Rotational Splittings in Diatomic Molecules of Interest to Searches for New Physics
Abstract
Diatomic molecules with an energetically low-lying state are attractive platforms to detect new physics beyond the Standard Model, such as parity- and time-reversal violating phenomena. One of the advantages of using a state is its tiny -splitting due to the coupling between the electronic and rotational angular momenta, which facilitates polarizing the molecules in small external electric fields. Theoretical estimation of the magnitude of the -splitting is helpful for planning new experiments. In this study, we present a theoretical model to calculate the -splitting. Our model integrates the relativistic four-component wavefunction and the traditional rotational Hamiltonian based on Hund’s case (a). The multireference character of the wavefunction is taken into account. Our calculations for PtH and molecules qualitatively agree with experiment. The -splitting of for the rotational ground state is predicted to be around 9 kHz. This tiny splitting can reduce the systematic uncertainty, but in a practical experiment, it may cause depolarization during rotation ramp-up.
I Introduction
Diatomic molecules are used as powerful low-energy probes in the search for physics beyond the Standard Model (SM) of elementary particles [1]. In particular, measurements and calculations on the hafnium flouride cation [2, 3, 4] currently yield the strongest constraint on the electric dipole moment (EDM) of the electron [5, 6]. Further advances are expected in the near future from work on the thorium monofluoride cation ThF+ [7, 8, 9, 10, 11, 12] and the tantalum monoxide cation TaO+ [13, 14, 15, 16] the latter of which can also be employed as a probe for nuclear charge-parity (CP) violation through the nuclear magnetic quadrupole moment [17, 18].
In state-of-the-art experiments [19, 2] aiming to measure a molecular EDM in the laboratory frame it is required to mix opposite parity states by polarizing the molecule through an external electric field. This mixing depends on the separation of the molecular target rovibronic energy levels that is induced by the coupling of intrinsic angular momenta to the angular momentum of the molecule rotating in the laboratory frame. As an example, the electronic state in which the EDM measurement is carried out in the thorium monoxide (ThO) and the ThF+ molecules is a state where the total electronic angular momentum projection onto the internuclear axis is . This state exhibits [20] quasi-degenerate pairs of rotational levels with well-defined parity and with energy splittings (or where is the total electronic orbital angular momentum projection). It is the purpose of this paper to present a method for calculating these so-called - or -doublings and its application to molecules of interest in low-energy searches of CP-violation beyond that already known to exist in Nature [21, 5, 6].
In the following section II we briefly discuss the theory underlying our approach and the specific approximations we make in view of the relevant experimental conditions. Next, we explain the mechanism in our model that leads to the -doublet in the state. Typically, an EDM measurement is carried out in the rovibrational ground levels of either the electronic ground state or an energetically low-lying excited electronic state of the molecule [20, 2]. Under these circumstances the required molecular vibrational overlap integrals can be approximated conveniently. In section III we discuss applications of our approach. The initial application concerns the platinum monohydride (PtH) molecule. Rotational couplings have been calculated earlier and quite extensively for this molecule [22, 23] which allows us to draw comparisons and to verify that our present method is correctly implemented. We then go on to apply our approach to molecular ions that are being prepared to become leading contenders in EDM measurements, the ThF+ and the TaO+ molecular ions. We conclude on our findings in section IV.
II Theory
II.1 -doublet structure
Earlier approaches to the calculation of molecular rotational couplings were based on a framework of scalar relativistic (or non-relativistic) wavefunctions and required the explicit calculation of matrix elements over the spin-orbit interaction Hamiltonian: the matrix elements were treated through perturbation theory [24, 25, 26, 27, 28, 29] or matrix diagonalization [30, 23].
The present theoretical approach closely follows the approach as described by Lefèbvre-Brion and Field [31] which uses an effective theory for many-body states in Born-Oppenheimer approximation represented in Hund’s case (a) for diatomic molecules. The choice of a Hund’s case (a) model is justified by earlier findings for the TaO+ cation [14] showing that molecular electronic states are represented to a very good approximation within this model. However, we use molecular electronic wavefunctions from a four-component Dirac-theory-based framework which includes the spin-orbit interaction more accurately and already in the zeroth-order wavefunctions. Calculations of -splittings including nuclear angular momenta have been reported [32, 33, 27, 28]. However, the hyperfine interaction presents only a minute perturbation that can be neglected given the other approximations made in the effective approach.
-type doubling matrix elements have been determined for molecular states by Brown et al. in 1987 [34]. Although our present theoretical formulation is quite different from that approach, the qualitative aspects of the coupling are the same.
The molecules’ energy is represented by the molecular Hamiltonian
| (1) |
where is an electronic Hamiltonian. For a rigid diatomic rotor the Hamiltonian representing the molecular rotational motion is
| (2) |
where is the reduced mass for the two fixed atomic nuclei with rest masses and , respectively, is the (constant) distance coordinate between the two nuclei and is the operator of rotational angular momentum. Its classical counterpart is angular momentum taken with respect to an origin lying in the center of mass of the diatomic molecule, and it is expressed in space-fixed (laboratory) coordinates.
