li et al. [39] carried out a comprehensive icMRCI+Q study of 14 Λ-S states, as well as 30 Ω states, deriving their spectroscopic constants. They also examined predissociation in the C 2 Π ( v 1 ) B 2 Σ + transition using the SOC matrix elements and calculated lifetimes for v = 0 5 in the C 2 Π and B 2 Σ + states.
In this work, the electronic structure and transition properties of the low-lying excited states of CdBr are investigated. Ab initio CASSCF/MRCI + Q calculations are employed to compute the adiabatic potential energy curves (PECs) and the permanent dipole moment curves (PDMCs) in the 2S+1Λ+/− and Ω(±) representations, both with and without spin–orbit coupling, respectively. For the bound states, the spectroscopic parameters (Te, Re, ωe, Be, α e , μ e , and De) are calculated and analyzed. In addition, the rovibrational constants (Ev, Bv, Dv, Rmin, and Rmax) are presented for the investigated bound states with and without the spin–orbit coupling. Based on these data, the transition dipole moments, the Franck−Condon factors, the Einstein coefficient, the radiative lifetimes, the vibrational branching ratio, and the slowing distance are determined for the transition X2Π1/2 − Ξ2Σ+1/2. These results show that the molecule CdBr is a good candidate for laser cooling and an optical laser cooling scheme is proposed by utilizing four lasers at a wavelength in the ultraviolet region, reaching a sub-microkelvin temperature limit.

2. Results and Discussion

2.1. The Potential Energy Curves and Spectroscopic Constants

The low-lying spin-free potential energy curves of the seven doublets and eight quartets electronic states of the CdBr molecule have been calculated and presented in Figure 1 and Figure 2, while those with spin–orbit coupling are given in Figure 3. In total, 15 Λ-S states correlating with the four lowest dissociation asymptotes Cd(1S) + Br(2P°), Cd(3P°) + Br(2P°), Cd(1P°) + Br(2P°), and Cd(3S) + Br(2P°) are calculated at the MRCI+Q level of theory. At the asymptotic limit of the separated atoms, cadmium (Cd) is in a singlet 1S state and bromine (Br) in a doublet 2P° state (odd parity). Following the correlation rules summarized by Huber and Herzberg [32], the combination of these atomic levels yields the formation of the molecular states X2Σ+ and (1)2Π states. Figure 1 confirms that our curves reproduce these states, which constitute the first dissociation channel. The second dissociation asymptote Cd(3P°) + Br(2P°) correlates with a set of bound states B2Σ+, C2Π, (1)4Δ, (1)4Σ+, (3)2Σ+, (1)2Δ, (1)4Σ, and (1)2Σ and predominantly repulsive states, namely (1)4Π, (2)4Π, (3)2Π, and (2)4Σ+. Similarly, higher channels also follow the expected dissociation correlations where (2)2Δ state correlates correctly with Cd(1P°) + Br(2P°), and (3)4Π and (3)4Σ+ states are attributed to the Cd(3S) + Br(2P°) dissociation channel. Any additional states not shown at specific limits are absent simply because they fall outside the set of states examined in the present work. A comparison of our calculated asymptotic-limit energy separations E theo (between each higher dissociation channel and the lowest one, without spin–orbit coupling) with the experimental atomic separations E exp from the National Institute of Standards and Technology website (NIST) [40] is summarized in Table 1. The second dissociation channel lies 26,228.93 cm−1 above the first, whereas the experimental value is 31,246.335 cm−1, resulting in a relative error of 16.0%. The third channel lies at 40,452.05 cm−1, compared with the experimental value of 43,692.384 cm−1, which yields a good agreement with a percentage relative error of 7.4%. On the other hand, the fourth channel lies at 40,452.05 cm−1; compared with the experimental value (45,889.49 cm−1), this corresponds to 10.9%.
The spectroscopic parameters such as the transition energy with respect to the ground state minimum Te, the equilibrium bond length Re, the harmonic frequency ωe, the rotational constant Be, the vibration–rotation interaction constant α e , the dissociation energy De, and the permanent dipole moment μ e are calculated and analyzed for the investigated bound electronic states upon fitting the corresponding potential energy curve into a polynomial around the internuclear distance at equilibrium Re. Table 2 illustrates the spectroscopic constants for the lowest electronic states of the CdBr molecule without spin–orbit coupling. The spectroscopic constants obtained for the spin-free Λ–S states of CdBr show good overall consistency with experiment [32] and prior theory [31,36,37,38,39]. For the ground state X2Σ+, our values (Re = 2.468 Å, ωe = 231.4 cm−1, Be = 5.928 cm−1) reproduce the available experimental data [32] very closely, with an (Obs. − Calc.) difference in only −0.9 cm−1 for ωe, confirming an accurately described equilibrium geometry and reliable representation of the potential energy surface. For the B2Σ+ state, the computed Te and Re values are in good agreement with earlier theoretical reports [31,39], and the corresponding vibrational constant ωe follows the expected trend of bond weakening upon electronic excitation. Similarly, for the C2Π state, the calculated Te = 29,340 cm−1 and ωe = 268 cm−1 show reasonable agreement with the corresponding experimental values (31,075 cm−1 and 254.5 cm−1) [32], with deviations of 1735 cm−1 in T e (≈5.6%) and 13.5 cm−1 in ω e (≈5.3%), which lie within the typical accuracy range of MRCI-level calculations. These results collectively confirm the internal consistency and predictive capability of the present calculations across all investigated Λ–S states of CdBr.
After considering the SOC effect, there are 18 Ω states investigated from the two lowest dissociation channels of the Λ–S states of CdBr, as shown in Figure 3. Table 3 summarizes the dissociation behavior of the lowest Ω states that arise from these dissociation asymptotes. For each channel, the energy separation from the lowest asymptote is reported along with the experimental data from the NIST Atomic Spectra Database (NIST) [40] and the corresponding percent error. Overall, the calculated energy separations in this work are in fair agreement with the experimental measurements, with an average percentage relative error of 10.5% over the examined asymptotes. As is well known, the SOC effect is very important to the spectroscopy and dynamics of molecules containing heavy atoms [41,42,43,44]. This requirement is validated by the notably large splitting energies observed for the C2Π state (approximately 930 cm−1 between C2Π1/2 and C2Π3/2) electronic states. These results highlight the significant influence of spin–orbit coupling on the electronic states of the CdBr molecule. Table 4 presents the spectroscopic constants for bound Ω states of CdBr, including spin–orbit coupling, compared to available experimental data [32] and previous theoretical calculations [31,36,37,38,39]. For the ground X2Σ+1/2 state, our harmonic frequency ωe differs from experiment [32] by only (Obs. − Cal.) = −0.1 cm−1, while earlier theoretical studies [31,36,37,38,39] reported deviations ranging from 2.7 to 32 cm−1. The excited (1)2Π1/2 state is found at Te = 1678 cm−1 and exhibits a small vibrational frequency ωe = 40.3 cm−1, indicating a weak bonding. The splitting of C2Π state is also well reproduced: C2Π1/2 is predicted at Te = 28,860 cm−1 with (Obs. − Cal.) = 1440 cm−1 (4.8%), and C2Π3/2 at Te = 29,790 cm−1 with (Obs. − Cal.) = 1673 cm−1 (5.3%), preserving the correct spin–orbit ordering with the experiment. For C2Π3/2, our ωe deviates from experiment by only 0.2 cm−1 compared with 17.6 cm−1 in prior calculation [39]. Overall, the explicit treatment of spin–orbit coupling yields spectroscopic parameters that follow experimental trends for the low-lying states and offers quantitatively reliable vibrational constants where benchmarks exist.

