Relativistic coupled-cluster calculations of RaOH pertinent to spectroscopic detection and laser cooling

Chaoqun Zhang a, Phelan Yu b, Chandler J. Conn b, Nicholas R. Hutzler *b and Lan Cheng *a
aDepartment of Chemistry, The Johns Hopkins University, Baltimore, MD 21218, USA. E-mail: lcheng24@jhu.edu
bDivision of Physics, Mathematics, and Astronomy, California Institute of Technology, Pasadena, CA 91125, USA. E-mail: hutzler@caltech.edu

Received 23rd August 2023 , Accepted 17th November 2023

First published on 24th November 2023


Abstract

A relativistic coupled-cluster study of the low-lying electronic states in the radium monohydroxide molecule (RaOH), a radioactive polyatomic molecule of interest to laser cooling and to the search of new physics beyond the Standard Model, is reported. The level positions of the A2Π1/2 and C2Σ states have been computed with an accuracy of around 200 cm−1 to facilitate spectroscopic observation of RaOH using laser induced fluorescence spectroscopy, thereby exploiting the systematic convergence of electron-correlation and basis-set effects in relativistic coupled-cluster calculations. The energy level for the B2Δ3/2 state has also been calculated accurately to conclude that the B2Δ3/2 state lies above the A2Π1/2 state. This confirms X2Σ ↔ A2Π1/2 as a promising optical cycling transition for laser cooling RaOH.


I. Introduction

Recent years have seen a surge of interest in heavy-atom-containing radioactive molecules.1–4 Small molecules containing radioactive elements, e.g., Ra, Ac, or Pa offer sensitive and versatile avenues for precision spectroscopy search of fundamental physics beyond the Standard Model (BSM), including the search of the electron's electric dipole moment (eEDM), nuclear magnetic quadrupole moments (nMQM), and Schiff moments.5–21 Furthermore, Ra(I) in many molecular species has the potential to serve as efficient optical cycling centers5,16,22–28; such radium-containing molecules are promising candidates for the application of laser-cooling and subsequent advanced quantum control techniques. High-resolution laser spectroscopy studies of the radium monofluoride (RaF) molecule have recently been reported,1,28–30 and a laser-cooling scheme has been designed for RaF.28 Optical mass spectroscopy detection of radium-containing molecular cations including RaOH+ and RaOCH3+ has also been reported.31

Among the radium-containing molecular species, the neutral radium monohydroxide (RaOH) molecule is of significant interest. Computational studies have predicted an effective internal electric field of over 50 GV cm−1 in the electronic ground state of RaOH; RaOH is a therefore a sensitive molecule to use in a search for the eEDM.5–7,14,17,22 Since the 225Ra nucleus has an octupole deformation and low-lying nuclear excited states of opposite parity,32,33225RaOH might also have good sensitivity for a Schiff moment search. Compared with radium-containing diatomic species such as RaF, the degenerate vibrational first excited states of the bending modes in RaOH give rise to parity doublets that provide crucial robustness against systematics errors in precision measurements.34 Precision measurement using laser-cooled triatomic molecules has recently been demonstrated using laser-cooled CaOH molecules loaded in an optical dipole trap (ODT).35 Recent computational studies have also shown that the X2Σ ↔ A2Π1/2 transitions in RaOH can serve as efficient optical cycling transitions for laser slowing and cooling.36 It seems to require fewer repumping lasers to slow a cryogenic RaOH molecular beam sufficiently to be loaded into a magneto-optical trap (MOT) than in the corresponding experiment for CaOH. Note that CaOH is currently the only triatomic molecule that has been laser-slowed in three dimension, loaded into a MOT,37 sub-Doppler cooled, and loaded into an ODT38 and optical tweezer arrays.39 RaOH is a promising heavy-atom-containing triatomic molecule to realize the same level of advanced quantum control. The combination of laser spectroscopy, cooling, trapping, and advanced quantum state control and measurement of RaOH in principle offers a clear pathway to use RaOH in eEDM, nMQM, and Schiff moment measurement with enhanced sensitivity and versatility.

High-resolution laser spectroscopic detection and measurement of RaOH is a prerequisite for laser cooling of this molecule and the subsequent precision measurement. Laser spectroscopy studies of radioactive molecules are in general challenging, because of the need to deal with radioactivity, the relatively short window for observation, and the low concentration of the molecular sample. Ab initio quantum-chemical prediction of energies and properties plays an indispensable role in the spectroscopic measurement of radioactive molecules. For example, relativistic Fock-space coupled-cluster calculations of RaF have accurately predicted the level positions for electronic excited states of RaF that have greatly facilitated the recent experimental measurements of this molecule.1 We focus this paper on relativistic coupled-cluster calculations of the low-lying electronic states of RaOH pertinent to its spectroscopic detection and laser cooling. Exploiting the ability of the exact two-component theory40–42 for accounting for relativistic effects and that of the coupled-cluster theory43,44 for treating electron-correlation effects, the present computational study aims to obtain accurate energy levels for the A2Π1/2 and C2Σ states of RaOH to facilitate the spectroscopic detection of this molecule. Furthermore, the computational study also aims to answer an important question of relevance to laser cooling of RaOH: is the A2Π1/2 state the lowest electronic excited state of RaOH? We present the computational methods in Section II. The results are discussed in Section III with a conclusion given in Section IV.

II. Computational details

The calculations are focused on the electronic ground state, X2Σ, and three low-lying electronic excited states, A2Π1/2, B2Δ3/2, and C2Σ, in both RaF and RaOH. The open-shell spinors in the X2Σ, A2Π1/2, and B2Δ3/2 states are mainly composed of a radium 6s spinor, a radium 6p1/2 spinor, and a radium 6d3/2 spinor, respectively. The open-shell spinor in the C2Σ state is a mixture of 50% a radium 6p3/2 spinor and 50% a radium 6d3/2 spinor. We have employed the Kramers unrestricted coupled-cluster singles and doubles (CCSD)45 method and CCSD augmented with a non-iterative triples [CCSD(T)]46 method to treat electron-correlation effects using Kramers unrestricted Hartree–Fock (HF) wave functions optimized specifically for each state as the reference functions. The HF wave functions for the A2Π1/2 and C2Σ states have been obtained by specifying occupation numbers based on the dominating orbital angular momenta of the HF spinors. We have also carried out calculations with the electron attachment version of equation-of-motion coupled-cluster singles and doubles (EOMEA-CCSD)47–49 method for comparison using the closed-shell cation as the reference function. Since radium is a heavy atom, it is necessary to treat relativistic effects rigorously to obtain accurate results. We have used the exact two-component theory with atomic mean-field integrals based on the Dirac–Coulomb–Breit Hamiltonian (the X2CAMF scheme)50,51 to treat relativistic effects in our calculations. All these calculations have been carried out using the CFOUR program.52–55 In addition, to study high-level correlation (HLC) contributions, namely, electron-correlation contributions beyond CCSD(T), we have carried out CC singles doubles triples (CCSDT)56,57 and CCSDT augmented with a non-iterative quadruples [CCSDT(Q)]58 calculations. These CCSDT and CCSDT(Q) calculations have used the MRCC program59–61 with an interface with the CFOUR program, using the implementation of the spin-free exact two-component theory in its one-electron variant (the SFX2C-1e scheme) in the CFOUR program62,63 to treat scalar-relativistic effects. These calculations have been performed for the term energies of the A2Π and B2Δ states. We have not performed SFX2C-1e calculations for the C2Σ state, since the specification of the HF occupation numbers based on orbital angular momenta has not been implemented in the CFOUR program for the calculations without spin–orbit coupling. We have performed calculations for RaOH and RaF using the same computational protocol. The computational results for RaF are then compared with available measured and computed values1,29,30 to help assess the accuracy of the calculations for RaOH. The present calculations aim at an accuracy of 100–200 cm−1 for the term energies of the A2Π1/2 and C2Σ states to facilitate using these two states in spectroscopic detection of RaOH. The present study also aims to calculate the term energy of the B2Δ3/2 state with sufficient accuracy to determine its relative position with respect to the A2Π1/2 state. We base our calculations on the generic idea of theoretical model chemistries (also referred to as additivity schemes or composite schemes),64–70 which have been extensively studied in calculations of atomization energies for molecules aiming at chemical or subchemical accuracies. The chemical accuracy and subchemical accuracy correspond to 1 kcal mol−1 (around 300 cm−1) and 1 kJ mol−1 (around 80 cm−1), respectively. The term energies targeted in the present study are energy differences between different electronic states of a same molecule. These calculations in general have more favorable error cancellation than calculations of atomization energies, which involve breaking of chemical bonds. An accuracy of 100–200 cm−1 is feasible with the available computational resources.

