Laser-induced fluorescence/dispersed fluorescence (LIF/DF) and cavity ring-down spectra of the A1̃2A/A2̃2AX̃2A electronic transition of the calcium ethoxide (CaOC2H5) radical have been obtained under jet-cooled conditions. An essentially constant Ã2Ã1 energy separation for different vibronic levels is observed in the LIF spectrum, which is attributed to both the spin–orbit (SO) interaction and non-relativistic effects. Electronic transition energies, vibrational frequencies, and spin–vibrational eigenfunctions calculated using the coupled-cluster method, along with results from previous complete active space self-consistent field calculations, have been used to predict the vibronic energy level structure and simulate the recorded LIF/DF spectra. Although the vibrational frequencies and Franck–Condon (FC) factors calculated under the Born–Oppenheimer approximation and the harmonic oscillator approximation reproduce the dominant spectral features well, the inclusion of the pseudo-Jahn–Teller (pJT) and SO interactions, especially those between the A1̃2A/A2̃2A and the B̃2A states, induces additional vibronic transitions and significantly improves the accuracy of the spectral simulations. Notably, the spin–vibronic interactions couple vibronic levels and alter transition intensities. The calculated FC matrix for the A1̃2A/A2̃2AX̃2A transition contains a number of off-diagonal matrix elements that connect the vibrational ground levels to the levels of the ν8 (CO stretch), ν11 (OCC bending), ν12 (CaO stretch), ν13 (in-plane CaOC bending), and ν21 (out-of-plane CaOC bending) modes, which are used for vibrational assignments. Transitions to the ν21(a″) levels are allowed due to the pJT effect. Furthermore, when LIF transitions to the Ã-state levels of the CaOC-bending modes, ν13 and ν21, are pumped, A1̃2A/A2̃2AX̃2A transitions to the combination levels of these two modes with the ν8, ν11, and ν12 modes are also observed in the DF spectra due to the Duschinsky mixing. Implications of the present spectroscopic investigation to laser cooling of asymmetric-top molecules are discussed.

Alkaline earth monoalkyl (MR) and monoalkoxide (MOR) free radicals, e.g., CaCH3, CaOCH3, SrCH3, and SrOCH3, are promising candidates for laser-cooling of polyatomic molecules.1,2 In 2020, CaOCH3, a symmetric top, became the first nonlinear molecule to be laser-cooled.3 Most recently, the 225RaOCH3+ cation has been proposed as a candidate molecule for testing fundamental symmetry due to its deformed 225Ra nucleus.4,5 Moreover, the practicality of laser-cooling asymmetric-top molecules has been investigated using ab initio calculated and experimentally determined molecular properties.6 Compared to diatomics and linear molecules, asymmetric-top molecules have certain advantages in optical cycling and laser cooling because they can be readily polarized in the laboratory frame due to the small asymmetry splitting in the K ≠ 0 levels. The fast advance of laser-cooling of molecules and its applications in the search for physics beyond the Standard Model7,8 requires more detailed investigation and quantitative understanding of spin–rovibronic (spin–rotational–vibrational–electronic) energy level structure and transition intensities of metal-containing molecules, both by laser-spectroscopy measurements and by first-principles calculations.

In 2019, we reported the first quantitative spin–vibronic analysis of the lowest electronic states of an alkaline earth monoalkoxide free radical, namely, calcium methoxide (CaOCH3).9 Laser-induced fluorescence (LIF) and dispersed fluorescence (DF) spectra of the Ã2EX̃2A1 electronic transition of CaOCH3 have been recorded under jet-cooled conditions. Vibronic transitions observed in the LIF/DF spectra were assigned unambiguously based on complete active space self-consistent field (CASSCF) and coupled-cluster singles and doubles (CCSD) calculations. Vibrational branching ratios (VBRs) of the Ã2EX̃2A1 transition were determined using the LIF/DF spectra, combined with a more precise intensity measurement in a jet-cooled cavity ring-down (CRD) measurement.

The calculated Franck–Condon (FC) matrix for the Ã2E–X̃2A1 electronic transition of CaOCH3 is highly diagonal. However, a number of weaker lines in the experimental spectra are attributable to vibronic transitions corresponding to off-diagonal FC matrix elements, and their intensities are reproduced by detailed calculations going beyond the Born–Oppenheimer (BO) approximation. The off-diagonal transitions include those from the vibrational ground level of the initial electronic state to the vibrational levels of the asymmetric CaOC stretch mode [ν3(a1)] and, most interestingly, the asymmetric CaOC bending mode [ν8(e)] of the final electronic state. (The numbers in parentheses correspond to the Herzberg convention for labeling the modes. See Table II of Ref. 9.) Transitions from the v = 0 level of an electronic state to the v = 1 level of an asymmetric mode of another electronic state is forbidden under the diabatic BO approximation and the FC approximation but induced by the Jahn–Teller (JT) effect10 of the Ã2E state in this case. A spin–vibronic spectroscopy model reveals that, in addition to the JT effect, the pseudo-Jahn–Teller (pJT) interactions10 between the Ã2E and B̃2A1 states, as well as the spin–orbit (SO) interaction of the Ã2E state,11 enhance the transition to the ν8 = 1 level. Additionally, the Duschinsky interaction12 induces transitions between the CaO-stretch (ν4) mode and non-CaO-stretch modes. Finally, anharmonic vibrational terms in the ground state induce transitions that are forbidden in the harmonic-oscillator approximation. The spectroscopic model we developed accounts for all these complex interactions, which enabled accurate prediction of transition frequencies and intensities.

The JT effect in molecules such as CaOCH3 that belong to the C3v point group is reduced to the pJT effect in molecules of the C2v, Cs, or C1 point groups. The intra-molecular interactions in pJT-active molecules are even more complex than JT-active ones due to the lowered symmetry. Spin–vibronic analysis of Cs or C1 molecules, similar to that done previously for C3v molecules, is necessary for future laser-cooling of asymmetric-top molecules and their applications in fundamental sciences.

A common approach to the spectroscopic investigation of low-symmetry molecules is the asymmetric alkyl substitution of a symmetric-top molecule such as CaOCH3. Using this approach, we have obtained LIF, DF, and CRD spectra of jet-cooled calcium ethoxide (CaOC2H5) radicals and analyzed their spin–vibronic structure. The spectroscopic analysis requires an in-depth understanding and modeling of couplings between the close-lying electronic states originating from the breaking of the degeneracy of the Ã2E state of CaOCH3 into nearly degenerate A1̃2A and A2̃2A states in CaOC2H5. Based on the previous work on CaOCH3, one expects that the spin–vibronic interactions between the A1̃2A/A2̃2A and the B̃2A states, including both the SO and the pJT effects, also affect the vibronic structure of CaOC2H5.

We have developed the methodology for treating pJT and SO interactions for Cs molecules and incorporated it in the SOCJT (Spin–Orbit Coupling Jahn–Teller) program,11 which was used to analyze the CaOCH3 spectra. The upgraded SOCJT program has been employed to analyze and simulate transition frequencies and intensities for the excitation and emission spectra of CaOC2H5 in the present work.

Previously, Bernath and co-workers reported the LIF/DF spectra of the ÃX̃ and B̃X̃ electronic transitions of alkaline earth monoalkoxide radicals (MOR; M = Ca, Sr, and Ba with R being an alkoxide group).13,14 These radicals were produced in a Broida-type oven by the reaction of a metal vapor with an oxidant (alcohol, acetone, or acetaldehyde). Paul et al. reported the DF spectra of CaOC2H5 by exciting its ÃX̃ origin band.15 The results reported in the present paper represent the first comprehensive spectroscopic investigation of the ÃX̃ electronic transition of CaOC2H5.

The rest of this paper is organized as follows: First, the experimental apparatuses for the jet-cooled LIF/DF and CRD measurements are described (Sec. II). Quantum chemistry calculations of related electronic states of the CaOC2H5 radical will then be presented (Sec. III A), followed by the ab initio vibronic calculations of the energy level structure and transition intensities that take into account both the pJT and SO interactions (Sec. III B). The spectral simulation using the calculated results will be compared with the experimentally obtained spectra for vibrational assignment in Sec. IV and discussed in Sec. V. Also to be discussed are the implications of the present work to laser-cooling of asymmetric-top molecules and detecting symmetry-violating interactions.

The LIF/DF and CRD apparatuses were described in our previous publication on the CaOCH3 radical9 and are only briefly described here. In the present work, CaOC2H5 radicals were produced by 1064 nm laser ablation of a calcium metal rod with a Nd:YAG laser (Continuum, Powerlite Precision II 8000) in the throat of a supersonic jet of helium seeded with ethanol. Ethanol was contained in a stainless steel reservoir at room temperature and entrained in the flow by passing high-pressure helium (backing pressure = 180 psi above 1 atm) through the reservoir. The supersonic jet molecular beam was produced by expanding the seeded flow through a pinhole valve (0.3 mm diameter) into the vacuum chamber (stagnation pressure = 20 mTorr). A 12 mm thick Teflon extension with a 1.5 mm diameter orifice at its center was attached to the supersonic nozzle for additional vibrational cooling.