The molecular rotational angular momentum operator can be represented in terms of electronic angular momentum operators as
| (3) |
where is the vector operator of total angular momentum, of total electronic orbital angular momentum and of total electronic spin. Inserting Eq. (3) into Eq. (2), exploiting the fact that components of different angular momentum operators commute and straightforward manipulations yield the rotational Hamiltonian in Hund’s case (a):
| (4) |
For the matrix representation of this operator we use explicit signed basis states defined as (see [31], p.221 ff.)
| (5) |
where , and are the projection quantum numbers of , and , respectively, onto the molecular axis and is the quantum number of the total angular momentum. Here we do not explicitly show another quantum number, parity, which can be obtained with for each state [35].
In practice, the evaluation of corresponding matrix elements requires the expansion of the operators and with molecule-fixed commutation rules in terms of operators with anomalous (space-fixed) commutation rules. It is found ([31], p.76) that
| (6) |
where and act in space-fixed coordinates and are the direction-cosine matrix elements with a space-fixed and a molecule-fixed coordinate and specifies a unit vector. In such a representation can be evaluated in a basis of states labeled as where and are the space-fixed total angular momentum and total angular momentum projection quantum numbers, respectively. The matrix element is expressed by ([31], p.78)
| (7) |
We use this expression in the explicit evaluation of our matrix elements in Hund’s case (a) formalism.
Even if the quantum numbers and are known to sufficient accuracy for a given state, the evaluation of the matrix elements of the Hamiltonian in Eq. (4) using the basis states in Eq. (5) requires the knowledge of the total orbital angular-momentum quantum number when the expression , for example, needs to be calculated. However, is not an exact quantum number in a (relativistic) molecular field, and thus the value of in the respective “ complex” [36] used for the evaluation is always approximate. In the PtH molecule is rather well defined, but in ThF+ and TaO+ the situation is more ambiguous. We discuss these cases and our reasoned choices in the applications section below.
In our current approach vibrational degrees of freedom are treated as follows. In the framework of the Born-Oppenheimer approximation [37] the molecular wavefunction is separated [38] into an electronic part and a vibrational part ,
| (8) |
for an -electron diatomic molecule where the electronic wavefunction depends parametrically on the nuclear positions. The basis functions given in Eq. (5) purely describe electronic degrees of freedom and we denote these as in the corresponding bra and ket vectors. The vibrational wavefunction is given in terms of nuclear degrees of freedom and will be denoted by . As an example for a given matrix element, we take one term of the rotational Hamiltonian representing the spin-uncoupling and write its matrix element as
| (9) | |||||
with
| (10) |
where the first factor on the rhs. of Eq. (9) is an integral over nuclear coordinates and the second factor is an electronic integral. Although we do not explicitly write it in the equation, the electronic energy associated with is added to the diagonal part of the matrix elements.
Supposing that the potential-energy curves of the respective electronic states are sufficiently parallel near the equilibrium internuclear distance of the diatomic molecule, the overlap of the corresponding ground-state vibrational wavefunctions can be approximated as from which
| (11) |
where is the rotational constant of the target electronic state. Approximating the latter by the equilibrium rotational constant (or by for the vibrational ground state, if available) the sample matrix element in Eq. (9) becomes
| (12) |
The current approach is justified for the purposes mentioned in the introduction.
We formulate the -splitting based on Hund’s case (a) above, but our electronic Hamiltonian includes the spin-orbit interaction, and thus and are not exactly good quantum numbers for our wavefunction. In relativistic wavefunctions, the signed and basis states defined in Eq. (II.1) can be generalized as follows
| (13) |
The and bases are orthogonalized, and the linear combination coefficients are determined to satisfy the normalization condition. This linear expansion of the basis is in our model key to the description of the -splitting in states, as shown in later sections.
In addition, for and molecules, we take the electronic configuration of each basis into account because of the multireference character of these wavefunctions. This is not required for the PtH molecule since here the interaction space is comprised by one electronic configuration only. The coupling occurs only when the electronic configurations (i.e., spinor structures) of bra and ket Hilbert-space vectors are the same, and the weight of the target electronic configuration has to be included as a factor. The target electronic configurations are and in the case of , and in the case of . When we denote the weight by the coefficient , the matrix element of the spin-uncoupling term in the basis and basis can be expressed by
Examples of the coefficients and are described in Secs. III.2 and III.3.
II.2 Mechanism of -doubling in states
In the following, we refer to vectors with exact quantum numbers in the picture as “basis vectors” or “basis functions” and those with approximate quantum numbers simply as “states”, in order to avoid confusion.
The - splitting does not occur in the state in the lowest approximation that considers only the terms in Eq. (II.1) because neither () nor () operators can couple and basis functions [34].