Following the idea of an additivity scheme, we perform geometry optimizations at the X2CAMF-CCSD(T) level of theory that provide sufficiently accurate molecular structures, and then carry out X2CAMF-CCSD(T) calculations of electronic energies with systematically enlarged basis sets using these structures. Specifically, the equilibrium structures of the X2Σ, A2Π1/2, B2Δ3/2, and C2Σ states of RaOH and RaF have been computed at the X2CAMF-CCSD(T) level using the uncontracted ANO-RCC basis set for Ra71,72 and the uncontracted cc-pVTZ basis sets for O, H, and F.73 The Ra 5s, 5p, 5d, 6s electrons, the 2s and 2p electrons of O and F, and the 1s electrons of H have been correlated in these calculations. Based on benchmark studies of CCSD(T) accuracy for molecular structures,74 the present calculations aim at an accuracy of 0.01 Å for the bond lengths. X2CAMF-EOMEA-CCSD calculations have also been carried out for comparison. The harmonic vibrational frequencies have been calculated to obtain the zero-point vibrational energies within the harmonic approximation. The equilibrium structures and harmonic vibrational frequencies for RaF have been obtained by fitting local potential energy surfaces around the equilibrium bond lengths. The calculations of equilibrium structures for RaOH have been facilitated by the recent development of analytic X2CAMF-CCSD(T) and EOM-CCSD gradients.75 The harmonic vibrational frequencies of RaOH have been evaluated by means of numerical differentiation of analytically evaluated gradients.76

The electronic energies for the X2Σ, A2Π1/2, B2Δ3/2, and C2Σ states of RaF and RaOH have been calculated at the HF, CCSD, and CCSD(T) levels with systematically enlarged basis sets using the X2CAMF-CCSD(T) equilibrium structures. It is essential to saturate the treatment of basis-set effects to obtain accurate electronic energies. We have systematically extended the uncontracted ANO-RCC basis set of Ra to form uncontracted basis sets of triple-, quadruple-, and quintuple-zeta quality, denoted as the ANOTZ, ANOQZ, and ANO5Z sets, respectively. Specifically, the primitive set of s-, p-, d-, and f-type functions in the ANO-RCC set (28s25p17d13f) is augmented with two additional diffuse d-type functions with exponents 0.0130943 and 0.0043648, as well as one diffuse f-type functions with an exponent 0.0135, to form the ANOTZ (28s25p19d14f) set. The exponents of these additional functions are obtained by multiplying those of the most diffuse functions in the ANO-RCC set with a geometric factor of 1/3. The ANOQZ set is obtained by augmenting the ANOTZ set with four g-type functions with exponents 0.9705, 0.3235, 0.1075, and 0.043. The ANO5Z set is constructed by augmenting the ANOTZ set with six g-type functions with exponents 9.1368, 3.0456, 1.0152, 0.3384, 0.0919, and 0.0368, and four h-type functions with exponents 1.1268, 0.3756, 0.1114, 0.0446. These exponents been obtained to cover the exponent range of the g- and h-type functions in the Ra aug-cc-pCV5Z-PP set.77 The combined sets of ANOXZ (X = T, Q, 5) with the corresponding uncontracted aug-cc-pVXZ sets of O, H, and F are denoted as the XZ (X = T, Q, 5) sets.73 The X2CAMF-HF/5Z energies are used to estimate the basis-set-limit values of the HF energies. The X2CAMF-CCSD and CCSD(T) correlation energies obtained using the QZ and 5Z sets are extrapolated using a simple two-point formula78

 
image file: d3cp04040b-t1.tif(1)
to estimate the basis-set-limit values for correlation energies. These calculations have been carried out with the correlation of 27 valence and semicore electrons, namely, the Ra 5s, 5p, 5d, 6s electrons, the 2s and 2p electrons of O and F, and 1s electrons of H. Finally, the contributions from correlating inner-shell electrons are obtained as the differences between the X2CAMF-CCSD(T)/TZ energies with the correlation of 69 electrons (freezing only the Ra 1s, 2s, 2p, 3s, 3p, and 3d electrons) and those with the correlation of 27 electrons. For the calculation of HLC contributions, correlation-consistent triple-zeta basis sets contracted for the SFX2C-1e scheme77 have been used with the correlation of the Ra 5s, 5p, 5s electrons, the O and F 2s and 2p electrons, and the H 1s electrons. A Gaussian-type nuclear model79 has been used for the calculations presented here.