The excitation laser used in the LIF/DF experiment is a pulsed dye laser (Spectra-Physics, Cobra Stretch) pumped by the second harmonic of an Nd:YAG laser (Spectra-Physics, GCR-4). The laser-induced fluorescence was collected by a lens system perpendicular to both the excitation laser beam and the jet expansion. A photomultiplier tube (PMT, Hamamatsu, H10721-01) was used to detect the focused fluorescence in the LIF measurement. The spectral linewidth of the LIF spectrum is ≈0.1 cm−1, limited mainly by the linewidth of the pulsed dye laser (≈0.06 cm−1) and the residual Doppler width. The frequency of the pulsed dye laser was calibrated by a wavemeter (HighFinesse, W7). The accuracy of the vibronic transition frequencies in the LIF spectrum is limited to ≈1 cm−1, mainly by the width of the rotational contour. For the DF experiment, the fluorescence is dispersed by a monochromator (Acton Research, SpectraPro 300i) equipped with an intensified CCD camera (Princeton Instruments, PI-MAX 512). The width of the entrance slit of the monochromator was adjusted to balance the spectral linewidth and the signal-to-noise ratio (S/N). The typical spectral resolution of DF spectra recorded in the present work is ≈20 cm−1. The wavelength of the DF spectra was calibrated using a mercury arc lamp, and the frequency uncertainty is ≈5 cm−1, limited mainly by the linewidth.

Intensities of the recorded LIF transitions were calibrated by the wavelength-dependent sensitivity of the PMT. Intensities of the DF spectra were calibrated by the grating efficiency of the spectrograph and the quantum efficiency of the iCCD camera, both of which are wavelength-dependent. Further details of intensity calibration can be found in our previous work on CaOCH3.9 The origin-band transitions in the LIF experiment are significantly saturated so that an accurate measurement of their transition intensities is impractical. Therefore, pulsed-laser CRD spectra were also recorded to obtain accurate transition intensities. The CRD mirrors (Los Gatos Research, R ≥ 99.995%, center wavelength = 620 nm) were mounted on the two arms of the vacuum chamber to form a ring-down cavity with a length of L = 76 cm. The ring-down mirrors were purged by a nitrogen flow continuously to prevent contamination (mainly by the metal vapor). Transmission of the ring-down beam from a pulse dye laser (Sirah, PrecisionScan) through the cavity was focused onto a PMT (Hamamatsu, H10721-01). The empty-cavity ring-down time (τ0) was about 250 µs. Ring-down signals were recorded at each wavelength with the ablation laser on and off. The “ablation laser off” CRD spectrum is subtracted from the “ablation laser on” spectrum to obtain the absorption spectrum of radicals.

In this section, we describe the computational and theoretical analyses used to assign and understand the observed spectra. There are essentially two levels to these calculations. At the first level (Sec. III A), quantum chemistry calculations for the relevant diabatic electronic states are performed, including the ground electronic X̃2A state, the first (A1̃2A) and the second (A2̃2A) excited electronic state originating from the Ã2E state of CaOCH3, and the third electronic state (B̃2A). These calculations at the first level conform to the BO approximation, as does the subsequent analyses of the vibrational structure within the electronic states. The second-level spin–vibronic calculations (Sec. III B) include interactions neglected in the BO approximation, which improves the quality of spectral simulations and provides insight into the pJT and SO coupling between the lowest electronic states of CaOC2H5.

In a previous work of ours,15 the X̃2A and A1̃2A/A2̃2A states of CaOC2H5 were calculated using the complete active space self-consistent field (CASSCF) method with the cc-PVTZ basis set, employing the Gaussian 09 program package.16 The active space used for the calculations and the electron promotions involved in the excited-state calculations can be found therein. The Ã1 and Ã2 states of CaOC2H5 are nearly degenerate and strongly coupled. Therefore, a state-averaged (SA) calculation with equal weights for the two states was carried out for the geometric optimization, which converges to the minimum of the conical intersection (CI) seam. Notably, the CASSCF calculations failed to predict the symmetry (A′ or A″) of these two nearly degenerate electronic states correctly,15 probably due to the limited active space and the inaccurate weights used in the calculation.

In the present work, the coupled cluster singles and doubles (CCSD) method was implemented with the CFOUR program suite17 to calculate the lowest four electronic states mentioned above. A cc-pVTZ basis set was employed for the CCSD calculation of the X̃2A state. Properties of the A1̃2A and A2̃2A states were calculated using the equation-of-motion electron excitation coupled-cluster theory (EOMEE-CCSD/cc-pVTZ). Both the CASSCF and CCSD methods conform to the BO approximation and calculate the adiabatic states. The rationale for the new CCSD calculations is first to allow cross-checking of the previous results. More importantly, the CCSD results are used directly in the second level of calculation that involved a vibronic treatment (see Sec. III B).

Geometry optimization of all the states provides equilibrium bond lengths and bond angles, as well as rotational constants, which are summarized in Table I, along with values calculated on the CASSCF(3,6)/cc-PVTZ level of theory as reported in Ref. 15. The largest change in geometry upon the Ã1/Ã2X̃ excitation is the reduction of the CaO bond length (≈28 mÅ). Therefore, qualitatively one expects vibronic bands to the excited CaO stretch levels in the LIF/DF spectra to be the strongest non-diagonal transitions.

TABLE I.

Geometric parameters and rotational constants of the X̃ and Ã1/Ã2 states of the CaOC2H5 radical and its A1̃/A2̃X̃ excitation energy calculated using the CASSCF and CCSD methods.

CASSCFaCCSDbEOM-CCSDc
X̃Ã1/Ã2X̃Ã1/Ã2
rCaO (Å) 2.014 1.984 2.115 2.087 
rOC (Å) 1.372 1.375 1.417 1.424 
rCC (Å) 1.523 1.521 1.532 1.530 
∠CaOC (deg) 179 179 179 179 
∠OCC (deg) 112 112 112 112 
A (cm−10.816 0.821 0.778 0.773 
B (cm−10.066 0.067 0.062 0.063 
C (cm−10.063 0.064 0.059 0.059 
ΔEÃX̃ (cm−1)d 15 390  15 869 
CASSCFaCCSDbEOM-CCSDc
X̃Ã1/Ã2X̃Ã1/Ã2
rCaO (Å) 2.014 1.984 2.115 2.087 
rOC (Å) 1.372 1.375 1.417 1.424 
rCC (Å) 1.523 1.521 1.532 1.530 
∠CaOC (deg) 179 179 179 179 
∠OCC (deg) 112 112 112 112 
A (cm−10.816 0.821 0.778 0.773 
B (cm−10.066 0.067 0.062 0.063 
C (cm−10.063 0.064 0.059 0.059 
ΔEÃX̃ (cm−1)d 15 390  15 869 
a

At the CASSCF(3,6)/cc-PVTZ level of theory. Calculated values from Ref. 15.

b

At the CCSD/cc-PVTZ level of theory.

c

At the EOM-CCSD/cc-PVTZ level of theory.

d

Defined as the energy separation between the X̃ state and the center of the A1̃2A and A2̃2A states. Compared to the experimental value of 15 882 cm−1.

Quantitatively, the vibrational assignment of LIF and DF spectra in the present work is guided by the electronic structure calculations. CaOC2H5 has 21 vibrational modes, including 13 a′ modes and 8 a″ modes. Harmonic frequencies of the X̃ and à states calculated using the CASSCF and CCSD methods are summarized in Table II. In most cases, the harmonic frequencies calculated using either the CASSCF or CCSD methods well reproduce the experimentally determined frequencies. In general, the CCSD calculations are more accurate than the CASSCF ones. Notably, they better reproduce the experimental frequencies of the in-plane and out-of-plane CaOC bending modes. The discrepancies between the CCSD-calculated harmonic frequencies are within 20 cm−1 of the experimental values with the exception of the CO stretch mode (ν8), the highest-frequency vibrational mode for which transitions have been observed in the LIF/DF spectra. For the X̃2A and A1̃2A/A2̃2A states, the calculated values are 74 and 40 cm−1 higher than the experimentally determined values, respectively. These discrepancies may partly be attributed to the anharmonicity of the CO stretch mode, which was also observed in the LIF and DF spectra of alkoxy radicals. However, the anharmonicity (2ωeχe) of the CO stretch mode of alkoxy radicals is usually smaller than 10 cm−1.18–21 A more plausible explanation of the discrepancy is, therefore, the vibronic interaction of the high-lying Ã-state ν8 level with the B̃2A state, which lowers the Ã-state ν8 level (see Sec. IV A).

TABLE II.

Harmonic vibrational frequencies of the X̃ and Ã1/Ã2 states of the CaOC2H5 radical calculated using the CASSCF and (EOM-)CCSD methods compared to the experimentally determined values. Vibrational modes are numbered following the Herzberg convention.