Furthermore, there is also no direct rotational coupling between and other electronic states. Any -splitting must, therefore, be due to rotational couplings between excited electronic states that can mix with the target state through a different mechanism.
We first use a simple model, assuming that the - splitting of the state occurs due to the contribution from a nearby term which rotationally couples with another energetically close state. The contribution from the term is due to spin-orbit coupling and can be written as
| (15) |
and correspond to and defined in Eq. (13). The lhs. of Eq. (15) is to be understood as a physical state and the terms on the rhs. of Eq. (15) are basis functions. Furthermore, and are now approximate quantum numbers that are close to integer/half-integer values. From the deviation between and ( and ), we can assume the contribution from another basis function with a different value of (). When , a basis function contributes to the state, as shown in Eq. (15).
The coefficients and can be obtained by solving the simultaneous equations
| (16) | |||||
The normalization condition () is automatically satisfied in our model, as follows:
| (17) | ||||
where corresponds to the value associated with the state of the corresponding coefficient (). For example, in the case of the state of , and were obtained from and .
Although the magnitudes of and are determined without arbitrariness, the following approximations are included in this model: (i) We can consider the contribution from one basis vector (), and cannot determine the contributions from other non- basis vectors. (ii) We cannot determine the (relative) sign of the linear combination coefficients ( and ). The sign would not always be positive, as shown in the case of PtH (Table 5). (iii) Since the model takes only and into account, the spin multiplicity of state is arbitrary. We selected the spin states that are energetically closest to the lowest-energy state: for and for .
The - splitting of the state is due to a small contribution from the basis function that causes the -doubling due to the coupling between the basis functions. As an example, we show the analytical expression for the coupling between the and basis functions:
The ordering of the quantum numbers is defined in Eq. (II.1). To obtain the matrix element shown in Eq. (II.1), the coefficients ( and ) need to be determined. In the case of the coupling between the state and the basis functions of the ground state of (cf. Table 7),
| ; | (19) | ||||
| ; |
The other coupling terms are zero because of the cancellation
| (20) | |||||
Similarly we can show that the S-uncoupling terms do not contribute to the splitting: the matrix elements between the and basis functions satisfy and .
The coefficient for the electronic configuration defined in Eq. (II.1) is obtained from the ratio of the target electronic configurations. For example, the of the ground of is obtained from , where 0.75 and 0.12 are the squares of the respective expansion coefficients of the and Slater determinants (cf. Table 6). The value of of the correction basis function of the state, , is obtained from the closest state with the energy of 6639 , where . of is obtained in the same manner.
III Application
We applied the developed code to three molecules. The application to PtH, for which experimental values have been reported, largely serves for verification purposes of the present method. is an important molecule in its own right and an example for showing the ambiguity of the quantum number . Also here we can compare with experimental results. For we are then able to make confident predictions for the expected -splitting. The employed input parameters (electronic energy and rotational constants) and configuration coefficients ( and ) are summarised in the Appendix.
III.1 Platinum hydride
The -splittings of the five lowest-energy electronic states of PtH are listed Table 1. These five states arise from the two atomic states, Pt’s and H’s [39]. Our theory and code successfully reproduce the order of magnitude for the available experimental values, but some input-parameter dependence is observed.
The input-parameter dependence of the -splitting becomes significant when the relative difference of the energy gaps between the coupled electronic states is changed. For the (I) state, TW-B shows the lowest splitting, while TW-D shows approximately twice the splitting. This is due to the employed energy difference between the (I) and (I) state: The energy difference of the former (latter) is 1 935 (1 213 ), whose relative difference is approximately 1.6. From the comparison between TW-A and TW-B, the contributions from the (II), 1/2 (II) states to the (I) state reach more than 10% at small , and goes beyond 20% at large , even though its energy gap is ca. 8 000 . When the experimental energy is used (TW-D and TW-E), -splitting is overestimated. The calculated values would become closer to the experiment if the vibrational wavefunction provided in Eq. (11) is taken into account and the coupling between the states provided in Eq. (9) is diminished.
The coefficient matrix of PtH at the rotational state is visualized in FIG. 1. basis dominantly contributes to the electronic ground states, but a strong mixing is observed between the (II) and (II) states. Note that the eigenvalues and diagonal parts of the matrix elements are far from the electronic energies shown in Table 5 because of the contribution from the first line of Eq. (II.1) which are the rotational energies for the states in question.