III. Results and discussions

A. The equilibrium structures and harmonic vibrational frequencies

The equilibrium structures and harmonic vibrational frequencies of the X2Σ, A2Π1/2, B2Δ3/2, and C2Σ states of RaOH computed at X2CAMF-CCSD(T) and EOMEA-CCSD levels are summarized in Table 1. Since the electronic excitations are localized at the radium atom, these four electronic states share similar structures. The Ra–O bond lengths of these states differ from each other by less than 0.01 Å, while the differences in the O–H bond lengths are smaller than 0.001 Å. The X2CAMF-EOMEA-CCSD structures are consistent with the X2CAMF-CCSD(T) ones, with discrepancies less than 0.01 Å for the bond lengths presented here. The X2CAMF-EOMEA-CCSD harmonic frequencies also agree well with the X2CAMF-CCSD(T) ones for the X2Σ and A2Π1/2 states. Because of the difficulty to converge the corresponding Kramers unrestricted Hartree–Fock wave functions in the displaced bent structures, we have not obtained X2CAMF-CCSD(T) harmonic frequencies for the B2Δ3/2 and C2Σ states. We have included X2CAMF-EOMEA-CCSD harmonic vibrational frequencies for these two states. Interestingly, while the four electronic states studied here have similar Ra–O and O–H stretching frequencies, the bending frequencies show relatively large variations. The bending frequency of 435 cm−1 for the C2Σ state is around 70 cm−1 greater than that of the X2Σ state.
Table 1 Computed structural parameters, harmonic vibrational frequencies, and equilibrium term energies for the electronic ground and low-lying excited states of RaOH and RaF. The uncontracted ANO-RCC basis set for Ra and the uncontracted cc-pVTZ basis sets for O, H, and F have been used for all calculations
X2Σ A2Π1/2 B2Δ3/2 C2Σ
a X2CAMF-CCSD(T) values with EOMEA-CCSD results in the parentheses. b Not calculated due to difficulty in HF convergence for displaced structures. c Computed values in ref. 29. d Measured values in ref. 1.
RaOHa R e (Ra–O) (Å) 2.265 (2.269) 2.260 (2.264) 2.268 (2.276) 2.268 (2.272)
R e (O–H) (Å) 0.956 (0.953) 0.956 (0.952) 0.956 (0.952) 0.956 (0.952)
ω e [Ra–O stretch] (cm−1) 490 (488) 495 (492) b (495) b (487)
ω e [Ra–O–H bend] (cm−1) 354 (363) 387 (395) b (345) b (435)
ω e [O–H stretch] (cm−1) 3912 (3969) 3919 (3973) b (3977) b (3970)
T e (cm−1) 0 (0) 12924 (12[thin space (1/6-em)]929) 15[thin space (1/6-em)]247 (15[thin space (1/6-em)]257) 15[thin space (1/6-em)]955 (16[thin space (1/6-em)]019)
RaFa R e (Ra–F) (Å) 2.251 (2.253) 2.254 (2.256) 2.266 (2.268) 2.271 (2.273)
ω e [Ra–F stretch] (cm−1) 438 (437) 434 (433) 424 (423) 425 (423)
T e (cm−1) 0 (0) 13[thin space (1/6-em)]499 (13[thin space (1/6-em)]502) 14[thin space (1/6-em)]935 (14[thin space (1/6-em)]936) 16[thin space (1/6-em)]884 (16[thin space (1/6-em)]944)
RaFc R e (Ra–F) (Å) 2.244 2.247 2.258 2.262
ω e [Ra–F stretch] (cm−1) 441 437 431 432
T e (cm−1) 0 (0) 13[thin space (1/6-em)]303 14[thin space (1/6-em)]352 16[thin space (1/6-em)]644
RaFd ω e [Ra–F stretch] (cm−1) 442 436
T e (cm−1) 0 (0) 13[thin space (1/6-em)]288 16[thin space (1/6-em)]621


We have also included X2CAMF-CCSD(T) and EOMEA-CCSD results for RaF in Table 1 and compared them with available computed and measured values.1,29 Similar to the case of RaOH, these four electronic states of RaF have similar bond lengths. The difference between the Ra–F bond lengths in the C2Σ state and the X2Σ state amounts to around 0.02 Å, slightly larger than that in RaOH. The X2CAMF-EOMEA-CCSD bond lengths and harmonic frequencies are consistent with the X2CAMF-CCSD(T) values. The present computational results for RaF agree reasonably well with available computed and measured results. The X2CAMF-CCSD(T) Ra–F bond lengths are around 0.007–0.009 Å longer than the four-component CC results in ref. 29. We expect further enlargement of basis sets in X2CAMF-CCSD(T) calculations to slightly reduce the Ra–F bond lengths. The X2CAMF-CCSD(T) harmonic frequencies for the B2Δ3/2 and C2Σ states are slightly smaller than the computed values in ref. 29, by around 7 cm−1.

B. Electronic energies

The equilibrium term energies obtained in the structural optimizations are also listed in Table 1. The results for the A2Π1/2 and C2Σ states of RaF agree reasonably well with the measured values and lie around 200–300 cm−1 above the measured values. The equilibrium term energy for the B2Δ3/2 state exhibits a relatively large discrepancy, lying around 600 cm−1 above the computed value in ref. 29. The corresponding equilibrium term energies for RaOH in Table 1 are expected to have similar accuracies. The discrepancies for RaF observed here can be attributed to basis-set effects. In particular, the B2Δ3/2 state has the unpaired electron occupying a radium 6d orbital. Since the differential angular momentum with respect to the ground state is large, the term energy of the B2Δ3/2 state is sensitive to basis-set effects. It thus is necessary to treat basis-set effects more thoroughly to improve the accuracy of the results.

Basis-set convergence of the HF, CCSD, and (T) contributions to the equilibrium term energies are summarized in Table 2. As expected, the HF energies converge rapidly with respect to the enlargement of basis sets. The HF/5Z term energies differ from the HF/QZ results by less than 3 cm−1. The electron-correlation contributions including the CCSD and (T) contributions exhibit larger basis-set effects. Since the excited states have the unpaired electrons in p- or d-type spinors with angular momenta higher than that in the ground state, the basis-set effects in general reduce the term energies. The TZ CCSD contributions to the term energies of the A2Π1/2 and C2Σ states amounts to 1448 cm−1 and 1689 cm−1, respectively. They are 148 cm−1 and 118 cm−1 greater than the corresponding ∞Z values of 1300 cm−1 and 1571 cm−1. The basis-set effects on the energies of the B2Δ3/2 state having the unpaired electron in a radium 6d spinor is substantially more significant. The TZ electron-correlation contribution to the term energy of the B2Δ3/2 state, namely, the sum of CCSD and (T) contributions, amount to 278 cm−1. This is to be compared with the basis-set-limit value of −611 cm−1; the remaining basis-set effects for the TZ result is as large as around 900 cm−1. The correlation of inner-shell electrons make minor contributions to these equilibrium term energies, amounting to 22 cm−1, 56 cm−1, and 37 cm−1 for the A2Π1/2, B2Δ3/2, and C2Σ states, respectively.

Table 2 Equilibrium term energies (cm−1) for the low-lying excited states of RaOH. The X2CAMF scheme has been used to account for relativistic effects
A2Π1/2 B2Δ3/2 C2Σ
HF-SCF TZ 11[thin space (1/6-em)]280 14[thin space (1/6-em)]925 13[thin space (1/6-em)]873
QZ 11[thin space (1/6-em)]273 14[thin space (1/6-em)]913 13[thin space (1/6-em)]874
5Z 11[thin space (1/6-em)]270 14[thin space (1/6-em)]916 13[thin space (1/6-em)]875
CCSD TZ 1448 248 1689
QZ 1368 −241 1626
5Z 1335 −397 1599
∞Z 1300 −559 1571
(T) TZ 138 62 153
QZ 151 12 172
5Z 158 −15 181
∞Z 166 −44 191
Core correlation TZ 22 56 37
Total 12[thin space (1/6-em)]758 14[thin space (1/6-em)]368 15[thin space (1/6-em)]673


A corresponding study of basis-set effects on the equilibrium term energies of the A2Π1/2, B2Δ3/2, and C2Σ states in RaF are summarized in Table 3. Electron-correlation contributions and convergence patterns similar to those of RaOH have been observed. This similarity might indicate that the quality of the present computational results for RaOH is similar to that for RaF.