2.2. Rovibrational Calculations

A theoretical calculation of the rovibrational constants for different vibrational levels of a given transition can predict the absorption/emission line positions in the spectrum of a molecule [45,46]. Moreover, vibrational states of molecules have been suggested as qubit encodings for quantum computers [47]. By using the canonical function approach [48,49,50] and the cubic spline interpolation of the PEC between every two consecutive points, the rovibrational constants such as the vibrational energy Eν, the rotational constant Bν, the centrifugal distortion constant Dν, and the abscissas of turning points Rmin and Rmax have been calculated for the lowest vibrational levels of the CdBr molecule for states in both the (Λ-S) and Ω representations. The rovibrational constants of the states X2Σ1/2 and C2Π1/2 are presented in Table 5. The values for five electronic states are provided up to v = 75 for the spin-free and two electronic states up to v = 14 for the spin–orbital coupling, as listed, respectively, in Tables S1 and S2 of the Supplementary Materials. To the best of current knowledge, no prior rovibrational constants for CdBr are available for direct comparison. The present values, therefore, constitute the first comprehensive dataset for this molecule.

2.3. Transition Properties and Laser Cooling Scheme

Laser cooling is a process that slows atoms or molecules by repeatedly scattering photons through fast and well-controlled optical transitions [51]. The momentum transfer from each scatter decreases the sample’s kinetic energy and entropy. Although molecules can also be cooled and trapped by direct methods [52,53,54,55] (e.g., buffer gas cooling, Stark deceleration) or by indirect routes [56,57] (e.g., photoassociation). Laser cooling has achieved sub-millikelvin temperatures across many diatomic [58,59,60,61] and linear triatomic systems [62,63,64]. Selecting appropriate cooling transitions among vibronic levels in a diatomic molecule requires the consideration of several important criteria, which are detailed below.

2.3.1. Highly Diagonal Franck–Condon Array

A highly diagonal Franck–Condon pattern in the considered band system minimizes the number of repump lasers and maintains a closed cycling loop; this condition is typically identified by a very small difference in the equilibrium internuclear separation (ΔRe) between the involved electronic states [65,66].

2.3.2. Absence of Intervening Electronic States

An intervening state is an intermediate level between the excited and ground states of the loop that has a sufficiently large transition probability from the excited state to act as a loss channel; if that probability is small, the state is non-intervening. Recent studies indicate that intermediate states can still be incorporated into the cooling dynamics in some schemes [67].

2.3.3. Short Radiative Lifetimes

The radiative lifetime of the chosen vibrational transition should be short to achieve high photon-scattering rates during the cooling process; typical useful lifetimes fall in the ns–ms range [68,69].
Through the comparison of the values of the internuclear distance Re of the two electronic states X2Σ+1/2 and C2Π1/2 in Table 4, a shows a very small difference, ΔRe = 0.007 Å, which leads to a diagonal Franck–Condon factor (FCF) for the transition between vibrational levels of these two electronic states. The calculated values of the Franck–Condon factor (FCF) for the transition C2Π1/2 − X2Σ+1/2 are represented in Figure 4. With the diagonality of the FCF for this transition, the first criterion (Section 2.3.1) of a closed laser cooling cycle is verified.
As shown in Figure 3, there are two unbound intervening states (1)2Π1/2 and (1)2Π3/2 between the two considered electronic states X2Σ+1/2 and C2Π1/2, which have no influence on the laser cycling loop between these two because of the very low transition probability between C2Π1/2-(1)2Π3/2 and C2Π1/2-(1)2Π1/2 [70]. Consequently, these states do not perturb the laser cooling cycle between the X2Σ+1/2 and C2Π1/2 states. Hence, the second criterion (Section 2.3.2.), the absence of intervening states, is fulfilled.
Aside from the diagonality of the FCF and the absence of intervening states, a sufficiently short lifetime (criterion (iii)) is desirable to ensure a sufficient photon scattering rate for rapid laser cooling. This radiative lifetime (τv′) can be evaluated by the inverse of the Einstein coefficient Av′v, (τv′ = 1/ΣvAv′v). The vibrational Einstein Coefficient among the transition X2Σ+1/2 − C2Π1/2 is calculated using the LEVEL 11 program [71] according to the following formula [46]:
A v v = ( 3.1361891 ) ( 10 7 ) ( E ) 3 ( ψ ν M ( r ) ψ ν ) 2 2
where Aνν′ has as units s−1, E is the emission frequency (in cm−1) and M(r) is the electronic transition dipole moment between the two considered electronic states X2S+1/2 and C2P1/2 calculated by using the MOLPRO program [72] and represented in Figure 5.
The vibrational branching ratio (Rνν), which represents the transition probability between two vibrational levels of the considered electronic states, is obtained by using the following formula [73]:
R ν ν   =   A ν ν A ν ν
The corresponding calculated values of Einstein coefficients and the vibrational branching ratio are given in Table 6, along with the values of the radiative lifetime. However, with the diagonal FCF (Figure 4), the short radiative lifetime (of 16.07 ns < t < 41.35 ns) for the transition C2Π1/2 − X2Σ+1/2 and the absence of an influencing intervening electronic state, the CdBr molecule is a good candidate for Doppler laser cooling. For this transition, our suggested laser cooling scheme is given in Figure 6. In the ultraviolet range, the red lines represent the cycling pumping lasers, where the wavelength of the main pumping laser is λ0′0= 349.6 nm. To close the leaks from the higher vibrational level v′ = 0 of the electronic state, C2Π1/2, there is a need for the three repumping lasers of wavelengths λ0′1 = 352.5 nm, λ0′2 = 355.4 nm, and λ0′3 = 358.3 nm, where the dotted green lines represent the spontaneous decay from the higher electronic state. For the transition C2Π1/2 − X2Σ+1/2 of the CdBr molecule. In Figure 6, f0′v (v = 0, 1, 2, 3) denotes the calculated Franck–Condon factors for spontaneous decay from the excited level v′ = 0 to the ground-state level v, and R0′v (v = 0, 1, 2, 3) represent the vibrational branching ratios, which correspond to the transition probabilities for spontaneous emission from the v′ = 0 of C2Π1/2 to v of X2Σ+1/2. For this transition, the total number N of cycles for photon absorption/emission, which is the reciprocal of the total loss, is given by
N = 1/(1 − ɳ) = 34,573
with ɳ = R0′0 + R0′1 + R0′2 + R0′3 [74]. The corresponding experimental parameters for the considered transitions are as follows [75]:
V r m s = h N m λ 00   =   207.2   m / s
T ini = m V 2   2 k B   =   496.0   K
a max = h N e N tot m λ 00 τ   =   3.89 × 10 4   m / s 2
L min = k B T ini m a max   =   55.2   cm
TD = 1.24 × 10−4 K and Tr = 8.31 10−7 K
where h, m, kB, Tini, and amax represent the Planck constant, the mass, the Boltzmann constant, the initial temperature, and the maximum acceleration of the molecule, respectively; Vrms is the rms velocity, Lmin is the minimum slowing distance. In the main cycling transition, Ne is the number of excited states, and Ntot is the number of excited states connected to the ground state plus Ne. Tr and TD are the recoil and the Doppler temperatures, respectively.
One must be cautious, however, that the proposed parameters in this work are primarily based on theoretical work. Consequently, one should pay attention to various caveats before proceeding with any experimental initiative. We cite a few in the following:
(1)
Our calculations show a Te value for the state C2Π1/2 that differs by about 1400 cm−1 with respect to previous experimental work. Even though minimal when taking into account the state C2Π1/2 is a high-lying state, ~30,000 cm−1 above the ground state, such a difference has still to be considered when setting up pumping laser wavelengths.
(2)
The calculated harmonic vibrational frequency ωe, shows very good agreement with the experiment for the ground X2Σ+1/2 and C2Π3/2 states. An experimental validation of the same parameter for the C2Π1/2 state is highly recommended.
(3)
A preliminary investigation of the position of the vibrational level v′ = 0 has shown no crossings with the neighboring C2Π3/2 and B2Σ+1/2; however, a direct confirmation is yet to be made available.