X̃Ã1/Ã2
ModeSymmetryCASSCFCCSDExpt.CASSCFEOM-CCSDExpt.Description
ν1 a′ 3198 3118  3202 3136   
ν2 a′ 3141 3064  3145 3052   
ν3 a′ 3080 3001  3096 3010   
ν4 a′ 1658 1556  1658 1520   
ν5 a′ 1610 1526  1609 1495   
ν6 a′ 1557 1444  1558 1426   
ν7 a′ 1514 1415  1517 1393   
ν8 a′ 1294 1244 1170 1299 1210 1170 CO stretch 
ν9 a′ 1174 1123  1178 1109   
ν10 a′ 974 942 916 983 930  CC stretch 
ν11 a′ 514 522 514 558 528 532 OCC bending 
ν12 a′ 367 393 386 411 391 389 CaO stretch 
ν13 a′ 96 97 93 114 90 93 CaOC bending (in-plane) 
ν14 a″ 3201 3124  3206 3141   
ν15 a″ 3084 3002  3104 3045   
ν16 a″ 1597 1516  1598 1480   
ν17 a″ 1431 1342  1437 1304   
ν18 a″ 1289 1207  1291 1177   
ν19 a″ 861 817  863 806   
ν20 a″ 302 308  303 288  Methyl torsion 
ν21 a″ 140 136 129 161 118 130 CaOC bending (out-of-plane) 
X̃Ã1/Ã2
ModeSymmetryCASSCFCCSDExpt.CASSCFEOM-CCSDExpt.Description
ν1 a′ 3198 3118  3202 3136   
ν2 a′ 3141 3064  3145 3052   
ν3 a′ 3080 3001  3096 3010   
ν4 a′ 1658 1556  1658 1520   
ν5 a′ 1610 1526  1609 1495   
ν6 a′ 1557 1444  1558 1426   
ν7 a′ 1514 1415  1517 1393   
ν8 a′ 1294 1244 1170 1299 1210 1170 CO stretch 
ν9 a′ 1174 1123  1178 1109   
ν10 a′ 974 942 916 983 930  CC stretch 
ν11 a′ 514 522 514 558 528 532 OCC bending 
ν12 a′ 367 393 386 411 391 389 CaO stretch 
ν13 a′ 96 97 93 114 90 93 CaOC bending (in-plane) 
ν14 a″ 3201 3124  3206 3141   
ν15 a″ 3084 3002  3104 3045   
ν16 a″ 1597 1516  1598 1480   
ν17 a″ 1431 1342  1437 1304   
ν18 a″ 1289 1207  1291 1177   
ν19 a″ 861 817  863 806   
ν20 a″ 302 308  303 288  Methyl torsion 
ν21 a″ 140 136 129 161 118 130 CaOC bending (out-of-plane) 

FC factors for both absorption and emission transitions were computed with ground- and excited-state vibrational wavefunctions from either the CASSCF or CCSD calculations. The FC calculations were performed using the ezSpectrum software.22 Both the LIF and DF spectra of the A1̃2A/A2̃2AX̃2A transition were simulated under the harmonic oscillator approximation using the calculated vibrational frequencies and FC factors [see Figs. 1(b) and 1(c) and Sec. IV for details].

FIG. 1.

(a) and (a′) The experimentally obtained A1̃2A/A2̃2AX̃2A LIF spectrum of CaOC2H5. Wavenumbers are relative to the center of the origin band at 15 882 cm−1. (b) Simulation using the transition frequencies and intensities calculated at the CASSCF(3,6)/cc-PVTZ level of theory. (c) Simulation at the EOM-CCSD/cc-PVTZ level of theory. (d) The complete spin–vibronic simulation. The spin–vibronic simulation is truncated at 1000 cm−1 (see text).

FIG. 1.

(a) and (a′) The experimentally obtained A1̃2A/A2̃2AX̃2A LIF spectrum of CaOC2H5. Wavenumbers are relative to the center of the origin band at 15 882 cm−1. (b) Simulation using the transition frequencies and intensities calculated at the CASSCF(3,6)/cc-PVTZ level of theory. (c) Simulation at the EOM-CCSD/cc-PVTZ level of theory. (d) The complete spin–vibronic simulation. The spin–vibronic simulation is truncated at 1000 cm−1 (see text).

Close modal

In addition to the quantum chemistry calculations described in Sec. III A, a complete simulation of CaOC2H5 spectra was done from first principles with the values of all spectral parameters derived from the electronic structure calculations, including vibronic and SO interactions. The approach is similar to our previous work9 on CaOCH3 and the prior work on CH3O by Weichman et al.23 However, in both these cases, the molecules are symmetric tops with nominal C3v symmetry and a degenerate electronic state. On the other hand, CaOC2H5 is an asymmetric top with only Cs symmetry and non-degenerate electronic states, and so the previous treatments must be generalized.

The complete spin–vibronic Hamiltonian matrix for the CaOC2H5 radical in a quasi-diabatic basis includes pJT and SO couplings between the Ã1, Ã2, and B̃ electronic states. These terms were neglected in our treatment in Subsection III A because, consistent with the BO approximation, those treatments ignored spin–vibronic terms coupling different eigenstates of the electronic Hamiltonian.

The spin–vibronic Hamiltonian is written in the quasi-diabatic basis containing Ã1, Ã2, and B̃ electronic states,

Ĥ=T̂N+V̂,
(1)

where T̂N is the nuclear kinetic energy and V̂ is the potential energy. The potential energy matrix is expanded in the normal coordinates (qj) of the X̃ state (from here on referred to as R0) and is a complex 6 × 6 matrix in the spin-electronic basis, |Λ⟩ ⊗ |Σ⟩, where Λ and Σ are the projections of the orbital angular momentum (L) and the spin (S) of the electron, respectively,

V̂=VÃ1VpJTÃ1Ã2+iVzÃ1Ã2VpJTÃ1B̃+iVzÃ1B̃0iVxÃ1Ã2iVxÃ1B̃VpJTÃ1Ã2iVzÃ1Ã2VÃ2+ΔEÃ1Ã2VpJTÃ2B̃iVxÃ1Ã20VyÃ2B̃VpJTÃ1B̃iVzÃ1B̃VpJTÃ2B̃VB̃+ΔEÃ1B̃iVxÃ1B̃VyÃ2B̃00iVxÃ1Ã2iVxÃ1B̃VÃ1VpJTÃ1Ã2iVzÃ1Ã2VpJTÃ1B̃iVzÃ1B̃iVxÃ1Ã20VyÃ2B̃VpJTÃ1Ã2+iVzÃ1Ã2VÃ2+ΔEÃ1Ã2VpJTÃ2B̃iVxÃ1B̃VyÃ2B̃0VpJTÃ1B̃+iVzÃ1B̃VpJTÃ2B̃VB̃+ΔEÃ1B̃.
(2)

In the matrix, from left to right, the columns are labeled by the Ã1, Ã2, and B̃ vibronic basis sets with Σ = +1/2 for the first three and correspondingly Σ = −1/2 for the last three. ΔEii (i = Ã1, Ã2, or B̃) is the electronic energy difference between the ith and i′th electronic states at R0. Vi is the diagonal term corresponding to each electronic state and contains the harmonic oscillator terms and the derivative coupling,

Vi=12jωjiqj2+jdjiqj,
(3)

where ωji is the harmonic frequency of the jth mode. The derivative coupling parameter, dji, is non-zero along totally symmetric normal coordinates only and is given by

dji=qjΛi|Vn|ΛiR0,
(4)

where |Λi⟩ is the electronic wave function of the ith state and Vn is the non-relativistic potential energy of the molecule and consists of nuclear–electron, nuclear–nuclear, and electron–electron interactions.

The pJT coupling between the Ã1, Ã2, and B̃ states is given by VpJTii,

VpJTii=jqjΛi|Vn|ΛiR0qj.
(5)

The SO coupling is given as

Vαii=Λi|aαLαSα|ΛiR0,
(6)

where aα is the SO coupling constant and α (= x, y, z) denotes the principal axes. Following our previous work on alkoxy radicals, the principal axis perpendicular to the Cs plane, the c-axis in the case of CaOC2H5, is defined as the y axis.24–26 The x and z axes lie in the Cs plane. The type-I right-hand (Ir) convention is followed here, i.e., the traditional a, b, and c axes correspond to z, x, and y, respectively. For symmetry reasons, VyÃ1Ã2, VyÃ1B̃, VxÃ2B̃, and VzÃ2B̃ vanish.

The CFOUR quantum chemistry package was used to calculate the values for ωji, dji, VpJTii, and Vαii. The ωji and dji values were calculated using finite differences of adiabatic potential energy surfaces. Vαii was calculated using the mean-field approach for EOM-CCSD wavefunctions. Table III contains the calculated derivative coupling constants and pJT coupling constants, while the SO constants between the Ã1, Ã2, and B̃ states are summarized in Table IV.

TABLE III.

Harmonic frequencies (ωji), derivative coupling constants (dji), and pJT coupling constants VpJTii along normal coordinates defined at the minimum of the X̃ state (R0) calculated at the EOM-CCSD/cc-pVTZ level of theory. All values are in cm−1.