Table 3 Equilibrium term energies (cm−1) for the low-lying excited states of RaF. The X2CAMF scheme has been used to account for relativistic effects
A2Π1/2 B2Δ3/2 C2Σ
HF-SCF TZ 11[thin space (1/6-em)]859 14[thin space (1/6-em)]855 14[thin space (1/6-em)]693
QZ 11[thin space (1/6-em)]859 14[thin space (1/6-em)]848 14[thin space (1/6-em)]712
5Z 11[thin space (1/6-em)]860 14[thin space (1/6-em)]852 14[thin space (1/6-em)]716
CCSD TZ 1447 206 1817
QZ 1351 −195 1736
5Z 1311 −390 1705
∞Z 1268 −594 1672
(T) TZ 140 72 173
QZ 152 −1 194
5Z 161 −9 205
∞Z 170 −17 216
Core correlation TZ 26 64 40
Total 13[thin space (1/6-em)]324 14[thin space (1/6-em)]305 16[thin space (1/6-em)]643


C. Assessment of the computational accuracy

The remaining errors for the computed equilibrium term energies include the errors due to the molecular structures, the errors in the estimate of the basis-set-limit values, the high-level correlation contributions beyond CCSD(T), and the quantum electrodynamics (QED) corrections. In this subsection we analyze these contributions and estimate the magnitude of the remaining errors for the computed equilibrium term energies.

The accuracy of the computed bond lengths is around 0.01 Å. We have carried out calculations by varying the bond lengths and found that a variation of 0.02 Å for the Ra–O or Ra–F bond length correspond to a variation of around 15 cm−1 for the computed term energies. It thus is safe to conclude that the remaining errors for the term energies due to the errors in the molecular structures are smaller than 20 cm−1. We should mention that an accuracy of 0.01 Å in bond length is not sufficient to guarantee an accuracy of 80 cm−1 for the calculation of bond energies, e.g., an O–H bond energy. It has been shown that an accuracy of 0.005 Å for the bond lengths is required to achieve subchemical accuracy.66 On the other hand, in the present calculations, all the electronic states have essentially the same O–H bond lengths; the errors in the O–H bond energies do not contribute significantly to the targeted term energies.

As shown in Tables 2 and 3, the basis-set convergence for both HF and electron-correlation energies has been observed to be smooth. Since the 5Z term energies of the A2Π1/2 and C2Σ states differ from the basis-set-limit values estimated through basis-set extrapolation by around 30 cm−1, we expect that the errors in the estimated basis-set-limit values are smaller than 20 cm−1. The (T) contributions to the term energies of the A2Π1/2 and C2Σ states are 166 cm−1 and 191 cm−1, respectively. Both of them are an order of magnitude smaller than the corresponding CCSD contributions. The high-level correlation (HLC) contributions, those beyond CCSD(T), thus are expected to further decrease significantly compared with the (T) contributions. It seems logical to assume that they are smaller than 60 cm−1. We have computed the HLC contributions to the term energy of the A2Π state as the difference between SFX2C-1e-CCSDT(Q), CCSDT, and CCSD(T) calculations using correlation-consistent triple-zeta basis sets. The full triples correction, namely, the difference between CCSDT and CCSD(T), amounts to around 2 cm−1. The quadruples contribution obtained as the difference between the CCSDT(Q) and CCSDT results is around 4 cm−1. While they used smaller basis sets and correlation space without including spin–orbit coupling, these calculations serve as a confirmation that 60 cm−1 is a conservative estimate for the HLC contributions. While we have included the contribution from the Breit term in the X2CAMF scheme, we have not included the QED corrections. The QED contributions to the term energies of the A2Π1/2 and C2Σ states in RaF have been estimated to be around −50 cm−1.29 An uncertainty of 50 cm−1 will be used here for RaOH as well.

If we assume that these four remaining uncertainties are statistically independent of each other, the remaining errors for the computed equilibrium term energies of the A2Π1/2 and C2Σ states is estimated to be around 80 cm−1. To be conservative about the error estimate, we use twice this estimate, namely 160 cm−1, as the estimate of the errors for these computed term energies. The final results for the equilibrium term energies of the A2Π1/2 and C2Σ states in RaOH thus are 12758 ± 160 cm−1 and 15673 ± 160 cm−1, respectively. We should mention that the computed term energies for the A2Π1/2 and C2Σ states of RaF are only around 40 cm−1 and 20 cm−1 larger than the measured values, as shown in Table 4. Since the electron-correlation contributions in RaOH take similar values and show similar convergence patterns as those in RaF, the present computational results for RaOH might have similar accuracies. Therefore, this estimated error of 160 cm−1 serves as a conservative estimate of the errors. Since the present study is focused on one molecule, the error estimate does not rigorously correspond to 95% confidence intervals defined in a statistical sense.80 On the other hand, the conservative estimate of the errors indicates that the actual error is unlikely larger than the estimate. This error of 160 cm−1 thus may be viewed as an estimate of 95% confidence intervals in practice.

Table 4 Computed equilibrium term energies (cm−1) for the low-lying excited states of RaOH together with the corresponding values for RaF computed using the same computational protocol and compared with results in the literature
A2Π1/2 B2Δ3/2 C2Σ
RaOH This work 12[thin space (1/6-em)]758 14[thin space (1/6-em)]368 15[thin space (1/6-em)]673
RaF This work 13[thin space (1/6-em)]324 14[thin space (1/6-em)]305 16[thin space (1/6-em)]643
Computed values in ref. 29 13[thin space (1/6-em)]303 14[thin space (1/6-em)]352 16[thin space (1/6-em)]644
Computed values in ref. 30 13[thin space (1/6-em)]299(36) 14[thin space (1/6-em)]300(61) 16[thin space (1/6-em)]615(69)
Measured values in ref. 1 13[thin space (1/6-em)]284.7(5) 16[thin space (1/6-em)]620.8(2)
Measured values in ref. 30 13[thin space (1/6-em)]284.427(1) 14[thin space (1/6-em)]332.82(13) 16[thin space (1/6-em)]613.6(12)


The contributions from the zero-point vibrational energies (ZPE) to the term energies of A2Π1/2 and C2Σ states in RaOH amount to 36 cm−1 and 68 cm−1, respectively, using the harmonic vibrational frequencies in Table 1. These agree well with the corresponding corrections of 40 cm−1 and 76 cm−1 obtained in our previous multi-state vibronic calculations.36 We adopt the latter values, since they include the anharmonic contributions. Adding these ZPE contributions to the equilibrium term energies, we recommend the values of 12798 ± 160 cm−1 and 15749 ± 160 cm−1 as the level positions for the vibrational origin transitions for the A2Π1/2–X2Σ and C2Σ–X2Σ electronic transitions, respectively, for spectroscopic detection of RaOH using laser induced fluorescence spectroscopy. In the case of RaOD, the ZPE corrections to the term energies of A2Π1/2 and C2Σ states amount to 31 cm−1 and 57 cm−1 in the multi-state vibronic calculations. The level positions for the vibrational origin transitions for the A2Π1/2–X2Σ and C2Σ–X2Σ electronic transitions in RaOD thus are predicted to take the values of 12789 ± 160 cm−1 and 15730 ± 160 cm−1, respectively.