3. Materials and Methods

The present theoretical work is based on ab initio calculations, which are performed using the state-averaged complete active space self-consistent field (CASSCF) method followed by the multi-reference configuration interaction (MRCI) method with Davidson correction (+Q) [76]. All computations are carried out using the computational chemistry program MOLPRO 2015.1 package [72], taking advantage of the graphical interface program GABEDIT [77].
The Cadmium (Cd) atom, consisting of forty-eight electrons, has been modeled using the relativistic energy-consistent pseudopotential ECP28MDF [78] together with the aug-cc-pV5Z-PP basis set [78,79], including spdfg functions. In this pseudopotential, twenty-eight electrons are frozen in the core of Cd, and the remaining twenty electrons are left for explicit treatment. For bromine (Br), the relativistic ECP28MDF pseudopotential has been used in combination with its (6s6p)/[4s4p] valence basis [80], replacing twenty-eight core electrons, and treating the remaining seven electrons explicitly. A complete active space self-consistent field (CASSCF) calculation was performed with seven valence electrons from CdBr distributed over nine molecular orbitals; hence, this active space is referred to as CAS(7,9). The active symmetry molecular orbitals 5a1, 2b1, 2b2, and 0a2 are distributed into the irreducible representation of the C2v point group symmetry. These arise from five σ-type orbitals (Cd: 5s, 5p0, 6s, Br: 4p0, 5s) and two π-type orbital pairs (Cd: 5p±1, Br: 4p±1). In the subsequent MRCI+Q calculations, the Cd (4s2 4p6 4d10) and Br (4s2) orbitals were kept as closed-shell orbitals and included in the correlation treatment.

4. Conclusions

A comprehensive ab initio study of the CdBr molecule was performed to investigate its electronic structure and assess its potential for direct laser cooling. Calculations based on a complete active space self-consistent field (CASSCF)/(MRCI + Q) method yielded 15 spin-free electronic states and 18 spin–orbit electronic states. The adiabatic potential energy curves for the ground and excited states, together with derived spectroscopic (Te, Re, ωe, Be, α e , μ e , and De) and rovibrational (Ev, Bv, Dv, Rmin, and Rmax) constants of the bound states have been studied. The results show good agreement with previously reported data, confirming the reliability and accuracy of the computational approach. A detailed analysis of the transition dipole moment, Franck–Condon factors (FCFs), the Einstein coefficients ( A v v ), and the spontaneous radiative lifetimes for the transition X2Σ1/2 − C2Π1/2 reveal a strongly diagonal Franck–Condon array, the absence of perturbing intermediate electronic states, and short radiative lifetimes. These characteristics fulfill the principal criteria required for direct laser cooling of molecules. Furthermore, vibrational branching ratios, the number of photon absorption–emission cycles (N), and the recoil and Doppler temperature limits were estimated, providing essential parameters for experimental implementation. A feasible laser cooling scheme is proposed, employing four pumping and repumping lasers in the ultraviolet region. Overall, the present results highlight CdBr as a promising candidate for direct laser cooling, opening up opportunities for future experimental investigations.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms27010184/s1.

Author Contributions

A.M.: Investigation and writing (original draft); I.Z.: data curation and writing (original draft); N.A.E.K.: validation and writing (review and editing); N.E.-K.: resources and writing (review and editing); M.K.: supervision and writing (review and editing). All authors have read and agreed to the published version of the manuscript.

Funding

This publication is based upon work supported by the Khalifa University of Science and Technology under Award No. RIG-2024-053 the Abu-Dhabi Department of Education and Knowledge under AARE Award AARE20-031.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding author.