Derivative coupling djipJT coupling VpJTii
Mode numberMode symmetryHarmonic frequencyÃ1Ã2B̃Ã1Ã2Ã1B̃Ã2B̃
ν1 a′ 3118.3 −13.22 −4.19 6.35 15.2 
ν2 a′ 3064.3 5.79 9.56 −1.97 6.1 
ν3 a′ 3001.2 −28.59 −28.78 −6.69 12 
ν4 a′ 1555.8 −37.96 −29.42 −26.06 −3 
ν5 a′ 1526 15.91 −20.85 −11.69 −22.6 
ν6 a′ 1443.7 −49.68 −59.14 −24.66 7.3 
ν7 a′ 1415 −32.73 −32.27 −25.84 −10.7 
ν8 a′ 1243.6 −162.86 −163.02 −125.73 −7.2 
ν9 a′ 1123.4 −32.07 −29.73 −34.29 −20.2 
ν10 a′ 942.6 69.16 73.15 73.13 −3.6 
ν11 a′ 521.9 166.04 166.4 184.19 34.8 
ν12 a′ 393.2 140.16 137.87 136.8 −71.7 
ν13 a′ 97.4 −12.87 −8.15 −3.25 −228.6 
ν14 a″ 3123.9 4.2 −16 
ν15 a″ 3001.5 −1 −18.6 
ν16 a″ 1515.6 9.4 −16 
ν17 a″ 1342 −8.4 −11.4 
ν18 a″ 1207.8 −5.9 −5.4 
ν19 a″ 817.2 −3.8 0.2 
ν20 a″ 307.5 −6.5 −53.4 
ν21 a″ 136.4 2.3 −229.6 
Derivative coupling djipJT coupling VpJTii
Mode numberMode symmetryHarmonic frequencyÃ1Ã2B̃Ã1Ã2Ã1B̃Ã2B̃
ν1 a′ 3118.3 −13.22 −4.19 6.35 15.2 
ν2 a′ 3064.3 5.79 9.56 −1.97 6.1 
ν3 a′ 3001.2 −28.59 −28.78 −6.69 12 
ν4 a′ 1555.8 −37.96 −29.42 −26.06 −3 
ν5 a′ 1526 15.91 −20.85 −11.69 −22.6 
ν6 a′ 1443.7 −49.68 −59.14 −24.66 7.3 
ν7 a′ 1415 −32.73 −32.27 −25.84 −10.7 
ν8 a′ 1243.6 −162.86 −163.02 −125.73 −7.2 
ν9 a′ 1123.4 −32.07 −29.73 −34.29 −20.2 
ν10 a′ 942.6 69.16 73.15 73.13 −3.6 
ν11 a′ 521.9 166.04 166.4 184.19 34.8 
ν12 a′ 393.2 140.16 137.87 136.8 −71.7 
ν13 a′ 97.4 −12.87 −8.15 −3.25 −228.6 
ν14 a″ 3123.9 4.2 −16 
ν15 a″ 3001.5 −1 −18.6 
ν16 a″ 1515.6 9.4 −16 
ν17 a″ 1342 −8.4 −11.4 
ν18 a″ 1207.8 −5.9 −5.4 
ν19 a″ 817.2 −3.8 0.2 
ν20 a″ 307.5 −6.5 −53.4 
ν21 a″ 136.4 2.3 −229.6 
TABLE IV.

SO coupling constants for CaOC2H5 calculated using the EOM-CCSD/cc-pVTZ method. All values are in cm−1.

iiVxiiVyiiVziiVii
A1̃2AA2̃2A 13.68 0.0 58.32 59.90 
A1̃2AB̃2A 43.08 0.0 10.20 44.27 
A2̃2AB̃2A 0.0 43.74 0.0 43.74 
iiVxiiVyiiVziiVii
A1̃2AA2̃2A 13.68 0.0 58.32 59.90 
A1̃2AB̃2A 43.08 0.0 10.20 44.27 
A2̃2AB̃2A 0.0 43.74 0.0 43.74 

To the best of our knowledge, none of the present quantum chemistry codes are capable of including both quadratic and SO terms in the potential of a Cs molecule and, hence, the former were excluded from Eq. (2) since they are likely to be small compared to the latter. To make the calculations feasible, we have ignored any coupling between the a′ and a″ modes, thereby making the Hamiltonian block diagonal. Using the Lanczos algorithm, blocks of the Hamiltonian matrix were diagonalized separately for the a′ and a″ modes to obtain eigenvalues. The a′ calculation includes ν8ν13 modes with vmax{ν8, ν9, ν10, ν11, ν12, ν13} = {10, 7, 9, 12, 12, 16}, and the a″ calculation includes ν16ν21 modes with vmax{ν16, ν17, ν18, ν19, ν20, ν21} = {8, 8, 10, 10, 12, 14}. The basis set for the a′ calculation contains ≈15 × 106 basis functions, and the a″ calculation contains ≈11 × 106 basis functions. Calculation of eigenfunctions up to 1000 cm−1 above the mean of the A1̃2A and A2̃2A states was converged. Therefore, in the present work, the simulated LIF spectrum based on the spin–vibronic calculations is truncated at 1000 cm−1 above the origin [dashed vertical line in Fig. 1(d)].

The calculated eigenvalues and eigenfunctions were then used to predict the transition frequencies and intensities of the LIF and DF spectra. The line strengths, Sα, are given by

Sα=Ψi|μα|Ψf2,
(7)

where μα is the electric dipole moment operator and Ψi and Ψf are the spin–vibronic wavefunctions of the initial and final energy levels, respectively, obtained by solving the Hamiltonian given above [Eq. (2)]. In order to obtain S’s without storing the eigenvectors, calculations are done using Ψi as the seed vector in the Lanczos algorithm.23 The calculated electronic transition dipole moments between the ground state and the lowest excited states are listed in Table V.

TABLE V.

Calculated electronic transition dipole moments (in Debye).

Transitionμxμyμzμ
X̃2AA1̃2A 0.0 2.28375 0.0 2.28375 
X̃2AA2̃2A 0.51931 0.0 2.21532 2.27537 
X̃2AB̃2A 1.85519 0.0 0.41683 1.90144 
Transitionμxμyμzμ
X̃2AA1̃2A 0.0 2.28375 0.0 2.28375 
X̃2AA2̃2A 0.51931 0.0 2.21532 2.27537 
X̃2AB̃2A 1.85519 0.0 0.41683 1.90144 

Comparisons between the experimental spectra and simulations based on the methods described in Sec. III allow the assignment of most of the observed vibronic transitions. This interplay of spectroscopic and computational results helps scrutinize the polyatomic asymmetric-top molecule CaOC2H5 as a potential candidate for laser-cooling.

Panels (a) and (a′) of Fig. 1 show the experimentally obtained LIF spectrum of the ÃX̃ transition of CaOC2H5. As indicated in the figure, the experimental LIF spectrum features six doublets, all with a frequency interval of ∼67 cm−1. The experimentally measured SO-free frequencies of the transitions, relative to the origin transition frequencies, are given in Table II. Using the harmonic frequencies given in Table II, we can assign a vibrational mode to each of the doublets [see Fig. 1(a)]. In addition, we label the lower- and higher-frequency peaks as representing the A1̃2AX̃2A and A2̃2AX̃2A transitions, respectively. The doublets are present in the EOM-CCSD electronic structure calculations, although, as we will see later, the A2̃2AA1̃2A separation is more determined by the SO interaction, which is included in the complete spin–vibronic simulation.

The strongest doublet centered at 15 882 cm−1 is assigned to the A1̃2AX̃2A and A2̃2AX̃2A origin transitions. The CASSCF and EOMEE-CCSD calculations predict the origin transition frequency to be 15 390 and 15 869 cm−1, respectively. In our previous work on CaOCH3,9 the corresponding CCSD calculation also predicted the origin transition frequency with excellent accuracy, a predicted value of 15 918 cm−1 compared to the experimental value of 15 925 cm−1.

Three doublets centered at 389, 532, and 1170 cm−1 to the blue of the center of the origin band were observed in the LIF spectrum. Compared to the calculated Ã-state vibrational frequencies (see Table II), these bands can be assigned as transitions to the v′ = 1 levels of the CaO stretch mode (1201), the OCC bending mode (1101), and the CO stretch mode (801), respectively. Two additional doublets are centered nearer to the origin at 93 and 130 cm−1 to the blue. The most reasonable assignments of these doublets are transitions to the v′ = 1 levels of the CaOC bending mode, in-plane (1301) and out-of-plane (2101), respectively.

It is instructive to compare these observations to the three spectral simulations displayed in panels (b)–(d) of Fig. 1. Panel (b) contains the simulation of the LIF spectrum using the SA-CASSCF method. Although this method does predict the Ã-state vibrational frequencies fairly well, it is insufficient to accurately reproduce the experimental spectrum. First, the doublet splitting cannot be predicted using the SA-CASSCF method because the Ã1Ã2 splitting is zero at the CI, and the calculation does not include the SO interaction. Furthermore, intensities of the vibronic bands are not accurately reproduced. The reasons for the intensity inaccuracy will be elucidated below.

Panel (c) of Fig. 1 contains the simulation resulting from the EOM-CCSD electronic structure calculations. This method clearly does better than SA-CASSCF as we now see the doublet structure of the vibrational transitions. However, the splitting of the doublets is only 17 cm−1 compared to 67 cm−1 observed experimentally. This is because, again, the SO coupling within the non-degenerate state vanishes, which affects the magnitude of the predicted doublet splitting.

Panel (d) of Fig. 1 is the complete spin–vibronic simulation of the LIF spectrum from first principles as described in Sec. III B. This calculation includes SO and first-order pJT couplings between the Ã1, Ã2, and B̃ states. This calculation does a quite good job of reproducing the experimentally observed spectrum. The calculated splitting of the origin doublet from first principles is ≈60 cm−1, in rather good agreement with the experimental splitting of 67 cm−1. In addition, the intensities of the various vibronic transitions are well predicted.

Figure 2 shows a zoomed-in plot of the 50–600 cm−1 region of the LIF spectrum. First, we see the 1301 doublet transition predicted at 90 cm−1, compared to the experimental SO-free transition frequency of 93 cm−1. There are also other lines calculated in the 50–300 cm−1 region, the strongest of which is shown in Fig. 2 at 129 cm−1. These are essentially transitions to dark states that gain intensity due to the pJT couplings. Limited by the S/N ratio, these transitions were not observed in the experiment. In the 300–600 cm−1 region, the 1201 and 1101 bands are observed. A complete list of calculated A1̃2A/A2̃2AX̃2A transitions and their intensities is given in Table S.1 of the supplementary material.

FIG. 2.

The low-frequency region of the experimental LIF spectrum (a) and the complete spin–vibronic simulation (b).

FIG. 2.

The low-frequency region of the experimental LIF spectrum (a) and the complete spin–vibronic simulation (b).