The computed equilibrium term energy for the B2Δ3/2 state of RaOH is around 1600 cm−1 greater than the A2Π1/2 state. The HLC contribution for the term energy of the B2Δ3/2 state obtained in the present calculations is around −3 cm−1. This confirms that the estimated error of 60 cm−1 in the treatment of electron correlation is also conservative for the B2Δ3/2 state. Since the equilibrium term energy of the B2Δ3/2 state is more sensitive to the basis-set effects, the overall estimated uncertainty is slightly larger than those of the A2Π1/2 and C2Σ states but is still smaller than 200 cm−1. The present computational study thus has definitely placed the B2Δ3/2 state above the A2Π1/2 state and predicted the A2Π1/2 state to be the lowest electronic excited state of RaOH. The X2Σ ↔ A2Π1/2 electronic transition hence does not have any intermediate electronic state and can serve as efficient optical cycling transition for laser cooling of RaOH.

IV. Conclusion

We report a relativistic coupled-cluster study of low-lying electronic states of RaOH, a radioactive triatomic molecule of significant interest to laser cooling and search of new physics. The calculated term energies of 12798 ± 160 cm−1 and 15749 ± 160 cm−1 for the A2Π1/2 and C2Σ states and the computed vibrational frequencies presented here will expedite the observation and identification of RaOH, which can be produced with on-line methods at radioactive beam facilities1,81,82 as well as table-top methods using cryogenic buffer gas cooling.83 The calculations also predict the B2Δ3/2 state to lie substantially above the A2Π1/2 state; the A2Π1/2 state is the lowest electronic excited state in RaOH. Laser slowing of RaOH using optical cycling of the X2Σ ↔ A2Π1/2 transitions, which have favorable vibronic branching ratios, thus is highly recommended.

Conflicts of interest

The authors report there are no competing interests to declare.

Acknowledgements

The computational work at the Johns Hopkins University has been supported by the Department of Energy, Office and Science, Office of Basic Energy Sciences under Award Number DE-SC0020317. The computations at the Johns Hopkins University were carried out at Advanced Research Computing at Hopkins (ARCH) core facility (rockfish.jhu.edu), which is supported by the NSF under Grant OAC-1920103. Work at Caltech was supported by the Heising-Simons Foundation (award 2022-3361) and the NSF (award PHY-230936).