Acknowledgments

The authors would like to acknowledge the use of the Al Mesbar High power computer to complete their work.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Carlson, K.D.; Claydon, C.R. Advances in High Temperature Chemistry; Eyring, L., Ed.; Academic Press: Cambridge, MA, USA, 1967; Volume 1, p. 43. [Google Scholar]
  2. Ma, L.Q.; Rao, G.N. Chemical fractionation of cadmium, copper, nickel, and zinc in contaminated soils. J. Environ. Qual. 1997, 26, 259–264. [Google Scholar] [CrossRef]
  3. Olajire, A.A.; Ayodele, E.T.; Oyedirdan, G.O.; Oluyemi, E.A. Levels and speciation of heavy metals in soils of industrial southern Nigeria. Environ. Monit. Assess. 2003, 85, 135–155. [Google Scholar] [CrossRef] [PubMed]
  4. Ediger, M.N.; McCown, A.W.; Eden, J.G. CdI and CdBr photodissociation lasers at 655 and 811 nm: CdI spectrum identification and enhanced laser output with 114CdI2. Appl. Phys. Lett. 1982, 40, 99–101. [Google Scholar] [CrossRef]
  5. Williams, C.R.; Harrison, R.M. Cadmium in the atmosphere. Experientia 1984, 40, 29–36. [Google Scholar] [CrossRef]
  6. Bewers, J.M.; Barry, P.J.; MacGregor, D.J. Distribution and cycling of cadmium in the environment. Adv. Environ. Sci. Tech. 1987, 19, 1. [Google Scholar]
  7. Krems, R.V. Molecules near absolute zero and external field control of atomic and molecular dynamics. Int. Rev. Phys. Chem. 2005, 24, 99–118. [Google Scholar] [CrossRef]
  8. Weinstein, J.D.; deCarvalho, R.; Guillet, T.; Friedrich, B.; Doyle, J.M. Magnetic trapping of calcium monohydride molecules at millikelvin temperatures. Nature 1998, 395, 148–150. [Google Scholar] [CrossRef]
  9. Friedrich, B.; Herschbach, D. Alignment and trapping of molecules in intense laser fields. Phys. Rev. Lett. 1995, 74, 4623–4626. [Google Scholar] [CrossRef]
  10. Bohn, J.L.; Rey, A.M.; Ye, J. Cold molecules: Progress in quantum engineering of chemistry and quantum matter. Science 2017, 357, 1002–1010. [Google Scholar] [CrossRef]
  11. Kajita, M. Sensitive measurement of mp/me variance using vibrational transition frequencies of cold molecules. New J. Phys. 2009, 11, 055010. [Google Scholar] [CrossRef]
  12. Hudson, E.R.; Lewandowski, H.J.; Sawyer, B.C.; Ye, J. Cold molecule spectroscopy for constraining the evolution of the fine-structure constant. Phys. Rev. Lett. 2006, 96, 143004. [Google Scholar] [CrossRef] [PubMed]
  13. Blatt, R.; Gill, P.; Thompson, R.C. Current perspectives on the physics of trapped ions. J. Mod. Opt. 1992, 39, 193–220. [Google Scholar] [CrossRef]
  14. Diddams, S.A.; Udem, T.; Bergquist, J.C.; Curtis, E.A.; Drullinger, R.E.; Hollberg, L.; Itano, W.M.; Lee, W.D.; Oates, C.W.; Vogel, K.R.; et al. An optical clock based on a single trapped 199Hg+ ion. Science 2001, 293, 825–828. [Google Scholar] [CrossRef]
  15. Micheli, A.; Brennen, G.K.; Zoller, P. A toolbox for lattice-spin models with polar molecules. Nat. Phys. 2006, 2, 341–347. [Google Scholar] [CrossRef]
  16. Góral, K.; Santos, L.; Lewenstein, M. Quantum phases of dipolar bosons in optical lattices. Phys. Rev. Lett. 2002, 88, 170406. [Google Scholar] [CrossRef]
  17. Barnett, R.; Petrov, D.; Lukin, M.; Demler, E. Quantum magnetism with multicomponent dipolar molecules in an optical lattice. Phys. Rev. Lett. 2006, 96, 190401. [Google Scholar] [CrossRef]
  18. Krems, R.V. Cold controlled chemistry. Phys. Chem. Chem. Phys. 2008, 10, 4079–4092. [Google Scholar] [CrossRef]
  19. 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]
  20. Fu, M.; Cao, J.; Ma, H.; Bian, W. Laser cooling of copper monofluoride: A theoretical study including spin–orbit coupling. RSC Adv. 2016, 6, 100568–100576. [Google Scholar] [CrossRef]
  21. Yang, Z.; Li, J.; Lin, Q.; Xu, L.; Wang, H.; Yang, T.; Yin, J. Laser-cooled HgF as a promising candidate for measuring the electric dipole moment of the electron. Phys. Rev. A 2019, 99, 032502. [Google Scholar] [CrossRef]
  22. Isaev, T.A.; Hoekstra, S.; Berger, R. Laser-cooled RaF as a promising candidate to measure molecular parity violation. Phys. Rev. A 2010, 82, 052521. [Google Scholar] [CrossRef]
  23. Zhao, S.; Cui, J.; Li, R.; Yan, B. Configuration interaction study of electronic structures of CdCl including spin–orbit coupling. Chem. Phys. Lett. 2017, 677, 92–98. [Google Scholar] [CrossRef]
  24. Yuan, X.; Gomes, A.S.P. Reassessing the potential of TlCl for laser cooling experiments via four-component correlated electronic structure calculations. arXiv 2022, arXiv:2203.12493. [Google Scholar] [CrossRef] [PubMed]
  25. Tao, C.; Deselnicu, M.; Mukarakate, C.; Reid, S.A. Electronic spectroscopy of the à 1 A″ ↔ X ~ 1 A′ system of CDBr. J. Chem. Phys. 2006, 125, 094305. [Google Scholar] [CrossRef]
  26. Elmoussaoui, S.; Korek, M. Electronic structure with dipole moment calculation of the low-lying electronic states of the molecule ZnI. J. Quant. Spectrosc. Radiat. Transf. 2015, 161, 131–135. [Google Scholar] [CrossRef]
  27. Elmoussaoui, S.; El-Kork, N.; Korek, M. Electronic structure of the ZnCl molecule with rovibrational and ionicity studies of the ZnX (X = F, Cl, Br, I) compounds. Comput. Theor. Chem. 2016, 1090, 94–104. [Google Scholar] [CrossRef]