Overall, our approach yields results of similar or even better quality than those obtained in Ref. 23 where vibrational overlap has been taken into account.
| PW[23] | TW-A | TW-B | TW-C | TW-D | TW-E | exp. [40] | |
| (I) | (I) | (I) | (I) | (II) | |||
| Energy | 23111Only the three lowest states (, (I), and (I)) are used in the matrix representation. | 23 | 41 | exp. [40]222The theoretical excitation energies of Ref. [41] are employed for the states. | exp. [40]2 | ||
| 0 () | |||||||
| 2.5 | 0.000 | 1 | 1 | 6 | 8 | 6 | |
| 3.5 | 0.000 | 7 | 8 | 4 | 5 | 4 | |
| 4.5 | 0.000 | 2 | 3 | 1 | 2 | 1 | |
| 5.5 | 0.000 | 6 | 7 | 4 | 5 | 4 | |
| 10.5 | 0.008 | 0.015 | 0.017 | 0.008 | 0.011 | 0.008 | |
| 15.5 | 0.053 | 0.099 | 0.114 | 0.055 | 0.071 | 0.055 | |
| 20.5 | 0.213 | 0.396 | 0.459 | 0.218 | 0.280 | 0.218 | |
| 2014.4 ( (I)) | |||||||
| 0.5 | 27.506 | 35.324 | 35.324 | 33.573 | 35.324 | 33.573 | |
| 1.5 | 55.000 | 70.630 | 70.631 | 67.107 | 70.604 | 67.107 | |
| 2.5 | 82.469 | 105.899 | 105.905 | 100.566 | 105.794 | 100.566 | |
| 3.5 | 109.898 | 141.114 | 141.128 | 133.913 | 140.853 | 133.911 | |
| 4.5 | 137.277 | 176.257 | 176.285 | 167.112 | 175.740 | 167.108 | |
| 5.5 | 164.592 | 211.310 | 211.360 | 200.129 | 210.414 | 200.123 | |
| 10.5 | 299.757 | 384.652 | 384.965 | 361.459 | 379.436 | 361.427 | |
| 15.5 | 431.375 | 553.452 | 554.413 | 514.628 | 539.148 | 514.540 | |
| 20.5 | 557.753 | 716.269 | 718.403 | 658.556 | 688.595 | 658.389 | |
| 3227.7 ( (I)) | |||||||
| 1.5 | 0.01 | 0.018 | 0.016 | 0.037 | 0.043 | 0.037 | 0.028 |
| 2.5 | 0.041 | 0.073 | 0.062 | 0.147 | 0.172 | 0.147 | 0.106 |
| 3.5 | 0.104 | 0.182 | 0.155 | 0.365 | 0.428 | 0.367 | 0.262 |
| 4.5 | 0.212 | 0.364 | 0.308 | 0.726 | 0.851 | 0.730 | 0.518 |
| 5.5 | 0.376 | 0.634 | 0.538 | 1.263 | 1.479 | 1.269 | 0.901 |
| 10.5 | 2.868 | 3.899 | 3.292 | 7.549 | 8.793 | 7.586 | 5.500 |
| 15.5 | 17.882 | 11.636 | 9.757 | 21.603 | 24.958 | 21.703 | 16.353 |
| 20.5 | 25.142 | 20.897 | 44.260 | 50.639 | 44.455 | 35.079 | |
| (51) | (30) | (41) | (36) | (58) | (36) | ||
| 11247.3 ( (II)) | |||||||
| 1.5 | –0.063 | –0.029 | –0.018 | –0.079 | –0.068 | –0.035 | |
| 2.5 | –0.158 | –0.115 | –0.070 | –0.312 | –0.270 | –0.130 | |
| 3.5 | –0.316 | –0.287 | –0.175 | –0.772 | –0.669 | –0.315 | |
| 4.5 | –0.553 | –0.570 | –0.349 | –1.524 | –1.323 | –0.621 | |
| 5.5 | –3.509 | –0.992 | –0.607 | –2.625 | –2.281 | –1.071 | |
| 10.5 | –10.971 | –5.923 | –3.683 | –14.584 | –12.813 | –6.200 | |
| 15.5 | –25.100 | –16.893 | –10.754 | –37.569 | –33.476 | –17.064 | |
| 20.5 | –34.398 | –22.553 | –68.635 | –61.985 | –33.553 | ||
| (66) | (8) | (42) | (133) | (104) | |||
| 11931.7 ( (II)) | |||||||
| 0.5 | –13.167 | –20.843 | –19.810 | –20.843 | –19.810 | ||
| 1.5 | –26.318 | –41.656 | –39.601 | –41.607 | –39.551 | ||
| 2.5 | –33.439 | –62.409 | –59.354 | –62.212 | –59.154 | ||
| 3.5 | –52.514 | –83.073 | –79.052 | –82.586 | –78.558 | ||
| 4.5 | –65.529 | –103.620 | –98.675 | –102.663 | –97.701 | ||
| 5.5 | –78.467 | –124.022 | –118.207 | –122.383 | –116.534 | ||
| 10.5 | –141.459 | –223.062 | –213.942 | –214.366 | –204.816 | ||
| 15.5 | –200.108 | –315.698 | –305.327 | –294.915 | –282.616 | ||
| 20.5 | –252.309 | –401.264 | –391.461 | –366.782 | –352.056 | ||
III.2 Thorium fluoride cation
III.2.1 KRCI computation and input parameters
The expectation values of and , and the CI coefficients were calculated in this study. The employed CI model for the lowest two electronic states is abbreviated as TZ/SD6_CAS2in3_SDTQ8/6a.u. and for all other states TZ/SD6_CAS2in3_SD8/6a.u. and comprises a triple-zeta basis set [42]. Both CI models allow for single and double holes in all of the F 2p shells, a complete active space where all occupations with two electrons in the three thorium 7s and 5dδ spinors and up to double (SD) excitations or up to quadruple (SDTQ) excitations into all virtual spinors below a cutoff of a.u.