References

  1. R. F. Garcia Ruiz, R. Berger, J. Billowes, C. L. Binnersley, M. L. Bissell, A. A. Breier, A. J. Brinson, K. Chrysalidis, T. E. Cocolios, B. S. Cooper, K. T. Flanagan, T. F. Giesen, R. P. deGroote, S. Franchoo, F. P. Gustafsson, T. A. Isaev, Á. Koszorús, G. Neyens, H. A. Perrett, C. M. Ricketts, S. Rothe, L. Schweikhard, A. R. Vernon, K. D. A. Wendt, F. Wienholtz, S. G. Wilkins and X. F. Yang, Spectroscopy of short-lived radioactive molecules, Nature, 2020, 5810(7809), 396–400,  DOI:10.1038/s41586-020-2299-4.
  2. J. Kłos, H. Li, E. Tiesinga and S. Kotochigova, Prospects for assembling ultracold radioactive molecules from laser-cooled atoms, New J. Phys., 2022, 240(2), 25005,  DOI:10.1088/1367-2630/ac50ea.
  3. G. Arrowsmith-Kron, M. Athanasakis-Kaklamanakis, M. Au, J. Ballof, R. Berger, A. Borschevsky, A. A. Breier, F. Buchinger, D. Budker, L. Caldwell, C. Charles, N. Dattani, R. P. De Groote, D. DeMille, T. Dickel, J. Dobaczewski, C. E. Düllmann, E. Eliav, J. Engel, M. Fan, V. Flambaum, K. T. Flanagan, A. Gaiser, R. G. Ruiz, K. Gaul, T. F. Giesen, J. Ginges, A. Gottberg, G. Gwinner, R. Heinke, S. Hoekstra, J. D. Holt, N. R. Hutzler, A. Jayich, J. Karthein, K. G. Leach, K. Madison, S. Malbrunot-Ettenauer, T. Miyagi, I. D. Moore, S. Moroch, P. Navrátil, W. Nazarewicz, G. Neyens, E. Norrgard, N. Nusgart, L. F. Pašteka, A. N. Petrov, W. Plass, R. A. Ready, M. P. Reiter, M. Reponen, S. Rothe, M. Safronova, C. Scheidenberger, A. Shindler, J. T. Singh, L. V. Skripnikov, A. V. Titov, S.-M. Udrescu, S. G. Wilkins and X. Yang, Opportunities for fundamental physics research with radioactive molecules, 2023 Search PubMed.
  4. J. Ballof, N. Nusgart, P. Lalain, M. Au, R. Heinke, D. Leimbach, S. Stegemann, M. Schütt, S. Rothe and J. T. Singh, Progress towards the FRIB-EDM3-Frontend: A tool to provide radioactive molecules from isotope harvesting for fundamental symmetry studies, Nucl. Instrum. Methods Phys. Res., Sect. B, 2023, 541, 224–227,  DOI:10.1016/j.nimb.2023.04.049.
  5. T. A. Isaev, S. Hoekstra and R. Berger, Laser-cooled RaF as a promising candidate to measure molecular parity violation, Phys. Rev. A: At., Mol., Opt. Phys., 2010, 820(5), 52521,  DOI:10.1103/PhysRevA.82.052521.
  6. A. D. Kudashov, A. N. Petrov, L. V. Skripnikov, N. S. Mosyagin, T. A. Isaev, R. Berger and A. V. Titov, Ab initio study of radium monofluoride (RaF) as a candidate to search for parity- and time-and-parity-violation effects, Phys. Rev. A: At., Mol., Opt. Phys., 2014, 900(5), 52513,  DOI:10.1103/PhysRevA.90.052513.
  7. S. Sasmal, H. Pathak, M. K. Nayak, N. Vaval and S. Pal, Relativistic coupled-cluster study of RaF as a candidate for the parity- and time-reversal-violating interaction, Phys. Rev. A, 2016, 930(6), 62506,  DOI:10.1103/PhysRevA.93.062506.
  8. K. Gaul, S. Marquardt, T. Isaev and R. Berger, Systematic study of relativistic and chemical enhancements of P,T-odd effects in polar diatomic radicals, Phys. Rev. A: At., Mol., Opt. Phys., 2019, 990(3), 32509,  DOI:10.1103/PhysRevA.99.032509.
  9. N. M. Fazil, V. S. Prasannaa, K. V. P. Latha, M. Abe and B. P. Das, RaH as a potential candidate for electron electric-dipole-moment searches, Phys. Rev. A: At., Mol., Opt. Phys., 2019, 990(5), 52502,  DOI:10.1103/PhysRevA.99.052502.
  10. A. Sunaga, M. Abe, M. Hada and B. P. Das, Merits of heavy-heavy diatomic molecules for electron electric-dipole-moment searches, Phys. Rev. A: At., Mol., Opt. Phys., 2019, 990(6), 62506,  DOI:10.1103/PhysRevA.99.062506.
  11. K. Talukdar, M. K. Nayak, N. Vaval and S. Pal, Electronic structure parameter of nuclear magnetic quadrupole moment interaction in metal monofluorides, J. Chem. Phys., 2020, 1530(18), 184306,  DOI:10.1063/5.0028669.
  12. A. N. Petrov and L. V. Skripnikov, Energy levels of radium monofluoride RaF in external electric and magnetic fields to search for P- and T,P-violation effects, Phys. Rev. A: At., Mol., Opt. Phys., 2020, 1020(6), 62801,  DOI:10.1103/PhysRevA.102.062801.
  13. K. Gaul and R. Berger, Ab initio study of parity and time-reversal violation in laser-coolable triatomic molecules, Phys. Rev. A, 2020, 1010(1), 12508,  DOI:10.1103/PhysRevA.101.012508.
  14. K. Gaul and R. Berger, Toolbox approach for quasi-relativistic calculation of molecular properties for precision tests of fundamental physics, J. Chem. Phys., 2020, 1520(4), 44101,  DOI:10.1063/1.5121483.
  15. T. A. Isaev and R. Berger, Lasercooled radium monofluoride: A molecular all-in-one probe for new physics, 2020 Search PubMed.
  16. T. Fleig and D. DeMille, Theoretical aspects of radium-containing molecules amenable to assembly from laser-cooled atoms for new physics searches, New J. Phys., 2021, 230(11), 113039,  DOI:10.1088/1367-2630/ac3619.
  17. C. Zhang, X. Zheng and L. Cheng, Calculations of time-reversal-symmetry-violation sensitivity parameters based on analytic relativistic coupled-cluster gradient theory, Phys. Rev. A, 2021, 1040(1), 012814 CrossRef.
  18. A. Zakharova and A. Petrov, P,T-odd effects for the RaOH molecule in the excited vibrational state, Phys. Rev. A, 2021, 1030(3), 32819,  DOI:10.1103/PhysRevA.103.032819.
  19. P. Yu and N. R. Hutzler, Probing Fundamental Symmetries of Deformed Nuclei in Symmetric Top Molecules, Phys. Rev. Lett., 2021, 1260(2), 23003,  DOI:10.1103/PhysRevLett.126.023003.
  20. A. V. Oleynichenko, L. V. Skripnikov, A. V. Zaitsevskii and V. V. Flambaum, Laser-coolable AcOH+ ion for CP-violation searches, Phys. Rev. A, 2022, 1050(2), 22825,  DOI:10.1103/PhysRevA.105.022825.
  21. A. Zakharova, Rotating and vibrating symmetric-top molecule RaOCH3 in fundamental P,T-violation searches, Phys. Rev. A, 2022, 1050(3), 32811,  DOI:10.1103/PhysRevA.105.032811.
  22. T. A. Isaev, A. V. Zaitsevskii and E. Eliav, Laser-coolable polyatomic molecules with heavy nuclei, J. Phys. B: At., Mol. Opt. Phys., 2017, 50, 225101,  DOI:10.1088/1361-6455/aa8f34.
  23. Y. Osika and M. Shundalau, Multi-reference perturbation theory study on the RaCl molecule promising for the laser cooling, Comput. Theor. Chem., 2020, 1188, 112972,  DOI:10.1016/j.comptc.2020.112972.
  24. Y. Osika and M. Shundalau, Fock-space relativistic coupled cluster study on the spectroscopic properties of the low-lying states of the radium monobromide RaBr molecule, J. Quant. Spectrosc. Radiat. Transfer, 2021, 276, 107947,  DOI:10.1016/j.jqsrt.2021.107947.
  25. Y. Osika and M. Shundalau, Fock-space relativistic coupled cluster study on the raf molecule promising for the laser cooling, Spectrochim. Acta, Part A, 2022, 264, 120274 CrossRef CAS PubMed.
  26. Y. Osika, M. Shundalau and Y.-C. Han, Ab initio study on the spectroscopic and radiative properties of the low-lying states of the radium monoiodide RaI molecule, J. Quant. Spectrosc. Radiat. Transfer, 2022, 285, 108144,  DOI:10.1016/j.jqsrt.2022.108144.