Close modal

In the 300–600 cm−1 region, there are also low-intensity transitions observed to the blue of 1101 and the red of 1201 in the experimental spectrum [see Fig. 3(a)]. These transitions may be attributed to hot bands. Figure 3(b) contains a simulation of the LIF spectrum at a vibrational temperature of 75 K. The simulation predicts the 11011311 hot band to the red of the 1101 band, which matches where it is observed in the experimental spectrum. As expected, the 12011311 hot band is also predicted to be lower in frequency than the 1201 transition, but experimentally the observed weak transitions have higher frequencies than the 1201 band. It is possible that higher-order couplings between the ν12 (CaO stretch) and ν13 (in-plane CaOC bending) modes, which are neglected in the present work (see above), cause the 121131 combination levels of the à states to occur at higher frequencies than would otherwise be expected. Unfortunately, there are no methods presently available to calculate such higher-order couplings for quasi-degenerate electronic states and, hence, it is not possible to test our hypothesis.

FIG. 3.

The experimental LIF spectrum (a) and the complete spin–vibronic simulation at a vibrational temperature of 75 K (b).

FIG. 3.

The experimental LIF spectrum (a) and the complete spin–vibronic simulation at a vibrational temperature of 75 K (b).

Close modal

Alternatively, the weak peaks at 517 cm−1, i.e., 15 cm−1 to the red of the 1101 band, can be assigned to the combination band of ν12 (389 cm−1) and ν21 (130 cm−1), as indicated in Fig. 2(a). Such an assignment can be used to explain the DF spectra obtained by pumping the 1201 transition in the LIF spectrum (see Sec. IV D).

The vibrational frequency of the Ã-state CO-stretch (ν8) mode is not predicted in the present ab initio spin–vibronic calculations because the Hamiltonian diagonalization converges up to 1000 cm−1 and the simulation of the A1̃2A/A2̃2AX̃2A transition is truncated as discussed in Sec. III B. The harmonic frequency of the ν8 mode is predicted to be 74 and 40 cm−1 higher than the experimental value by the CASSCF and CCSD calculations, respectively. It is expected that the pJT interaction with the B̃2A state lowers the ν8 levels as it does to other Ã-state vibrational levels observed in the LIF spectrum (e.g., ν11, ν12, ν13) (see Fig. 1). Therefore, inclusion of the spin–vibronic interactions would move the predicted ν8 vibrational frequency closer to the experimentally observed value.

The origin transitions are power-saturated in the LIF measurements with a pulsed laser. To determine the VBRs of the A1̃/A2̃X̃ electronic transitions, we need to know the degree of this saturation, and we turn to jet-cooled pulsed-laser CRD spectroscopy measurements. The LIF and CRD spectra of the origin, 1201, and 1101 bands are compared in Fig. 4. The determined saturation factor for the origin transitions in the LIF experiment is 2.75. Therefore, FC factors for non-origin transitions determined from the LIF spectrum are reduced by a factor of 2.75 relative to the origin transitions, followed by re-normalization. The details for this process are given in Ref. 9.

FIG. 4.

Comparison of LIF (top) and CRD (bottom) spectra of CaOC2H5 that shows the power saturation of the origin transitions in the LIF measurement.

FIG. 4.

Comparison of LIF (top) and CRD (bottom) spectra of CaOC2H5 that shows the power saturation of the origin transitions in the LIF measurement.

Close modal

The relative intensities, i.e., FC factors, of the ÃX̃ transitions so determined are listed in Table VI. The relative intensities of the A1̃2AX̃2A and A2̃2AX̃2A transitions are combined for the total FC factors. The experimental intensities of the origin transitions are estimated to have a relative error of 10% determined from the error of three repeated experimental traces. The main error source is the fluctuation of the free radical concentration in the jet expansion and the necessity to use the CRD calibration to eliminate the effects of saturation. The relative error of non-origin transitions is estimated to be 5%. The uncertainties of the relative intensities, i.e., the FC factors, have been determined using the error propagation method outlined in Sec. S.3 of the supplementary material of Ref. 9.

TABLE VI.

Relative intensities of vibronic transitions in the ÃX̃(v=0) LIF spectrum of CaOC2H5. The total intensity is normalized to unity.

Expt.Calc.
Ã1/Ã2-state vibrational levelRelative energy (cm−1)Ã1X̃Ã2X̃TotalCalc. 1aCalc. 2bCalc. 3c
00 0.460 ± 0.030 0.382 ± 0.029 0.841 ± 0.042 0.906 0.89 0.884 
131 93 0.018 ± 0.001 0.013 ± 0.001 0.031 ± 0.002 0.003  0.011 
211 130 0.025 ± 0.002 0.019 ± 0.001 0.045 ± 0.002    
121 389 0.018 ± 0.001 0.025 ± 0.002 0.044 ± 0.002 0.034 0.043 0.055 
111 532 0.018 ± 0.001 0.014 ± 0.001 0.032 ± 0.002 0.038 0.036 0.045 
81 1170 0.003 ± <0.001 0.005 ± <0.001 0.008 ± <0.001 0.011 0.006 0.006d 
Expt.Calc.
Ã1/Ã2-state vibrational levelRelative energy (cm−1)Ã1X̃Ã2X̃TotalCalc. 1aCalc. 2bCalc. 3c
00 0.460 ± 0.030 0.382 ± 0.029 0.841 ± 0.042 0.906 0.89 0.884 
131 93 0.018 ± 0.001 0.013 ± 0.001 0.031 ± 0.002 0.003  0.011 
211 130 0.025 ± 0.002 0.019 ± 0.001 0.045 ± 0.002    
121 389 0.018 ± 0.001 0.025 ± 0.002 0.044 ± 0.002 0.034 0.043 0.055 
111 532 0.018 ± 0.001 0.014 ± 0.001 0.032 ± 0.002 0.038 0.036 0.045 
81 1170 0.003 ± <0.001 0.005 ± <0.001 0.008 ± <0.001 0.011 0.006 0.006d 
a

At the CASSCF(3,6)/cc-PVTZ level of theory.

b

At the (EOM)-CCSD/cc-PVTZ level of theory.

c

Complete spin–vibronic calculations. The transition intensity for Ã1 and Ã2 components of each vibrational transition is given in Table S.1 of the supplementary material.

d

Fixed to the value in the EOM-CCSD calculations.

Figures 5(a) and 5(c) illustrate the DF spectra obtained by pumping the A1̃2AX̃2A and A2̃2AX̃2A origin transitions, respectively. DF spectra obtained under slightly different experimental conditions were reported in a previous publication of ours.15 In that paper, vibrational assignments were made by comparing the experimental transition frequencies and intensities to those calculated at the CASSCF(3,6)/cc-PVTZ level of theory [see Fig. 5(e)]. The vibrational assignment has been confirmed by the CCSD calculations in the present work [see Figs. 5(b) and 5(d)]. Frequencies and intensities of Ã(v=0)X̃ vibronic transitions calculated on the EOM-CCSD/cc-PVTZ level of theory are listed in Table S.2.

FIG. 5.

(a) and (c) Experimental A1̃2AX̃2A and A2̃2AX̃2A DF spectra of CaOC2H5 obtained by pumping the origin band in the LIF spectrum. Wavenumbers are redshifts relative to the pumped origin transitions. The asterisked peaks in the DF spectra (a) and (c) are due to collision-induced population transfer between the A1̃2A and A2̃2A states.9,15 (b) and (d) The A1̃2AX̃2A and A2̃2AX̃2A spectra simulated with transition frequencies and intensities calculated at the (EOM-)CCSD/cc-PVTZ level of theory. (e) The A1̃2A/A2̃2AX̃2A DF spectrum simulated at the SA-CASSCF(3,6)/cc-PVTZ level of theory.

FIG. 5.

(a) and (c) Experimental A1̃2AX̃2A and A2̃2AX̃2A DF spectra of CaOC2H5 obtained by pumping the origin band in the LIF spectrum. Wavenumbers are redshifts relative to the pumped origin transitions. The asterisked peaks in the DF spectra (a) and (c) are due to collision-induced population transfer between the A1̃2A and A2̃2A states.9,15 (b) and (d) The A1̃2AX̃2A and A2̃2AX̃2A spectra simulated with transition frequencies and intensities calculated at the (EOM-)CCSD/cc-PVTZ level of theory. (e) The A1̃2A/A2̃2AX̃2A DF spectrum simulated at the SA-CASSCF(3,6)/cc-PVTZ level of theory.

Close modal

In the present work, we mainly report the FC factors and VBRs of the DF spectra. It is worth noting that VBRs for the emission spectroscopy are determined by not only FC factors but also the transition frequencies (ν) because the Einstein A coefficient is proportional to ν3. The experimental determination of the FC factors or VBRs for the origin transitions from the DF spectra is difficult due to the contamination of the scattering of the excitation laser. The origin-transition FC factors are, therefore, fixed to the values determined in the LIF and CRD experiments (see Table VI). FC factors of the other transitions are determined by the maintenance of their experimentally determined intensity ratios scaled by ν3 while keeping the total FC factor normalized. VBRs are calculated by scaling the FC factors of all observed DF peaks by ν3 followed by re-normalization. The VBRs for the emission so determined are listed in Table VII.

TABLE VII.

Relative intensities of vibronic transitions in the Ã(v=0)X̃ DF spectrum of CaOC2H5. The intensities are normalized to unity.