  28. Elmoussaoui, S.; El-Kork, N.; Korek, M. Electronic structure with dipole moment and ionicity calculations of the low-lying electronic states of the ZnF molecule. Can. J. Chem. 2017, 95, 22–27. [Google Scholar] [CrossRef]
  29. Elmoussaoui, S.; Chmaisani, W.; Korek, M. Theoretical electronic structure with dipole moment and rovibrational calculation of the low-lying electronic states of the HgF molecule. J. Quant. Spectrosc. Radiat. Transf. 2017, 201, 64–74. [Google Scholar] [CrossRef]
  30. Elmoussaoui, S.; Korek, M. Electronic structure with dipole moment calculation of the low-lying electronic states of ZnBr molecule. Comput. Theor. Chem. 2015, 1068, 42–46. [Google Scholar] [CrossRef]
  31. Badreddine, K.; Korek, M. Theoretical electronic structure of the cadmium monohalide molecules CdX (X = F, Cl, Br, I). AIP Adv. 2017, 7, 105320. [Google Scholar] [CrossRef]
  32. Huber, K.P.; Herzberg, G. Molecular Spectra and Molecular Structure. IV. In Constants of Diatomic Molecules; Van Nostrand Reinhold: New York, NY, USA, 1979. [Google Scholar]
  33. Wieland, K. Bandenspektren der Quecksilber-, Cadmium- und Zinkhalogenide. Helv. Phys. Acta 1929, 2, 46. [Google Scholar]
  34. Howell, H.G. The ultraviolet spectra and electron configuration of HgF and related halide molecules. Proc. R. Soc. Lond. A 1943, 182, 95–112. [Google Scholar]
  35. Gosavi, R.K.; Greig, G.; Young, P.J.; Strausz, O.P. Reactions of metal atoms. IV. The UV spectra of CdBr, CdI, ZnBr, and ZnI. J. Chem. Phys. 1971, 54, 983–991. [Google Scholar] [CrossRef]
  36. Liao, M.; Zhang, Q.; Schwarz, W.H.E. Properties and stabilities of MX, MX2, and M2X2 compounds (M = Zn, Cd, Hg; X = F, Cl, Br, I). Inorg. Chem. 1995, 34, 5597–5605. [Google Scholar] [CrossRef]
  37. Shepler, B.C.; Peterson, K.A. Chemically accurate thermochemistry of cadmium: An ab initio study of Cd + XY (X = H, O, Cl, Br; Y = Cl, Br). J. Phys. Chem. A 2006, 110, 12321–12329. [Google Scholar] [CrossRef]
  38. Cheng, L.; Wang, M.Y.; Wu, Z.J.; Su, Z.M. Electronic structures and chemical bonding in 4d transition metal monohalides. J. Comput. Chem. 2007, 28, 2190–2202. [Google Scholar] [CrossRef]
  39. Li, R.; Wang, M.; Zhao, S.; Zu, N.; Yan, B. Spin–orbit calculations on the ground and excited electronic states of CdBr molecule. J. Quant. Spectrosc. Radiat. Transf. 2018, 221, 110–117. [Google Scholar] [CrossRef]
  40. Wiese, W.L. The NIST atomic spectra database. AIP Conf. Proc. 2000, 543, 299–303. [Google Scholar] [CrossRef]
  41. Minaev, B.F.; Knuts, S.; Ågren, H. On the interpretation of the external heavy atom effect on singlet–triplet transitions. Chem. Phys. 1994, 181, 15–28. [Google Scholar] [CrossRef]
  42. Liu, S.Y.; Zhang, X.M.; Zhai, H.S.; Liu, Y.F. Accurate predictions of spectroscopic and molecular properties of the NTe molecule. J. Quant. Spectrosc. Radiat. Transf. 2017, 202, 50–57. [Google Scholar] [CrossRef]
  43. El Kher, N.A.; Korek, M.; Alharzali, N.; El-Kork, N. Electronic structure with spin–orbit coupling effect of HfH molecule for laser cooling investigations. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2024, 314, 124106. [Google Scholar] [CrossRef] [PubMed]
  44. El Kher, N.A.; El-Kork, N.; Alharzali, N.; Assaf, J.; Korek, M. Fine structure investigation and laser cooling study of the LaH molecule. RSC Adv. 2025, 15, 41699–41709. [Google Scholar] [CrossRef]
  45. Brown, J.M.; Carrington, A. Rotational Spectroscopy of Diatomic Molecules; Cambridge University Press: Cambridge, UK, 2003. [Google Scholar]
  46. Bernath, P.F. Spectra of Atoms and Molecules, 4th ed.; Oxford University Press: Oxford, UK, 2020. [Google Scholar]
  47. Tesch, C.M.; Kurtz, L.; de Vivie-Riedle, R. Applying optimal control theory for elements of quantum computation in molecular systems. Chem. Phys. Lett. 2001, 343, 633–641. [Google Scholar] [CrossRef]
  48. Korek, M.; Kobeissi, H. New analytical expression for the rotational factor in Raman transitions. Can. J. Phys. 1995, 73, 559–565. [Google Scholar] [CrossRef]
  49. Korek, M.; Kobeissi, H. Highly accurate diatomic centrifugal distortion constants for high orders and high levels. J. Comput. Chem. 1992, 13, 1103–1108. [Google Scholar] [CrossRef]
  50. Korek, M. A one directional shooting method for the computation of diatomic centrifugal distortion constants. Comput. Phys. Commun. 1999, 119, 169–178. [Google Scholar] [CrossRef]
  51. Mitra, D.; Vilas, N.B.; Hallas, C.; Anderegg, L.; Augenbraun, B.L.; Baum, L.; Miller, C.; Raval, S.; Doyle, J.M. Direct laser cooling of a symmetric top molecule. Science 2020, 369, 1366–1369. [Google Scholar] [CrossRef]
  52. Lemeshko, M.; Krems, R.V.; Doyle, J.M.; Kais, S. Manipulation of molecules with electromagnetic fields. Mol. Phys. 2013, 111, 1648–1682. [Google Scholar] [CrossRef]
  53. Prehn, A.; Ibrügger, M.; Glöckner, R.; Rempe, G.; Zeppenfeld, M. Optoelectrical cooling of polar molecules to submillikelvin temperatures. Phys. Rev. Lett. 2016, 116, 063005. [Google Scholar] [CrossRef]
  54. Wu, X.; Gantner, T.; Koller, M.; Zeppenfeld, M.; Chervenkov, S.; Rempe, G. A cryofuge for cold-collision experiments with slow polar molecules. Science 2017, 358, 645–648. [Google Scholar] [CrossRef]