  27. T. Isaev, D. Makinskii and A. Zaitsevskii, Radium-containing molecular cations amenable for laser cooling, Chem. Phys. Lett., 2022, 140078 CrossRef CAS.
  28. S.-M. Udrescu, S. Wilkins, A. Breier, R. F. Garcia Ruiz, M. Athanasakis Kaklamanakis, M. Au, I. Belošević, R. Berger, M. Bissell, K. Chrysalidis, et al., Precision spectroscopy and laser cooling scheme of a radium-containing molecule, 2023 Search PubMed.
  29. A. Zaitsevskii, L. V. Skripnikov, N. S. Mosyagin, T. Isaev, R. Berger, A. A. Breier and T. F. Giesen, Accurate ab initio calculations of RaF electronic structure appeal to more laser-spectroscopical measurements, J. Chem. Phys., 2022, 1560(4), 44306,  DOI:10.1063/5.0079618.
  30. M. Athanasakis-Kaklamanakis, S. G. Wilkins, L. V. Skripnikov, A. Koszorus, A. A. Breier, M. Au, I. Belosevic, R. Berger, M. L. Bissell, A. Borschevsky, A. Brinson, K. Chrysalidis, T. E. Cocolios, R. P. de Groote, A. Dorne, C. M. Fajardo-Zambrano, R. W. Field, K. T. Flanagan, S. Franchoo, R. F. Garcia Ruiz, K. Gaul, S. Geldhof, T. F. Giesen, D. Hanstorp, R. Heinke, T. A. Isaev, A. A. Kyuberis, S. Kujanpaa, L. Lalanne, G. Neyens, M. Nichols, L. F. Pasteka, H. A. Perrett, J. R. Reilly, S. Rothe, S. M. Udrescu, B. van den Borne, Q. Wang, J. Wessolek, X. F. Yang and C. Zuelch, Pinning down electron correlations in raf via spectroscopy of excited states, 2023 Search PubMed.
  31. M. Fan, C. A. Holliman, X. Shi, H. Zhang, M. W. Straus, X. Li, S. W. Buechele and A. M. Jayich, Optical Mass Spectrometry of Cold RaOH+ and RaOCH3+, Phys. Rev. Lett., 2021, 1260(2), 23002,  DOI:10.1103/PhysRevLett.126.023002.
  32. V. V. Flambaum, Enhancement of parity and time invariance violation in the radium atom, Phys. Rev. A: At., Mol., Opt. Phys., 1999, 600(4), R2611–R2613,  DOI:10.1103/PhysRevA.60.R2611.
  33. V. A. Dzuba, V. V. Flambaum, J. S. M. Ginges and M. G. Kozlov, Electric dipole moments of Hg, Xe, Rn, Ra, Pu, and TlF induced by the nuclear Schiff moment and limits on time-reversal violating interactions, Phys. Rev. A: At., Mol., Opt. Phys., 2002, 660(1), 12111,  DOI:10.1103/PhysRevA.66.012111.
  34. I. Kozyryev and N. R. Hutzler, Precision Measurement of Time-Reversal Symmetry Violation with Laser-Cooled Polyatomic Molecules, Phys. Rev. Lett., 2017, 1190(13), 133002,  DOI:10.1103/PhysRevLett.119.133002.
  35. L. Anderegg, N. B. Vilas, C. Hallas, P. Robichaud, A. Jadbabaie, J. M. Doyle and N. R. Hutzler, Quantum Control of Trapped Polyatomic Molecules for eEDM Searches, 2023 Search PubMed.
  36. C. Zhang, N. R. Hutzler and L. Cheng, Intensity-Borrowing Mechanisms Pertinent to Laser Cooling of Linear Polyatomic Molecules, J. Chem. Theory Comput., 2023, 19, 4136–4148,  DOI:10.1021/acs.jctc.3c00408.
  37. N. B. Vilas, C. Hallas, L. Anderegg, P. Robichaud, A. Winnicki, D. Mitra and J. M. Doyle, Magneto-optical trapping and sub-Doppler cooling of a polyatomic molecule, Nature, 2022, 6060(7912), 70–74 CrossRef PubMed.
  38. C. Hallas, N. B. Vilas, L. Anderegg, P. Robichaud, A. Winnicki, C. Zhang, L. Cheng and J. M. Doyle, Optical Trapping of a Polyatomic Molecule in an [small script l]-Type Parity Doublet State, Phys. Rev. Lett., 2023, 1300(15), 153202,  DOI:10.1103/PhysRevLett.130.153202.
  39. N. B. Vilas, P. Robichaud, C. Hallas, G. K. Li, L. Anderegg and J. M. Doyle, An optical tweezer array of ultracold polyatomic molecules, 2023 Search PubMed.
  40. K. G. Dyall, Interfacing relativistic and nonrelativistic methods. I. Normalized elimination of the small component in the modified Dirac equation, J. Chem. Phys., 1997, 1060(23), 9618–9626 CrossRef.
  41. M. Iliaš and T. Saue, An infinite-order two-component relativistic Hamiltonian by a simple one-step transformation, J. Chem. Phys., 2007, 126, 064102 CrossRef PubMed.
  42. W. Liu and D. Peng, Exact two-component hamiltonians revisited, J. Chem. Phys., 2009, 1310(3), 031104 CrossRef PubMed.
  43. T. D. Crawford and H. F. Schaefer III, An introduction to coupled cluster theory for computational chemists, Rev. Comput. Chem., 2000, 14, 33–136 CAS.
  44. R. J. Bartlett and M. Musial, Coupled-cluster theory in quantum chemistry, Rev. Mod. Phys., 2007, 79, 291–352 CrossRef CAS.
  45. G. D. Purvis III and R. J. Bartlett, A full coupled-cluster singles and doubles model: The inclusion of disconnected triples, J. Chem. Phys., 1982, 76, 1910–1918 CrossRef.
  46. K. Raghavachari, G. W. Trucks, J. A. Pople and M. Head-Gordon, A fifth-order perturbation comparison of electron correlation thoeries, Chem. Phys. Lett., 1989, 1570(6), 479–483,  DOI:10.1016/S0009-2614(89)87395-6.
  47. J. F. Stanton and R. J. Bartlett, The equation of motion coupled‚Äêcluster method. A systematic biorthogonal approach to molecular excitation energies, transition probabilities, and excited state properties, J. Chem. Phys., 1993, 98, 7029–7039 CrossRef CAS.
  48. M. Nooijen and R. J. Bartlett, Equation of motion coupled cluster method for electron attachment, J. Chem. Phys., 1995, 102, 3629–3647 CrossRef CAS.
  49. J. F. Stanton and J. Gauss, A simple scheme for the direct calculation of ionization potentials with coupled-cluster theory that exploits established excitation energy methods, J. Chem. Phys., 1999, 111, 8785–8788 CrossRef CAS.
  50. J. Liu and L. Cheng, An atomic mean-field spin-orbit approach within exact two-component theory for a non-perturbative treatment of spin-orbit coupling, J. Chem. Phys., 2018, 148, 144108 CrossRef PubMed.
  51. C. Zhang and L. Cheng, Atomic Mean-Field Approach within Exact Two-Component Theory Based on the Dirac–Coulomb–Breit Hamiltonian, J. Phys. Chem. A, 2022, 126, 4537–4553,  DOI:10.1021/acs.jpca.2c02181.
  52. 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, A. Asthana, R. J. Bartlett, U. Benedikt, C. Berger, D. E. Bernholdt, S. Blaschke, Y. J. Bomble, S. Burger, O. Christiansen, D. Datta, F. Engel, R. Faber, J. Greiner, M. Heckert, O. Heun, M. Hilgenberg, C. Huber, T.-C. Jagau, D. Jonsson, J. Jusélius, T. Kirsch, K. Klein, G. M. KopperW. J. Lauderdale, F. Lipparini, J. Liu, T. Metzroth, L. A. Mück, D. P. O'Neill, T. Nottoli, D. R. Price, E. Prochnow, C. Puzzarini, K. Ruud, F. Schiffmann, W. Schwalbach, C. Simmons, S. Stopkowicz, A. Tajti, J. Vázquez, F. Wang, J. D. Watts, C. Zhang, X. Zheng, and 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 https://www.cfour.de.
  53. D. A. Matthews, L. Cheng, M. E. Harding, F. Lipparini, S. Stopkowicz, T.-C. Jagau, P. G. Szalay, J. Gauss and J. F. Stanton, Coupled-cluster techniques for computational chemistry: The CFOUR program package, J. Chem. Phys., 2020, 1520(21), 214108,  DOI:10.1063/5.0004837.