X̃2AExpt.Calc. 1aCalc. 2bCalc. 3c
-stateRelativeÃ12AX̃2AÃ22AX̃2AAveragedÃ12AX̃2AÃ22AX̃2AAveraged
vibrational energyFCFCFCFCFC
level(cm−1)VBRVBRVBRFactorVBRFactorVBRFactorVBRFactorVBRFactorVBR
00 0.858 ± 0.013 0.845 ± 0.014 0.855 ± 0.009 0.918 0.928 0.891 0.901 0.877 0.888 0.882 0.894 0.879 0.890 
131 93 0.007 ± 0.001 0.007 ± 0.001 0.003 ± <0.001 0.003 0.003   0.014 0.014 0.007 0.007 0.011 0.011 
121 386 0.059 ± 0.006 0.062 ± 0.006 0.061 ± 0.001 0.034 0.032 0.081 0.076 0.055 0.052 0.055 0.052 0.055 0.052 
111 514 0.058 ± 0.006 0.067 ± 0.006 0.062 ± 0.001 0.038 0.035 0.019 0.017 0.044 0.040 0.045 0.041 0.044 0.040 
101 916 0.007 ± 0.001 0.007 ± 0.001 0.007 ± <0.001 0.002 0.002 0.001 0.001       
81 1170 0.012 ± 0.001 0.013 ± 0.001 0.013 ± 0.001 0.001 0.001 0.006 0.005 0.008 0.006 0.008 0.006 0.008 0.006 
X̃2AExpt.Calc. 1aCalc. 2bCalc. 3c
-stateRelativeÃ12AX̃2AÃ22AX̃2AAveragedÃ12AX̃2AÃ22AX̃2AAveraged
vibrational energyFCFCFCFCFC
level(cm−1)VBRVBRVBRFactorVBRFactorVBRFactorVBRFactorVBRFactorVBR
00 0.858 ± 0.013 0.845 ± 0.014 0.855 ± 0.009 0.918 0.928 0.891 0.901 0.877 0.888 0.882 0.894 0.879 0.890 
131 93 0.007 ± 0.001 0.007 ± 0.001 0.003 ± <0.001 0.003 0.003   0.014 0.014 0.007 0.007 0.011 0.011 
121 386 0.059 ± 0.006 0.062 ± 0.006 0.061 ± 0.001 0.034 0.032 0.081 0.076 0.055 0.052 0.055 0.052 0.055 0.052 
111 514 0.058 ± 0.006 0.067 ± 0.006 0.062 ± 0.001 0.038 0.035 0.019 0.017 0.044 0.040 0.045 0.041 0.044 0.040 
101 916 0.007 ± 0.001 0.007 ± 0.001 0.007 ± <0.001 0.002 0.002 0.001 0.001       
81 1170 0.012 ± 0.001 0.013 ± 0.001 0.013 ± 0.001 0.001 0.001 0.006 0.005 0.008 0.006 0.008 0.006 0.008 0.006 
a

At the CASSCF(3,6)/cc-PVTZ level of theory.

b

At the (EOM)-CCSD/cc-PVTZ level of theory.

c

Complete spin–vibronic calculations.

DF spectra obtained by pumping strong non-origin vibronic bands in the LIF spectrum, namely, 1301, 2101, 1201, 1101, and 801, are represented in Figs. 6(b)6(f) and compared to those obtained by pumping the origin band [Fig. 6(a)]. As a general rule (but with exceptions), DF transitions from a vibronic level of the à state to those X̃-state vibrational levels of the pumped mode gain intensity because of favorable FC factors. Additionally, because of the Duschinsky mixing,12 transitions to combination levels of the pumped mode and other modes may also gain intensities. Following these principles and guided by the ab initio calculated vibrational frequencies and transition intensities, the assignment of most peaks observed in the DF spectra is straightforward (see Fig. 6). Individual DF spectra are discussed below.

FIG. 6.

DF spectra of CaOC2H5 obtained by pumping different vibronic bands in its A1̃2A/A2̃2AX̃2A LIF spectrum. Wavenumbers are redshifts relative to that of the pump laser.

FIG. 6.

DF spectra of CaOC2H5 obtained by pumping different vibronic bands in its A1̃2A/A2̃2AX̃2A LIF spectrum. Wavenumbers are redshifts relative to that of the pump laser.

Close modal

When the 1301 transition in the LIF spectrum is pumped, the DF spectrum [Fig. 6(b)] is dominated by transitions to the ν13 fundamental level of the X̃2A state and combination levels of the ν13 mode with ν12, ν11, and ν8. These ÃX̃ transitions gain intensities through the Duschinsky mixing between ν13 and the other three FC-favored modes.

Similarly, when the 2101 transition in the LIF spectrum is pumped, transitions to the fundamental level of ν21 and combination levels of 121211, 111211, and 81211 of the X̃2A state are observed in the DF spectrum, although the transitions to the combination levels are quite weak [see Fig. 6(c)]. Surprisingly, a strong transition to the 111131 level is observed in the DF spectrum obtained by pumping the 2101 transition. The ν13 and ν21 modes are the in-plane and out-of-plane CaOC bending modes, respectively. Both modes are pJT-active. The 211 and 111131 levels have the a″ and a′ vibrational symmetries, respectively. Therefore, the transition between these two levels (11101310 and 2101) is FC-forbidden under the BO approximation but allowed with the pJT interaction that mixes the 211 level of the à state with its (totally symmetric) vibrational ground level.

When the 1201 transitions in the LIF spectrum are pumped, the only strong transitions in the DF spectra are those to the 121 levels of the X̃2A state [see Fig. 6(d)].

The DF spectra obtained by pumping the 1101 (OCC bending) transitions [Fig. 6(e)] deserve special attention. One would expect strong transitions to the 111 level of the X̃2A state in the DF spectra. However, the strong peaks labeled “X1” and “X2” in Fig. 6(e) do not match the energies of the 111 level of the X̃2A state but are redshifted from the 111 level by ∼12 cm−1. We recall that the 111 levels of the à state are only ∼15 cm−1 above its 121211 combination levels as determined in the LIF spectrum (see Sec. IV A). Therefore, we attribute the aforementioned redshift of peaks “X1” and “X2” from the 111 level to collision-induced relaxation from the Ã-state 111 level to its 121211 level following the LIF excitation. After this non-radiative decay, the excited-state molecules decay to the nearly degenerate 111 (514 cm−1) and 121211 (386 + 129 cm−1) levels of the X̃2A state. The overall redshift for the DF transition compared to the pump frequency is, therefore, the energy difference between the 111 and 121211 levels of the à state (15 cm−1) combined with the vibrational energy of the 111 and 121211 level of the X̃2A state, hence the extra redshift with respect to the 111 level.

Two other strong bands, labeled “Y1” and “Y2” in Fig. 6(e), are observed in the DF spectra obtained by pumping the 1101 band in the LIF spectrum. The “Y1” band appears at 1160 cm−1 redshift when the transition to the upper SO component (ν00 + 568 cm−1) is pumped, while “Y2” appears at 1130 cm−1 redshift when the transition to the lower SO component (ν00 + 502 cm−1) is pumped. These two transitions lie significantly higher than the expected transitions to the 112 level (fundamental frequency = 514 cm−1) and are not lined up with strong transition around 1170 cm−1, assigned to the 81 transitions (see Fig. 6). Due to the near degeneracy between the 112 and 122212 levels, one expects strong Fermi-resonance between these vibrational levels. However, the vibrational assignment of these two peaks is not possible without high-resolution spectra or further spin–vibronic calculations.

Finally, when the 801 transitions in the LIF spectrum are pumped, the only transitions observed in the DF spectra are those to the 81 levels of the X̃2A state [see Fig. 6(f)].

The vibrational structure of the recorded DF spectra can be better illustrated by blueshifting them by the ground-state frequencies of the pumped modes. The resulting spectra are displayed in Fig. 7. The shifted spectra show that the CaO stretch mode (ν12), the OCC bending mode (ν11), and the CaO stretch mode (ν8) are mixed with both pJT-active modes, the in-plane CaOC bending (ν13) and the out-of-plane CaOC bending (ν21), although mixing with ν21(a″) is significantly weaker.

FIG. 7.

DF spectra of CaOC2H5 obtained by pumping different vibronic bands in its A1̃2A/A2̃2AX̃2A LIF spectrum. Wavenumbers are redshifts relative to that of the pump lase, and all DF spectra obtained by pump non-origin vibronic bands are blueshifted by the X̃-state vibrational frequency of the pumped mode.

FIG. 7.

DF spectra of CaOC2H5 obtained by pumping different vibronic bands in its A1̃2A/A2̃2AX̃2A LIF spectrum. Wavenumbers are redshifts relative to that of the pump lase, and all DF spectra obtained by pump non-origin vibronic bands are blueshifted by the X̃-state vibrational frequency of the pumped mode.