  55. Liu, Y.; Vashishta, M.; Djuricanin, P.; Zhou, S.; Zhong, W.; Mittertreiner, T.; Carty, D.; Momose, T. Magnetic trapping of cold methyl radicals. Phys. Rev. Lett. 2017, 118, 093201. [Google Scholar] [CrossRef] [PubMed]
  56. Ni, K.-K.; Ospelkaus, S.; de Miranda, M.H.G.; Pe’er, A.; Neyenhuis, B.; Zirbel, J.J.; Kotochigova, S.; Julienne, P.S.; Jin, D.S.; Ye, J. A high phase-space-density gas of polar molecules. Science 2008, 322, 231–235. [Google Scholar] [CrossRef]
  57. De Marco, L.; Valtolina, G.; Matsuda, K.; Tobias, W.G.; Covey, J.P.; Ye, J. A degenerate Fermi gas of polar molecules. Science 2019, 363, 853–856. [Google Scholar] [CrossRef]
  58. Shuman, E.S.; Barry, J.F.; DeMille, D. Laser cooling of a diatomic molecule. Nature 2010, 467, 820–823. [Google Scholar] [CrossRef]
  59. Truppe, S.; Williams, H.J.; Hambach, M.; Caldwell, L.; Fitch, N.J.; Hinds, E.A.; Sauer, B.E.; Tarbutt, M.R. Molecules cooled below the Doppler limit. Nat. Phys. 2017, 13, 1173–1176. [Google Scholar] [CrossRef]
  60. Anderegg, L.; Augenbraun, B.L.; Chae, E.; Hemmerling, B.; Hutzler, N.R.; Ravi, A.; Collopy, A.; Ye, J.; Ketterle, W.; Doyle, J.M. Radio frequency magneto-optical trapping of CaF with high density. Phys. Rev. Lett. 2017, 119, 103201. [Google Scholar] [CrossRef]
  61. Collopy, A.L.; Ding, S.; Wu, Y.; Finneran, I.A.; Anderegg, L.; Augenbraun, B.L.; Doyle, J.M.; Ye, J. 3D magneto-optical trap of yttrium monoxide. Phys. Rev. Lett. 2018, 121, 213201. [Google Scholar] [CrossRef]
  62. Kozyryev, I.; Baum, L.; Matsuda, K.; Augenbraun, B.L.; Anderegg, L.; Sedlack, A.P.; Doyle, J.M. Sisyphus laser cooling of a polyatomic molecule. Phys. Rev. Lett. 2017, 118, 173201. [Google Scholar] [CrossRef] [PubMed]
  63. Augenbraun, B.L.; Lasner, Z.D.; Frenett, A.; Sawaoka, H.; Miller, C.; Steimle, T.C.; Doyle, J.M. Laser-cooled polyatomic molecules for improved electron electric dipole moment searches. New J. Phys. 2020, 22, 022003. [Google Scholar] [CrossRef]
  64. Baum, L.; Vilas, N.B.; Hallas, C.; Augenbraun, B.L.; Raval, S.; Mitra, D.; Doyle, J.M. 1D magneto-optical trap of polyatomic molecules. Phys. Rev. Lett. 2020, 124, 133201. [Google Scholar] [CrossRef] [PubMed]
  65. Di Rosa, M.D. 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. [Google Scholar] [CrossRef]
  66. El-Kork, N.; AlMasri Alwan, A.; Abu El Kher, N.; Assaf, J.; Ayari, T.; Alhseinat, E.; Korek, M. Laser cooling with intermediate state of spin–orbit coupling of LuF molecule. Sci. Rep. 2023, 13, 7087. [Google Scholar] [CrossRef]
  67. Xiang, Y.; Guo, H.J.; Wang, Y.M.; Xue, J.L.; Xu, H.F.; Yan, B. Laser-cooling with an intermediate electronic state: Theoretical prediction on bismuth hydride. J. Chem. Phys. 2019, 150, 224305. [Google Scholar] [CrossRef]
  68. Yassine, K.; El Kork, N.; Abu El Kher, N.; Younes, G.; Korek, M. Theoretical study of spin–orbit coupling and laser cooling for HBr molecule with first-overtone spectral calculations. Front. Phys. 2025, 13, 1635859. [Google Scholar] [CrossRef]
  69. Rabah, F.; Chmaisani, W.; Younes, G.; El-Kork, N.; Korek, M. Theoretical spin–orbit laser cooling for AlZn molecule. J. Chem. Phys. 2024, 161, 154305. [Google Scholar] [CrossRef]
  70. Yang, Q.S.; Gao, T. The feasibility of laser cooling: An ab initio investigation of MgBr, MgI, and MgAt molecules. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2020, 231, 118107. [Google Scholar] [CrossRef]
  71. Le Roy, R.J. LEVEL: A computer program for solving the radial Schrödinger equation for bound and quasibound levels. J. Quant. Spectrosc. Radiat. Transf. 2017, 186, 167–178. [Google Scholar] [CrossRef]
  72. Werner, H.-J.; Knowles, P.J. Getting Started with MOLPRO, Version 2015.1; Molpro.net: Cardiff, UK, 2015; Available online: https://www.molpro.net (accessed on 31 October 2025).
  73. Li, R.; Yuan, X.; Liang, G.; Wu, Y.; Wang, J.; Yan, B. Laser cooling of the SiO+ molecular ion: A theoretical contribution. Chem. Phys. 2019, 525, 110412. [Google Scholar] [CrossRef]
  74. Metcalf, H.J.; van der Straten, P. Laser cooling and trapping of atoms. J. Opt. Soc. Am. B 2003, 20, 887. [Google Scholar] [CrossRef]
  75. Barry, J.F. Laser Cooling and Slowing of a Diatomic Molecule. Ph.D. Thesis, Yale University, New Haven, CT, USA, 2013. [Google Scholar]
  76. Werner, H.-J.; Knowles, P.J. An efficient internally contracted multiconfiguration–reference configuration interaction method. J. Chem. Phys. 1988, 89, 5803–5814. [Google Scholar] [CrossRef]
  77. Allouche, A.R. Gabedit—A graphical user interface for computational chemistry softwares. J. Comput. Chem. 2011, 32, 174–182. [Google Scholar] [CrossRef] [PubMed]
  78. Figgen, D.; Rauhut, G.; Dolg, M.; Stoll, H. Energy-consistent pseudopotentials for group 11 and 12 atoms: Adjustment to multi-configuration Dirac–Hartree–Fock data. Chem. Phys. 2005, 311, 227–244. [Google Scholar] [CrossRef]
  79. Peterson, K.A.; Puzzarini, C. Systematically convergent basis sets for transition metals. II. Pseudopotential-based correlation consistent basis sets for the group 11 (Cu, Ag, Au) and 12 (Zn, Cd, Hg) elements. Theor. Chem. Acc. 2005, 114, 283–296. [Google Scholar] [CrossRef]
  80. Stoll, H.; Metz, B.; Dolg, M. Relativistic energy-consistent pseudopotentials—Recent developments. J. Comput. Chem. 2002, 23, 767–778. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Potential energy curves of the doublet electronic states for the CdBr molecule without the spin–orbit coupling effect.