For the rotational-coupling calculations in the coupled electronic configurations are and . is the dominant configuration of the ground state, and is the only configuration that can contribute to the state and can couple with the configuration. The values of for the states with are set to 0, because we consider only the coupling between and the states coupled with it. The states represented with are treated as in the -doublet calculation. The values of for all basis functions are summarized in Table 6.
The choice of in a multireference system can be somewhat arbitrary. In this study, we employed and . The () corresponds to the ground state of the Th+ cation, and it is justified by comparing the electron affinity of F (27 400 ) and the ionization potential of Th+ (97 600 ). Another option is (), which is the first excited state of the cation at the energy of 63.3 from the ground state. The Mulliken population analysis of ThF+ spinors supports of an ionic-bonding like electronic configuration, -F-. The ground state of Th2+, does not significantly contribute to low-energy states of ThF+. In the calculation of the matrix elements of , we considered the bases with .
III.2.2 -doublet splitting
Table 2 lists the -splitting of the state of . Taking the electronic configuration of the coupled states into account (in the following denoted as “with configuration”), the -splitting is reduced to approximately half, which is a greater decrease than that predicted by the configuration coefficient : for , and for state, which are listed in Table 6. When we take the electronic configuration into account, the value is closer to the experiment than the value. However, all models listed in the table can qualitatively reproduce the experimental results. This indicates that our model is reliable in predicting at least the order of magnitude of -splitting for planning new experiments, even in systems like the present where a very complicated coupling mechanism makes an accurate calculation quite difficult. ’s relatively large -splitting – given that a state is concerned – is mainly due to the very small energy difference between the ground and states, amounting to only 315 (cf. Table 6). Our model can provide an opposite sign of the splitting because the relative sign of the correction basis function is not considered in Eq. (15).
| w/o configuration | with configuration | |||||
|---|---|---|---|---|---|---|
| exp. [43] | exp. [44] | |||||
| 1 | –4.4 | –8.8 | –2.2 | –4.5 | 5.29(5) | 10.0 |
| 2 | –13.2 | –26.3 | –6.7 | –13.4 | ||
| 3 | –26.3 | –52.6 | –13.4 | –26.9 | ||
| 4 | –43.8 | –87.7 | –22.4 | –44.8 | ||
| 5 | –65.8 | –131.5 | –33.6 | –67.1 | ||
| 10 | –241.1 | –482.1 | –123.1 | –246.1 | ||
| 15 | –525.9 | –1051.5 | –268.5 | –536.9 | ||
| 20 | –920.1 | –1839.2 | –469.8 | –939.5 | ||
III.3 Tantalum oxide cation
III.3.1 Input parameters
We employed as the coupled electronic configurations, contributing to both states and states (cf. Table 7). Another possible configuration pair would be and . However, we ignore the contributions from these configurations because the states to which the configuration mainly contributes are located in a higher energy range than those listed in Table 7. A total of fourteen electronic states with , and listed in Table 7 are employed for building the coupling matrix. Although also exhibits multireference character similar to , the complex with is a good choice here since the , and cations have , and ground states, respectively [45], all with .
| 7.003 | 0 | 0 | 0 | –0.291 | 0 | 0 | 0 | 0 | 0 | –0.012 | 0 | –0.012 | 0 | 0 | 0 | ||
|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
| 0 | 7.003 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | ||
| 0 | 0 | 3184 | 0 | 0 | 0 | 0 | 0 | –0.176 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | ||
| 0 | 0 | 0 | 3184 | 0 | 0 | 0 | 0 | 0 | –0.176 | 0 | 0 | 0 | 0 | 0 | 0 | ||
| –0.291 | 0 | 0 | 0 | 8935 | 0 | –10.085 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | –10.888 | 0 | ||
| 0 | 0 | 0 | 0 | 0 | 8935 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | ||
| 0 | 0 | 0 | 0 | –10.085 | 0 | 9381 | 0 | 0 | 0 | –0.428 | 0 | –0.428 | 0 | 0 | 0 | ||
| 0 | 0 | 0 | 0 | 0 | 0 | 0 | 9381 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | ||
| 0 | 0 | –0.176 | 0 | 0 | 0 | 0 | 0 | 17968 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | ||
| 0 | 0 | 0 | –0.176 | 0 | 0 | 0 | 0 | 0 | 17968 | 0 | 0 | 0 | 0 | 0 | 0 | ||
| –0.012 | 0 | 0 | 0 | 0 | 0 | –0.428 | 0 | 0 | 0 | 18683 | 0 | 0 | 0 | –0.462 | 0 | ||
| 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 18683 | 0 | 0 | 0 | 0 | ||
| –0.012 | 0 | 0 | 0 | 0 | 0 | –0.428 | 0 | 0 | 0 | 0 | 0 | 18854 | 0 | –0.462 | 0 | ||
| 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 18854 | 0 | 0 | ||
| 0 | 0 | 0 | 0 | –10.888 | 0 | 0 | 0 | 0 | 0 | –0.462 | 0 | –0.462 | 0 | 18946 | 0 | ||
| 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 18946 | ||
III.3.2 -doublet splitting
The coupling matrix in units of is presented in Table 3. The magnitude of the off-diagonal elements depends on the coefficients and defined in Eq. II.1. For example, the ratio between and is about 0.0267. This ratio is similar in amount to for the basis function of the ground state presented in Table 7. The small discrepancy is due to a small contribution from the basis function to the state, that is, the state does not purely consist of the basis function. The employed values of for the state with the excitation energy of 2.348 eV and correction basis function in the ground state are identical. Although a tiny coupling between and is also observed due to the correction term ( and basis functions, respectively), the coupling that dominantly causes the -splitting of state is the and states, through the matrix elements of the L-uncoupling term between and the analytical expression of which is shown in Eq. (II.2).