Close modal

The present work on the calcium ethoxide radical provides the first comprehensive experimental and computational study of the spin–vibronic structure of an alkaline earth monoalkoxide radical of Cs symmetry. It is a natural extension of our previous spectroscopic investigation of the calcium methoxide radical.9 The most significant difference between these two radicals is that, in CaOCH3, the à state is split by only the SO interaction and retains its 2E symmetry, while in CaOC2H5, it splits into two nearly degenerate states with different symmetries, A1̃2A and A2̃2A, and the separation between these two states (ΔEÃ2Ã1) is due to both the SO interaction and the non-relativistic effects (ΔE0). The non-relativistic effects separating these two electronic states include the “difference potential”27 (ε2) and the difference between their zero-point energies (ΔZPE).28–30 The overall separation of the A1̃2A and A2̃2A states can be calculated as26 

ΔEÃ2Ã1=ΔE02+SO2=ε2+ΔZPE2+SO2.
(8)

The spin–vibronic calculations of the Ã1 and Ã2 states based on the CCSD calculations predict the splitting between the vibrationless levels of these two states to be 62 cm−1, which is rather close to the observed value. The calculated value, 62 cm−1, is the root-sum-square of the non-relativistic separation between the Ã1 and Ã2 states (17 cm−1) and the SO interaction (60 cm−1) [see Eq. (8)]. The present work, therefore, exemplifies the importance of the SO interaction in the study of low-symmetry alkaline earth monoalkoxides and similar free radicals subject to both the SO and pJT interactions.

The SO splitting and the non-relativistic splitting cannot be determined independently in the current spin–vibronic analysis of the experimental LIF spectrum because both effects separate the A1̃2A and A2̃2A states. However, simulation and fitting of future rotationally resolved A1̃2A/A2̃2AX̃2A LIF spectrum using continuous-wave (cw) lasers or laser amplifiers can be used to differentiate and quantize these two mechanisms because they affect the rotational and fine structure differently.26 This has been demonstrated in our previous work on alkoxy radicals with Cs symmetry,24,25,31,32 which have nearly degenerate X̃ and à states with A′ and A″ symmetries that are coupled by both SO and pJT effects. Analysis of the rotational and fine structure of experimentally observed B̃Ã/X̃ vibronic transitions unraveled the interplay between the pJT interaction and the SO interaction. Furthermore, the magnitude of the mixing between the A′ and A″ states was determined in fitting the B̃à and B̃X̃ spectra simultaneously. Recently, the spectroscopic model for analyzing nearly degenerate electronic states of open-shell molecules has been extended to the case of C1 molecules26 and implemented in the analysis of large straight-chain alkoxy radicals.33–35 

In the present work, vibrational assignments were made by comparing experimental spectra to the SA-CASSCF and (EOM-)CCSD calculated harmonic vibrational frequencies and FC factors. In general, the EOM-CCSD calculations can target each of the two excited states (A1̃2A and A2̃2A) and provide significantly more accurate frequency and intensity values compared to the experiment than the SA-CASSCF calculations (see Table II). However, there is still a considerable discrepancy between the calculated Ã-state vibrational frequencies and experimentally determined ones [see Figs. 1(a) and 1(c)].

Introducing the pJT couplings between the A1̃2Aand A2̃2A states improved the calculated vibronic structure significantly so that the predicted ÃX̃ transition frequencies match the experimental ones very well. Furthermore, as described in Sec. V A, the inclusion of the Ã2-Ã1 SO interaction reproduces the observed splitting of the two states quite well (62 cm−1 vs 67 cm−1 in experiment). Including the pJT and SO couplings with the B̃2A state slightly improves the quality of the vibronic simulation in terms of SO-free transition frequencies and intensities. In addition, the inclusion of coupling with the B̃ state in the calculation quenches the SO splitting of the zero-point levels by ≈5%, to 60 cm−1. The simulated LIF spectrum is compared to the experimental spectrum in Fig. 1(d) vs Fig. 1(a)′. Tables S.1 and S.2 list the frequencies and intensities of the A1̃2A/A2̃2AX̃2A vibronic transitions calculated excluding and including the pJT and SO couplings with the B̃ state, respectively. There are quite a few lines in the 50–250 cm−1 region that, while remaining weak, do gain some intensity due to the coupling with the B̃ state as can be seen by comparing Tables S.1 and S.2. The enhancement of transition intensities of “dark states” in this low-frequency region is mainly attributed to strong B̃2A-A2̃2A pJT interaction via the ν13 mode, which has a calculated Ã-state harmonic frequency of 90 cm−1 (see Table III).

Transitions to levels of the ν21 (out-of-plan CaOC bending) mode of CaOC2H5 is FC-forbidden under the BO approximation because of its a″ symmetry. Such transitions gain intensities from the pJT coupling as well as the SO interaction. Previously, transitions to the levels of the out-of-plane mode (ν8) of CaOCH3 were also observed in its LIF/DF spectra. Similarly, FC-forbidden vibronic transitions have also been observed in the B̃Ã/X̃ transitions of alkoxy radicals with Cs symmetry, including isopropoxy32 and cyclohexoxy.25,31 As demonstrated therein, analysis and simulation of the rotational and fine structure of experimentally obtained spectra enable the determination of the pJT and SO coupling strengths.

The excitation scheme of laser-cooling molecules is mainly determined by FC factors and VBRs. A highly diagonal FC matrix for vibronic transitions between the involved electronic states is preferred to avoid population leakage and achieve the number of photon scatterings necessary for laser cooling.36 Previously, FC factors for the Ã-X̃ vibronic transitions of CaOC2H5 were calculated using the density functional theory method on the B3LYP/def2-TZVPP level of theory.6 FC factors of the origin transition and the transition from the vibrational ground level of the à state to the v″ = 1 level of the CaO stretch mode (1201) were found to be 0.892 and 0.102, respectively. These values are consistent with the calculated results without vibronic interactions reported in the present work (See Table VII). As demonstrated in the present calculations, spin–vibronic interactions induce off-diagonal FC matrix elements with considerable magnitudes, which have to be taken into account in future direct laser cooling experiments.

In the future laser-cooling experiment, pre-cooled CaOC2H5 molecules are excited from the X̃ state to the Ã1 state. Following the excitation of the origin transition, spontaneous emission to excited vibrational levels of the ground electronic state occurs due to off-diagonal transitions, causing loss of population from the cooling cycle. Therefore, a reasonable number of repump lasers are used to send the population on X̃-state vibrational levels back to the cooling cycle and limit the loss of population for laser cooling. Based on the spectroscopic investigation of the present work, in addition to the spontaneous emission to the vibrational ground level of the X̃ state, VBRs for transitions to both the 111 (OCC bending) and 121 (CaO stretch) levels are also significant. Two repumping lasers are, therefore, needed to return the population from these two ground-state vibrational levels to the cooling cycle. The combined VBR of the spontaneous decay to the 00, 111, and 121 levels of the X̃2A state is 0.975. As a result, a molecule will experience on average ≈40 scattering events before it decays to vibrational “dark” states that are not addressed by the repumping lasers. If a third repumping laser is added to return the population on the 81 level of the X̃2A state to the cooling cycle, the combined VBR will be increased to 0.987 so that a molecule will experience on average 77 scattering events before decaying to vibrational dark states.

We report vibrationally resolved LIF, DF, and CRD spectra of the A1̃2A/A2̃2AX̃2A transition of the calcium ethoxide radical. The vibrational assignment has been made based on vibrational frequencies and FC factors calculated using the CASSCF and CC methods. The calculations predict the vibrational frequencies to a significant degree of accuracy and the intensities of the allowed transitions quasi-quantitatively. However, the ab initio calculations do not reproduce weak transitions in the LIF spectrum induced by the pJT effect, including transitions to the ν21 (out-of-plane CaOC bending with the a″ symmetry) level and combination levels of this pJT-active mode. In the DF spectra obtained by pumping the origin band, strong transitions to the ν12 (CaO stretch) and ν11 (OCC bending) levels of the X̃2A state were observed, with weaker transitions to a few other vibrational levels. When vibronic bands other than the origin are pumped, the obtained DF spectra contain off-diagonal transitions attributed to the Duschinsky rotation. In the case of the DF spectra obtained by the 1101 transitions, a non-radiative decay to the nearby 121211 combination level precedes the ÃX̃ radiative decay.

The current spin–vibronic calculation does not include quadratic coupling terms and is truncated at 1000 cm−1 due to convergence issues. Development of quantum chemistry codes that include both the quadratic coupling terms and SO interaction is expected to help further improve the accuracy of spin–vibronic calculations on CaOC2H5 and other open-shell molecules with Cs symmetry. High-resolution spectra with resolved rotational and fine structure are desired for experimental determination of the SO and non-relativistic interactions that separate the A1̃2A and A2̃2A states. They will also help confirm certain tentative vibrational assignments in the present work.

Relative intensities of A1̃2A/A2̃2AX̃2A and A1̃2A/A2̃2AX̃2A vibronic transitions were determined in LIF and DF spectra, respectively. However, the excitation transitions of the origin band are power saturated in the current LIF measurement using a pulsed dye laser, while the fluorescence signal in the DF measurement of the origin transitions is contaminated by the laser scattering. Therefore, CRD measurements were performed to determine the saturation factor. Future LIF/DF and CRD spectroscopy measurements with cw lasers can determine the FC factors and VBRs with better accuracy. The S/N ratio of the current experiment is mainly limited by the shot-to-shot fluctuation and long-term variation of the concentration of free radicals produced by laser ablation. In future experiments, the S/N ratio can be improved by monitoring the signals of the origin transition and the vibronic transitions simultaneously, followed by intensity normalization.

See the supplementary material for frequencies and intensities of the A1̃2A/A2̃2AX̃2A vibronic transitions calculated excluding and including the pJT and SO couplings with the B̃ state.