Figure 1. Potential energy curves of the doublet electronic states for the CdBr molecule without the spin–orbit coupling effect.
Ijms 27 00184 g001
Figure 2. Potential energy curves of the quartet electronic states for the CdBr molecule without the spin–orbit coupling effect.
Figure 2. Potential energy curves of the quartet electronic states for the CdBr molecule without the spin–orbit coupling effect.
Ijms 27 00184 g002
Figure 3. Potential energy curves of the doublet and quartet electronic states for the CdBr molecule with the spin–orbit coupling effect.
Figure 3. Potential energy curves of the doublet and quartet electronic states for the CdBr molecule with the spin–orbit coupling effect.
Ijms 27 00184 g003
Figure 4. Franck–Condon factor (FCF) for the transition between the X2Σ+1/2 and C2Π1/2 electronic states of the CdBr molecule.
Figure 4. Franck–Condon factor (FCF) for the transition between the X2Σ+1/2 and C2Π1/2 electronic states of the CdBr molecule.
Ijms 27 00184 g004
Figure 5. Transition dipole moment curve of the X2Σ+1/2 − C2Π1/2 transition of the CdBr molecule.
Figure 5. Transition dipole moment curve of the X2Σ+1/2 − C2Π1/2 transition of the CdBr molecule.
Ijms 27 00184 g005
Figure 6. Laser cooling scheme of the X2Σ+1/2 − C2Π1/2 transition of the CdBr molecule.
Figure 6. Laser cooling scheme of the X2Σ+1/2 − C2Π1/2 transition of the CdBr molecule.
Ijms 27 00184 g006
Table 1. The dissociation limits of the Λ-S states of the CdBr molecule.
Table 1. The dissociation limits of the Λ-S states of the CdBr molecule.
Atomic States
Cd + Br
Molecular States
CdBr
Energy Separation (cm−1)
Etheo. aEexp. b% Relative Error
Cd (4d105s2, 1S) + Br (4s24p5, 2P°)X2Σ+, (1)2Π 0.000.000.00
Cd (4d105s5p, 3P°) + Br (4s24p5, 2P°)B2Σ+, X2Π, (1)4Δ, (1)4Σ+, (3)2Σ+, (1)2Δ, (1)4Σ, (1)4Π, (2)4Π, (2)4Σ+
(3)2Π, (1)2Σ
26,245.1531,246.33516.0
Cd (4d105s5p, 1P°) + Br (4s24p5, 2P°)(2)2Δ40,452.0543,692.3847.4
Cd (4d105s6s, 3S) + Br (4s24p5, 2P°)(3)4Π, (3)4Σ+45,889.4951,483.98010.9
a Present work.; b J-weighted averaged values, from ref. [40]. Note that the table is energy ordered.
Table 2. Spectroscopic constants of bound Λ-S states (spin-free) of the CdBr molecule. Numbers in parentheses, given in italics, represent (Obs. − Calc.) values based on the only available experimental data [32], shown in bold.
Table 2. Spectroscopic constants of bound Λ-S states (spin-free) of the CdBr molecule. Numbers in parentheses, given in italics, represent (Obs. − Calc.) values based on the only available experimental data [32], shown in bold.
StateRef.Te
(cm−1)
Re
(Å)
ωe
(cm−1)
Be
(cm−1)
α e
(cm−1)
De
(cm−1)
µe
(a.u)
X2Σ+Exp. [32]0-230.5--7258–12,904
This work0 2.468231.4 (−0.9)5.9280.248314,5521.196
Calc. [36]02.52214 (16.5)----
Calc. [37]02.4664233.2 (−2.7)--11,856-
Calc. [38]02.563198 (32.5)--11,1301.395
Calc. [31] 02.564210.5 (20)5.48- -
Calc. [39] *02.4738227.7 (2.8)5.91-12,098-
Calc. [39] **02.5206223.3 (7.2)5.69-14,840-
B2Σ+This work23,033 3.410118.43.104-17,7711.555
Calc. [31] 19,5283.4679119.03.00- -
Calc. [39] *22,2723.2646133.93.39-19,841-
Calc. [39] **21,5733.5386113.22.89-20,405-
C2ΠExp. [32]31,075-254.5----
This work29,340 (1735)2.466268.0 (−13.5)5.9390.767811,4411.013
Calc. [31] 28,850 (2225)2.5522222.1 (32.4)5.53---
Calc. [39] *30,688 (387)2.4462243.2 (11.3)6.04-11,533-
Calc. [39] **30,333 (742)2.4973259.4 (−4.9)5.80-11,937-
(1)4Σ+This work38,853 2.97585.14.0780.518319330.860
Calc. [31] 35,9373.382346.93.13----
Calc. [39] *40,3953.019477.23.97-1774-
(1)4ΔThis work39,683 3.07168.93.8270.512111230.829
Calc. [31] 36,4913.621332.32.75---
Calc. [39] *41,0353.111264.53.74-1209-
(1)2ΔThis work40,220 3.26749.33.3491.80165870.624
Calc. [31] 36,6474.005124.82.26---
Calc. [39] *41,4753.301545.83.32-725-
(1)4ΣThis work40,243 3.19451.03.5380.81235680.733
Calc. [39] *41,5043.226351.53.48-725-
(1)2ΣThis work40,548 3.42829.93.0690.99232630.536
Calc. [31] 36,8034.4471-1.68---
Calc. [39] *41,7303.385838.63.15-483-
(3)4ΠThis work56,280 2.904128.64.2780.276540500.784
Calc. [31] 54,5653.051499.993.88--1.7874
* Evaluated values using MRCI+Q, including the correlated effect of 4d10 electrons of Cd; ** Evaluated values using MRCI+Q with exclusion of correlated effect of 4d10 electrons of Cd.
Table 3. The lowest dissociation limits of possible Ω states of the CdBr molecule obtained by MRCI+Q calculations in comparison with the experimental values.
Table 3. The lowest dissociation limits of possible Ω states of the CdBr molecule obtained by MRCI+Q calculations in comparison with the experimental values.
Atomic StatePossible Ω StatesRelative Energy (cm−1)
This WorkExp. [40]% Relative Error
Cd (1S) + Br (2P3/2)1/2, 3/20.000.000.00
Cd (1S) + Br (2P1/2)1/23580.423685.24 2.8
Cd (3P0) + Br (2P3/2)1/2, 3/225,634.9430,113.990 14.9
Cd (3P1) + Br (2P3/2)5/2, 3/2, 3/2, 1/2, 1/2, 1/226,586.1230,656.087 13.8
Table 4. Spectroscopic constants of bound W states (spin–orbit coupling) of the CdBr. Numbers in parentheses, given in italics, represent (Obs. − Calc.) values based on the only available experimental data [32], shown in bold.