The -splitting of the state of are listed in Table 4. As to be seen, the employed electronic excitation energy does not affect the order of magnitude of the -splitting. The -splitting is much smaller than in ThF+ (–4.5 MHz, shown in Table 2). The reason for this is that in TaO+ (i) the energy difference between the states responsible for the -splitting ( and ) is much larger (ii) the contribution from the state to the state is smaller, and (iii) the contributions from the configurations to the corresponding states are smaller (cf. Table 7).
| 13 and 14 | only 14 | |
|---|---|---|
| 1 | 8.7 | 8.2 |
| 2 | 26.1 | 24.6 |
| 3 | 52.1 | 49.2 |
| 4 | 86.9 | 82.0 |
| 5 | 130.3 | 123.0 |
| 10 | 477.7 | 451.0 |
| 15 | 1042.4 | 984.0 |
| 20 | 1824.2 | 1721.9 |
III.4 Conditions for -splitting
Molecules with small -splitting are suitable for CP-violation search, since small external electric fields lead to full polarization. For diatomic molecules with a low-lying multiplet of states such as , , and , we can identify several decisive characteristics. One factor for achieving small -splitting is the energy difference between the (or and in case they are the target states) and states, because the -splitting decreases as the energy difference increases in perturbative expressions [31, 28]. Another factor is the strength of the spin-orbit interaction. The -splitting of the state becomes zero when there is no contribution to the state. This contribution is driven by the magnitude of SO coupling, a relativistic effect. From this point of view intermediately heavy molecular systems, e.g., and , have an advantage over heavier systems such as ThO and , although the nuclear charge is not the only factor that determines the size of the SO coupling. For instance, Table 5 shows larger mixing between and basis functions of PtH than the corresponding mixing in the case of . Finally, in the case of multi-reference systems, the rotational coupling between and basis functions is reduced when the electronic configurational overlap between them is small.
IV Conclusion
We present a theoretical model and calculations of the rotational-coupling effects in three molecular species, PtH, , and . Our model integrates the multi-reference four-component wavefunction and the -splitting based on Hund’s case (a). The matrix elements for the rotational-coupling Hamiltonian are constructed employing the electronic excitation energies and rotational constants as input parameters, and are diagonalized in an - and -type basis. In our model, the -splitting of the state occurs indirectly through the operators acting on and states. Thus, it is an -uncoupling effect entering through higher orders in perturbation theory. The dominant factors that determine the magnitude of the splitting of the target state are (i) the contribution of to the state (i.e., the strength of the spin-orbit coupling), (ii) the energy difference between the state and the state, (iii) the electronic configurational overlap of the and the basis functions. Our calculations of the -splitting qualitatively agree with the available experimental data for PtH and .
The -splitting of is in our present work predicted to be about kHz in the electronic ground state and for rotational quantum number . This small - (-) splitting can reduce the systematic uncertainty due to the external electric field. In addition, this is small enough to avoid the use of larger ring traps and the study of ion dynamics in the ring trap. On the other hand it may not be large enough to avoid depolarization during rotation ramp-up [46].
Our program is also applicable for estimating the -doubling of other target molecules for CP-violation search, such as TaN () [47, 48], WC () [49], and PbO () [50, 51, 52] molecules. -doublings from a Hund’s case (c) point of view [31, 53, 54] could be obtained with only minor modifications of the present formulation.
Acknowledgements.