This work was supported by the National Science Foundation under Grant Nos. CHE-1454825 and CHE-1955310. K.S. gratefully acknowledges a Terry A. Miller Postdoctoral Fellowship from the Ohio State University. T.A.M. acknowledges support from the Ohio Supercomputer via Project No. PAS0540.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
T. A.
Isaev
and
R.
Berger
,
Phys. Rev. Lett.
116
,
063006
(
2016
).
2.
I.
Kozyryev
,
L.
Baum
,
K.
Matsuda
, and
J. M.
Doyle
,
ChemPhysChem
17
,
3641
(
2016
).
3.
D.
Mitra
,
N. B.
Vilas
,
C.
Hallas
,
L.
Anderegg
,
B. L.
Augenbraun
,
L.
Baum
,
C.
Miller
,
S.
Raval
, and
J. M.
Doyle
,
Science
369
,
1366
(
2020
).
4.
M.
Fan
,
C. A.
Holliman
,
X.
Shi
,
H.
Zhang
,
M. W.
Straus
,
X.
Li
,
S. W.
Buechele
, and
A. M.
Jayich
,
Phys. Rev. Lett.
126
,
023002
(
2021
).
5.
P.
Yu
and
N. R.
Hutzler
,
Phys. Rev. Lett.
126
,
023003
(
2021
).
6.
B. L.
Augenbraun
,
J. M.
Doyle
,
T.
Zelevinsky
, and
I.
Kozyryev
,
Phys. Rev. X
10
,
031022
(
2020
).
7.
T.
Steimle
and
W.
Ubachs
,
J. Mol. Spectrosc.
300
,
1
(
2014
).
8.
M. S.
Safronova
,
D.
Budker
,
D.
DeMille
,
D. F. J.
Kimball
,
A.
Derevianko
, and
C. W.
Clark
,
Rev. Mod. Phys.
90
,
025008
(
2018
).
9.
A. C.
Paul
,
K.
Sharma
,
M. A.
Reza
,
H.
Telfah
,
T. A.
Miller
, and
J.
Liu
,
J. Chem. Phys.
151
,
134303
(
2019
).
10.
I. B.
Bersuker
,
The Jahn-Teller Effect
(
Cambridge University Press
,
Cambridge, UK
,
2006
).
11.
T. A.
Barckholtz
and
T. A.
Miller
,
Int. Rev. Phys. Chem.
17
,
435
(
1998
).
12.
F.
Duschinsky
,
Acta Physicochim. URSS
7
,
551
(
1937
).
13.
C. R.
Brazier
,
L. C.
Ellingboe
,
S.
Kinsey-Nielsen
, and
P. F.
Bernath
,
J. Am. Chem. Soc.
108
,
2126
(
1986
).
14.
P. F.
Bernath
,
Science
254
,
665
(
1991
).
15.
A. C.
Paul
,
M. A.
Reza
, and
J.
Liu
,
J. Mol. Spectrosc.
330
,
142
(
2016
).
16.
M. J.
Frisch
,
G. W.
Trucks
,
H. B.
Schlegel
,
G. E.
Scuseria
,
M. A.
Robb
,
J. R.
Cheeseman
,
G.
Scalmani
,
V.
Barone
,
B.
Mennucci
,
G. A.
Petersson
,
H.
Nakatsuji
,
M.
Caricato
,
X.
Li
,
H. P.
Hratchian
,
A. F.
Izmaylov
,
J.
Bloino
,
G.
Zheng
,
J. L.
Sonnenberg
,
M.
Hada
,
M.
Ehara
,
K.
Toyota
,
R.
Fukuda
,
J.
Hasegawa
,
M.
Ishida
,
T.
Nakajima
,
Y.
Honda
,
O.
Kitao
,
H.
Nakai
,
T.
Vreven
,
J. A.
Montgomery
, Jr.
,
J. E.
Peralta
,
F.
Ogliaro
,
M.
Bearpark
,
J. J.
Heyd
,
E.
Brothers
,
K. N.
Kudin
,
V. N.
Staroverov
,
R.
Kobayashi
,
J.
Normand
,
K.
Raghavachari
,
A.
Rendell
,
J. C.
Burant
,
S. S.
Iyengar
,
J.
Tomasi
,
M.
Cossi
,
N.
Rega
,
J. M.
Millam
,
M.
Klene
,
J. E.
Knox
,
J. B.
Cross
,
V.
Bakken
,
C.
Adamo
,
J.
Jaramillo
,
R.
Gomperts
,
R. E.
Stratmann
,
O.
Yazyev
,
A. J.
Austin
,
R.
Cammi
,
C.
Pomelli
,
J. W.
Ochterski
,
R. L.
Martin
,
K.
Morokuma
,
V. G.
Zakrzewski
,
G. A.
Voth
,
P.
Salvador
,
J. J.
Dannenberg
,
S.
Dapprich
,
A. D.
Daniels
,
Ö.
Farkas
,
J. B.
Foresman
,
J. V.
Ortiz
,
J.
Cioslowski
, and
D. J.
Fox
, Gaussian 09 Revision A.1,
Gaussian, Inc.
,
Wallingford, CT
,
2009
.
17.
J. F.
Stanton
,
J.
Gauss
,
L.
Cheng
,
M. E.
Harding
,
D. A.
Matthews
, and
P. G.
Szalay
, CFOUR, coupled-cluster techniques for computational chemistry, a quantum-chemical program package, with contributions from
A. A.
Auer
,
R. J.
Bartlett
,
U.
Benedikt
,
C.
Berger
,
D. E.
Bernholdt
,
Y. J.
Bomble
,
O.
Christiansen
,
F.
Engel
,
R.
Faber
,
M.
Heckert
,
O.
Heun
,
M.
Hilgenberg
,
C.
Huber
,
T.-C.
Jagau
,
D.
Jonsson
,
J.
Jusélius
,
T.
Kirsch
,
K.
Klein
,
W. J.
Lauderdale
,
F.
Lipparini
,
T.
Metzroth
,
L. A.
Mück
,
D. P.
O’Neill
,
D. R.
Price
,
E.
Prochnow
,
C.
Puzzarini
,
K.
Ruud
,
F.
Schiffmann
,
W.
Schwalbach
,
C.
Simmons
,
S.
Stopkowicz
,
A.
Tajti
,
J.
Vázquez
,
F.
Wang
, and
J. D.
Watts
, the integral packages MOLECULE,
J.
Almlöf
and
P. R.
Taylor
, PROPS,
P. R.
Taylor
, ABACUS,
T.
Helgaker
,
H. J. Aa.
Jensen
,
P.
Jørgensen
, and
J.
Olsen
, and ECP routines by
A. V.
Mitin
and
C.
van Wüllen
, for the current version, see http://www.cfour.de.
18.
D. E.
Powers
,
M. B.
Pushkarsky
, and
T. A.
Miller
,
J. Chem. Phys.
106
,
6863
(
1997
).
19.
C. C.
Carter
,
J. R.
Atwell
,
S.
Gopalakrishnan
, and
T. A.
Miller
,
J. Phys. Chem. A
104
,
9165
(
2000
).
20.
J.
Jin
,
I.
Sioutis
,
G.
Tarczay
,
S.
Gopalakrishnan
,
A.
Bezant
, and
T. A.
Miller
,
J. Chem. Phys.
121
,
11780
(
2004
).
21.
J.
Liu
,
N. J.
Reilly
,
A.
Mason
, and
T. A.
Miller
,
J. Phys. Chem. A
119
,
11804
(
2015
).
22.
V. A.
Mozhayskiy
and
A. I.
Krylov
, ezspectrum, http://iopenshell.usc.edu/downloads.
23.
M. L.
Weichman
,
L.
Cheng
,
J. B.
Kim
,
J. F.
Stanton
, and
D. M.
Neumark
,
J. Chem. Phys.
146
,
224309
(
2017
).
24.
J.
Liu
,
D.
Melnik
, and
T. A.
Miller
,
J. Chem. Phys.
139
,
094308
(
2013
).
25.
J.
Liu
and
T. A.
Miller
,
J. Phys. Chem. A
118
,
11871
(
2014
).
26.
J.
Liu
,
J. Chem. Phys.
148
,
124112
(
2018
).
27.
P. D. A.
Mills
,
C. M.
Western
, and
B. J.
Howard
,
J. Chem. Phys.
90
,
3331
(
1986
).
28.
L.
Yu
,
D. W.
Cullin
,
J. M.
Williamson
, and
T. A.
Miller
,
J. Chem. Phys.
98
,
2682
(
1993
).
29.
D.
Melnik
,
J.
Liu
,
R. F.
Curl
, and
T. A.
Miller
,
Mol. Phys.
105
,
529
(
2007
).
30.
D. G.
Melnik
,
J.
Liu
,
M.-W.
Chen
,
T. A.
Miller
, and
R. F.
Curl
,
J. Chem. Phys.
135
,
094310
(
2011
).
31.
L.
Zu
,
J.
Liu
,
G.
Tarczay
,
P.
Dupré
, and
T. A.
Miller
,
J. Chem. Phys.
120
,
10579
(
2004
).
32.
R.
Chhantyal-Pun
,
M.
Roudjane
,
D. G.
Melnik
,
T. A.
Miller
, and
J.
Liu
,
J. Phys. Chem. A
118
,
11852
(
2014
).
33.
Y.
Yan
,
K.
Sharma
,
T. A.
Miller
, and
J.
Liu
,
J. Chem. Phys.
153
,
174306
(
2020
).
34.
J.
Liu
and
T. A.
Miller
,
J. Phys. Chem. A
125
,
1391
(
2021
).
35.
J.
Liu
,
M.-W.
Chen
, and
T. A.
Miller
,
J. Phys. Chem. A
125
,
1402
(
2021
).
36.
M. D.
Di Rosa
,
Eur. Phys. J. D
31
,
395
(
2004
).

Supplementary Material