Table 4. Spectroscopic constants of bound W states (spin–orbit coupling) of the CdBr. Numbers in parentheses, given in italics, represent (Obs. − Calc.) values based on the only available experimental data [32], shown in bold.
StateRef.Te
(cm−1)
Re
(Å)
ωe
(cm−1)
Be × 102
(cm−1)
X2Σ+1/2Exp. [32] -230.5-
This work0.02.468230.6 (−0.1)5.92
Calc. [36] 2.52214.0 (16.5)-
Calc. [37] 2.4664233.2 (−2.7)-
Calc. [38] 2.563198 (32.5)-
Calc. [31] 2.564210.5 (20)5.48
Calc. [39] 2.4749226.8 (3.7)-
(1)2Π1/2This work16,7983.68140.32.66
B2Σ+1/2This work23,1433.431118.13.07
Calc. [39]22,2633.2798133.33.36
C2Π1/2Exp. [32]30,300---
This work28,860 (1440)2.475271.15.89
Calc. [39]30,140 (160)2.4582199.85.64
C2Π3/2Exp. [32]31,463-253.8-
This work29,790 (1673)2.469254.0 (−0.2)5.92
Calc. [39]31,148 (315)2.447236.2 (17.6)6.05
(1)4Σ+1/2This work38,2123.0224.033.954
(1)4Σ+3/2This work38,6283.1701.403.619
(1)4Δ1/2This work39,0013.0729.003.825
Table 5. The rotational constants for the ground X2Σ1/2 and C2Π1/2 states with spin–orbit of the CdBr.
Table 5. The rotational constants for the ground X2Σ1/2 and C2Π1/2 states with spin–orbit of the CdBr.
X2Σ1/2
vEv
(cm−1)
Bv × 102
(cm−1)
Dv × 108
(cm−1)
Rmin
(Å)
Rmax
(Å)
0115.195.84621.45052.4172.526
1349.735.83731.51522.3752.569
2581.025.82111.39222.3532.601
3813.335.79001.18812.3362.627
41049.295.75861.41682.3202.652
51283.515.74701.60422.3072.574
61513.975.72771.29462.2952.695
71744.615.69851.31592.2842.714
81974.775.67791.52932.2742.733
92202.705.66141.49432.2652.751
102428.845.63811.29842.2562.769
112654.495.61271.43652.2482.786
122878.745.59481.56062.2402.802
133100.995.57401.37512.2332.819
143322.255.54951.42022.2262.835
153542.245.52901.51582.2202.850
163760.605.50791.45082.2132.866
173977.645.48541.46032.2072.881
184193.315.46371.48042.2022.896
194407.555.44161.48322.1962.911
204620.395.41981.50302.1912.926
C2Π1/2
vEv
(cm−1)
Bv × 102
(cm−1)
Dv × 108
(cm−1)
Rmin
(Å)
Rmax
(Å)
0133.935.85221.08772.4302.534
1403.345.80331.04102.4012.580
2672.205.76020.924322.3812.615
3944.525.72950.772532.3682.639
41223.155.69660.859742.3582.663
51500.685.63661.15762.3482.689
61767.485.57641.05292.3402.714
72029.135.54221.03652.3332.737
82287.835.50881.05482.3262.758
92543.545.47681.12322.3192.779
102795.015.43470.997512.3132.799
113044.115.40351.06332.3082.819
123290.205.36351.17832.3022.838
133532.475.33571.18232.2972.857
Table 6. The radiative lifetimes τ and the vibrational branching ratio of the vibrational transitions between the electronic states X2Σ+1/2 − C2Π 1/2 of the CdBr molecule.
Table 6. The radiative lifetimes τ and the vibrational branching ratio of the vibrational transitions between the electronic states X2Σ+1/2 − C2Π 1/2 of the CdBr molecule.
CdBr X2Σ+1/2 − C2Π1/2
ν′(C2Π1/2) = 0123456
ν (X2Σ+1/2) = 0Aν′v 30,882,256.771,746,179.3385,206.7187127,778.6242,465.01213,967.0074883.3049
Rν′v 0.952152090.028056630.006300630.002126310.000726130.000249380.00020192
ν = 1Aν′v 1,416,739.2455,179,9784,685,587.1431,408,663.4551,980.73209,044.6276,279.25
Rν′v 0.043680460.886601020.076639780.023440950.009438620.003732420.00315406
ν = 2Aν′v 132,692.07585,016,532.945,183,875.217,673,5342,848,681.21,181,868.7537,702.81
Rν′v 0.004091120.080602840.739049830.127691890.048711130.021101840.02223341
ν = 3Aν′v 1537.496581289,325.5410,456,180.6434,819,5148,703,381.13,984,275.91,833,160.8
Rν′v 0.000047400.004648720.171026470.579416140.148823800.071137800.07579917
ν = 4Aν′v 429.583547119.191812397,518.641515,396,83326,303,4057,904,342.24,730,441.4
Rν′v 0.000013240.000000310.006502010.256211890.449776080.141129170.19559851
ν = 5Aν′v 469.05577062659.475622,832.35746519,034.5219,480,61618,272,5687,084,431
Rν′v 0.000014460.000042730.000373460.008637020.333109540.326250080.29293337
ν = 6Aν′v 39.509368262971.68156592.117738148,781.89550,582.3224,441,7919,917,546.6
Rν′v 0.000001220.000047750.000107820.002475810.009414700.436399320.41007956
Sum(s−1) = Aν′v 32,434,163.7362,237,66661,137,792.8260,094,13958,481,11256,007,85724,184,445
τ:(s) = /Aν′v 3.08317 × 10−81.607 × 10−81.63565 × 10−81.664 × 10−81.71 × 10−81.785 × 10−84.135 × 10−8
τ:(ns) 30.8316.0716.3616.6417.1017.8541.35
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

Mostafa, A.; Zeid, I.; Abu El Kher, N.; El-Kork, N.; Korek, M. Fine Structure Investigation and Laser Cooling Study of the CdBr Molecule. Int. J. Mol. Sci. 2026, 27, 184. https://doi.org/10.3390/ijms27010184

AMA Style

Mostafa A, Zeid I, Abu El Kher N, El-Kork N, Korek M. Fine Structure Investigation and Laser Cooling Study of the CdBr Molecule. International Journal of Molecular Sciences. 2026; 27(1):184. https://doi.org/10.3390/ijms27010184

Chicago/Turabian Style

Mostafa, Ali, Israa Zeid, Nariman Abu El Kher, Nayla El-Kork, and Mahmoud Korek. 2026. "Fine Structure Investigation and Laser Cooling Study of the CdBr Molecule" International Journal of Molecular Sciences 27, no. 1: 184. https://doi.org/10.3390/ijms27010184

APA Style

Mostafa, A., Zeid, I., Abu El Kher, N., El-Kork, N., & Korek, M. (2026). Fine Structure Investigation and Laser Cooling Study of the CdBr Molecule. International Journal of Molecular Sciences, 27(1), 184. https://doi.org/10.3390/ijms27010184

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