We thank Dr. Yan Zhou (Las Vegas) for helpful discussions. A.S. acknowledges financial support from the Japan Society for the Promotion of Science (JSPS) KAKENHI (Grant No. 21K14643).*
Appendix A Electronic structure data
Tables 5-7 list the employed excitation energies, CI coefficients (), and configuration parameters (, defined in Eq. (II.1)) of the PtH, , and molecules, respectively. Table 6 lists both the singlet and triplet states, but only the singlet states are employed for the calculation of Table 2. Although our previous work [14] provided higher-energy electronic excited states, Table 7 presents only the states to which and orbitals dominantly contribute.
| Theory | Experiment | ||
| 23333Taken from states listed in Table 5. | 41444Taken from the zero-point energies shown in a file v-dependent_constants_PtH.xlsx in the Supplementary Material. | 40555Taken from Table VII. | |
| 0 | 0 | 0 | |
| 1479.2 | 2014.4 | ||
| 3414.2 | 3227.7 | 3224.89 | |
| 11625.9 | 11247.3 | 11581.55 | |
| 12208.4 | 11931.7 | ||
| KRCI | dominant | correction | configuration | |||||||
| energy | basis | basis | Th | |||||||
| 0 | 0.0000 | 0.0000 | 1.0000 | 0.87 | – | 0.0000 | 0 | 0.75 | ||
| 0.12 | ||||||||||
| 315666Experimental data [11] | 0.9822 | 1.9822 | 0.9911 | 0 | 0.1334 | 0.78 | 0.94 | |||
| 3395 | 0.0100 | 0.0100 | 0.9950 | 0 | 0.1000 | 0 | 0.58 | |||
| 0.12 | ||||||||||
| 6528777Calculated data (Table 9 of Ref. [55]) | 0.9471 | 0.9471 | 0.9732 | 0 | 0.2300 | 0.87 | 0.29 | |||
| 0.29 | ||||||||||
| 0.11 | ||||||||||
| 0.11 | ||||||||||
| 6639 | 0.1092 | 0.8908 | 0.9438 | 0.78 | 0.3305 | 0 | 0.41 | |||
| 0.18 | ||||||||||
| 0.16 | ||||||||||
| 6747 | 0.9485 | 0.9485 | 0.9739 | 0 | 0.2269 | 0.87 | 0.23 | |||
| 0.23 | ||||||||||
| 0.11 | ||||||||||
| 0.09 | ||||||||||
| 0.09 | ||||||||||
| 7490 | 0.9890 | 0.0110 | 0.9945 | 0 | 0.1049 | 0.78 | 0.74 | |||
| 7918 | 0.0003 | 0.9997 | 0.9998 | 0.20 | 0.0173 | 0 | 0.42 | |||
| 0.19 | ||||||||||
| 0.15 | ||||||||||
| Energy (eV) | KRCI | dominant | correction | configuration | |||||||
| Ref. 13 | Ref. 14 | basis | basis | Ta | |||||||
| 0.000 | 0.000 | –0.9992 | 1.9992 | 0.9996 | 0 | 0.0283 | 0.43 | 0.89 | |||
| 0.163 | 0.158 | 0.0000 | 2.0000 | 1.0000 | 0 | - | - | 0 | 0.60 | ||
| 0.29 | |||||||||||
| 0.405 | 0.395 | 0.9974 | 2.0026 | 0.9987 | 0 | 0.0510 | 1.00 | 0.89 | |||
| 0.466 | 0.414 | 0.0000 | 0.0000 | 1.0000 | 0 | - | - | 0 | 0.62 | ||
| 0.22 | |||||||||||
| 1.025 | 1.106 | 0.0000 | 0.0000 | 1.0000 | 1.00 | - | - | 0 | 0.50 | ||
| 0.30 | |||||||||||
| 1.043 | 1.162 | 0.9992 | 0.0008 | 0.9996 | 1.00 | 0.0283 | 0.43 | 0.88 | |||
| 1.421 | 1.342 | 0.0006 | 1.9994 | 0.9997 | 0 | 0.0245 | 0 | 0.57 | |||
| 0.26 | |||||||||||
| 1.804 | 0.0043 | 3.9957 | 0.9978 | 0 | 0.0656 | 1.00 | 0.88 | ||||
| 2.228 | –0.9965 | 2.9965 | 0.9982 | 1.00 | 0.0592 | 0 | 0.89 | ||||
| 2.315 | 0.9982 | –0.9982 | 0.9991 | 0 | 0.0424 | 1.00 | 0.44 | ||||
| 0.44 | |||||||||||
| 2.336 | –0.9982 | 0.9982 | 0.9991 | 0 | 0.0424 | 1.00 | 0.40 | ||||
| 0.40 | |||||||||||
| 2.348 | 0.0039 | 0.9961 | 0.9980 | 0.43 | 0.0624 | 1.00 | 0.50 | ||||
| 0.37 | |||||||||||
| 2.417 | 0.0000 | 0.0000 | 1.0000 | 0 | - | - | 0 | 0.47 | |||