  54. J. Liu, Y. Shen, A. Asthana and L. Cheng, Two-component relativistic coupled-cluster methods using mean-field spin-orbit integrals, J. Chem. Phys., 2018, 148, 034106 CrossRef PubMed.
  55. A. Asthana, J. Liu and L. Cheng, Exact two-component equation-of-motion coupled-cluster singles and doubles method using atomic mean-field spin-orbit integrals, J. Chem. Phys., 2019, 150, 074102 CrossRef PubMed.
  56. J. Noga and R. J. Bartlett, The full CCSDT model for molecular electronic structure, J. Chem. Phys., 1987, 86, 7041–7050 CrossRef CAS.
  57. G. E. Scuseria and H. F. Schaefer III, A new implementation of the full CCSDT model for molecular electronic structure, Chem. Phys. Lett., 1988, 152, 382–386 CrossRef CAS.
  58. Y. J. Bomble, J. F. Stanton, M. Kállay and J. Gauss, Coupled-cluster methods including noniterative corrections for quadruple excitations, J. Chem. Phys., 2005, 123, 054101 CrossRef PubMed.
  59. M. Kállay and P. R. Surján, Higher excitations in coupled-cluster theory, J. Chem. Phys., 2001, 115, 2945–2954 CrossRef.
  60. M. Kállay, P. R. Nagy, D. Mester, Z. Rolik, G. Samu, J. Csontos, J. Csóka, P. Bernát Szabó, L. Gyevi-Nagy, B. Hégely, I. Ladjánszki, L. Szegedy, B. Ladóczki, K. Petrov, M. Farkas, P. D. Mezei and Á. Ganyecz, The MRCC program system: Accurate quantum chemistry from water to proteins, J. Chem. Phys., 2020, 1520(7), 74107,  DOI:10.1063/1.5142048.
  61. M. Kállay, P. R. Nagy, D. Mester, L. Gyevi-Nagy, J. Csóka, P. B. Szabó, Z. Rolik, G. Samu, J. Csontos, B. Hégely, Á. Ganyecz, I. Ladjánszki, L. Szegedy, B. Ladóczki, K. Petrov, M. Farkas, P. D. Mezei and R. A. Horváth, MRCC, a quantum chemical program suite, see https://www.mrcc.hu.
  62. K. G. Dyall, Interfacing relativistic and nonrelativistic methods. IV. One- and two-electron scalar approximations, J. Chem. Phys., 2001, 115, 9136–9143 CrossRef CAS.
  63. L. Cheng and J. Gauss, Analytic energy gradients for the spin-free exact two-component theory using an exact block diagonalization for the one-electron Dirac Hamiltonian, J. Chem. Phys., 2011, 135, 084114 CrossRef PubMed.
  64. L. A. Curtiss, K. Raghavachari, P. C. Redfern, V. Rassolov and J. A. Pople, Gaussian-3 (G3) theory for molecules containing first and second-row atoms, J. Chem. Phys., 1998, 1090(18), 7764–7776,  DOI:10.1063/1.477422.
  65. M. S. Schuurman, S. R. Muir, W. D. Allen and H. F. Schaefer III, Toward subchemical accuracy in computational thermochemistry: Focal point analysis of the heat of formation of NCO and [H,N,C,O] isomers, J. Chem. Phys., 2004, 120, 11586–11599 CrossRef CAS PubMed.
  66. A. Tajti, P. G. Szalay, A. G. Császár, M. Kállay, J. Gauss, E. F. Valeev, B. A. Flowers, J. Vázquez and J. F. Stanton, Heat: High accuracy extrapolated ab initio thermochemistry, J. Chem. Phys., 2004, 1210(23), 11599–11613 CrossRef PubMed.
  67. A. Karton, E. Rabinovich, J. M. L. Martin and B. Ruscic, W4 theory for computational thermochemistry: In pursuit of confident sub-kJ/mol predictions, J. Chem. Phys., 2006, 1250(14), 144108,  DOI:10.1063/1.2348881.
  68. N. J. DeYonker, T. R. Cundari and A. K. Wilson, The correlation consistent composite approach (ccCA): An alternative to the Gaussian-n methods, J. Chem. Phys., 2006, 1240(11), 114104,  DOI:10.1063/1.2173988.
  69. L. A. Curtiss, P. C. Redfern and K. Raghavachari, Gaussian-4 theory, J. Chem. Phys., 2007, 1260(8), 084108,  DOI:10.1063/1.2436888.
  70. K. A. Peterson, D. Feller and D. A. Dixon, Chemical accuracy in ab initio thermochemistry and spectroscopy: current strategies and future challenges, Theor. Chem. Acc., 2012, 1310(1), 1079,  DOI:10.1007/s00214-011-1079-5.
  71. K. Fægri, Relativistic Gaussian basis sets for the elements K – Uuo, Theor. Chem. Acc., 2001, 105, 252–258 Search PubMed.
  72. B. O. Roos, V. Veryazov and P.-O. Widmark, Relativistic atomic natural orbital type basis sets for the alkaline and alkaline-earth atoms applied to the ground-state potentials for the corresponding dimers, Theor. Chem. Acc., 2004, 1110(2), 345–351,  DOI:10.1007/s00214-003-0537-0.
  73. T. H. Dunning, Jr, Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen, J. Chem. Phys., 1989, 90, 1007–1023 CrossRef.
  74. S. Coriani, D. Marchesan, J. Gauss, C. Hättig, T. Helgaker and P. Jørgensen, The accuracy of ab initio molecular geometries for systems containing second-row atoms, J. Chem. Phys., 2005, 1230(18), 184107,  DOI:10.1063/1.2104387.
  75. X. Zheng, C. Zhang, J. Liu and L. Cheng, Geometry optimizations with spinor-based relativistic coupled-cluster theory, J. Chem. Phys., 2022, 1560(15), 151101,  DOI:10.1063/5.0086281.
  76. J. F. Stanton, C. L. Lopreore and J. Gauss, The equilibrium structure and fundamental vibrational frequencies of dioxirane, J. Chem. Phys., 1998, 108, 7190–7196 CrossRef CAS.
  77. J. G. Hill and K. A. Peterson, Gaussian basis sets for use in correlated molecular calculations. xi. pseudopotential-based and all-electron relativistic basis sets for alkali metal (k–fr) and alkaline earth (ca–ra) elements, J. Chem. Phys., 2017, 1470(24), 244106 CrossRef PubMed.
  78. T. Helgaker, W. Klopper, H. Koch and J. Noga, Basis-set convergence of correlated calculations on water, J. Chem. Phys., 1997, 106, 9639–9646 CrossRef CAS.
  79. L. Visscher and K. G. Dyall, Dirac-fock atomic electronic structure calculations using different nuclear charge distributions, At. Data Nucl. Data Tables, 1997, 670(2), 207–224,  DOI:10.1006/adnd.1997.0751.
  80. B. Ruscic, Uncertainty quantification in thermochemistry, benchmarking electronic structure computations, and Active Thermochemical Tables, Int. J. Quantum Chem., 2014, 1140(17), 1097–1101,  DOI:10.1002/qua.24605.
  81. Y. Blumenfeld, T. Nilsson and P. Van Duppen, Facilities and methods for radioactive ion beam production, Phys. Scr., 2013, 2013(T152), 14023,  DOI:10.1088/0031-8949/2013/T152/014023.
  82. T. E. Cocolios, H. H. Al Suradi, J. Billowes, I. Budincević, R. P. de Groote, S. De Schepper, V. N. Fedosseev, K. T. Flanagan, S. Franchoo, R. F. Garcia Ruiz, H. Heylen, F. Le Blanc, K. M. Lynch, B. A. Marsh, P. J. R. Mason, G. Neyens, J. Papuga, T. J. Procter, M. M. Rajabali, R. E. Rossel, S. Rothe, G. S. Simpson, A. J. Smith, I. Strashnov, H. H. Stroke, D. Verney, P. M. Walker, K. D. A. Wendt and R. T. Wood, The Collinear Resonance Ionization Spectroscopy (CRIS) experimental setup at CERN-ISOLDE, Nucl. Instrum. Methods Phys. Res., Sect. B, 2013, 317, 565–569,  DOI:10.1016/j.nimb.2013.05.088.
  83. N. R. Hutzler, H.-I. Lu and J. M. Doyle, The Buffer Gas Beam: An Intense, Cold, and Slow Source for Atoms and Molecules, Chem. Rev., 2012, 1120(9), 4803–4827,  DOI:10.1021/cr200362u.

This journal is © the Owner Societies 2023
Click here to see how this site uses Cookies. View our privacy policy here.