| 0.13 | |||||||||||
| 0.13 | |||||||||||
| 2.543 | 0.0009 | 0.9991 | 0.9995 | 0.34 | 0.0300 | 1.00 | 0.39 | ||||
| 0.30 | |||||||||||
| 0.19 | |||||||||||
| 2.619 | 0.0053 | 2.9947 | 0.9973 | 1.00 | 0.0728 | 0 | 0.45 | ||||
| 0.45 | |||||||||||
| 2.717 | 1.0000 | 1.0000 | 1.0000 | 0 | - | - | 0 | 0.84 | |||
| 2.880 | 0.9951 | –0.9951 | 0.9975 | 1.00 | 0.0700 | 1.00 | 0.44 | ||||
| 0.44 | |||||||||||
| 2.892 | –0.9951 | 0.9951 | 0.9975 | 1.00 | 0.0700 | 1.00 | 0.44 | ||||
| 0.44 | |||||||||||
| 2.913 | 0.9982 | 1.0018 | 0.9991 | 1.00 | 0.0424 | 0 | 0.83 | ||||
| 3.018 | 0.9964 | 3.0036 | 0.9982 | 1.00 | 0.0600 | 0 | 0.90 | ||||
| 3.024 | 0.0004 | 0.9996 | 0.9998 | 0 | 0.0200 | 1.00 | 0.45 | ||||
| 0.27 | |||||||||||
| 0.11 | |||||||||||
| 3.629 | 0.0005 | 2.9995 | 0.9997 | 0 | 0.0224 | 0 | 0.43 | ||||
| 0.42 | |||||||||||
References
- Alarcon et al. [2022] R. Alarcon, J. Alexander, V. Anastassopoulos, T. Aoki, R. Baartman, S. Baeßler, L. Bartoszek, D. H. Beck, F. Bedeschi, R. Berger, M. Berz, H. L. Bethlem, T. Bhattacharya, M. Blaskiewicz, T. Blum, T. Bowcock, A. Borschevsky, K. Brown, D. Budker, S. Burdin, B. C. Casey, G. Casse, G. Cantatore, L. Cheng, T. Chupp, V. Cianciolo, V. Cirigliano, S. M. Clayton, C. Crawford, B. P. Das, H. Davoudiasl, J. de Vries, D. DeMille, D. Denisov, M. V. Diwan, J. M. Doyle, J. Engel, G. Fanourakis, R. Fatemi, B. W. Filippone, V. V. Flambaum, T. Fleig, N. Fomin, W. Fischer, G. Gabrielse, R. F. G. Ruiz, A. Gardikiotis, C. Gatti, A. Geraci, J. Gooding, B. Golub, P. Graham, F. Gray, W. C. Griffith, S. Haciomeroglu, G. Gwinner, S. Hoekstra, G. H. Hoffstaetter, H. Huang, N. R. Hutzler, M. Incagli, T. M. Ito, T. Izubuchi, A. M. Jayich, H. Jeong, D. Kaplan, M. Karuza, D. Kawall, O. Kim, I. Koop, W. Korsch, E. Korobkina, V. Lebedev, J. Lee, S. Lee, R. Lehnert, K. K. H. Leung, C.-Y. Liu, J. Long, A. Lusiani, W. J. Marciano, M. Maroudas, A. Matlashov, N. Matsumoto, R. Mawhorter, F. Meot, E. Mereghetti, J. P. Miller, W. M. Morse, J. Mott, Z. Omarov, L. A. Orozco, C. M. O’Shaughnessy, C. Ozben, S. Park, R. W. Pattie, A. N. Petrov, G. M. Piacentino, B. R. Plaster, B. Podobedov, M. Poelker, D. Pocanic, V. S. Prasannaa, J. Price, M. J. Ramsey-Musolf, D. Raparia, S. Rajendran, M. Reece, A. Reid, S. Rescia, A. Ritz, B. L. Roberts, M. S. Safronova, Y. Sakemi, P. Schmidt-Wellenburg, A. Shindler, Y. K. Semertzidis, A. Silenko, J. T. Singh, L. V. Skripnikov, A. Soni, E. Stephenson, R. Suleiman, A. Sunaga, M. Syphers, S. Syritsyn, M. R. Tarbutt, P. Thoerngren, R. G. E. Timmermans, V. Tishchenko, A. V. Titov, N. Tsoupas, S. Tzamarias, A. Variola, G. Venanzoni, E. Vilella, J. Vossebeld, P. Winter, E. Won, A. Zelenski, T. Zelevinsky, Y. Zhou, and K. Zioutas, Electric dipole moments and the search for new physics (2022), arXiv:2203.08103 [hep-ph].
- Roussy et al. [2023] T. S. Roussy, L. Caldwell, T. Wright, W. B. Cairncross, Y. Shagam, K. B. Ng, N. Schlossberger, S. Y. Park, A. Wang, J. Ye, and E. A. Cornell, An improved bound on the electron’s electric dipole moment, Science 381, 46 (2023).
- Fleig [2017a] T. Fleig, -odd and magnetic hyperfine-interaction constants and excited-state lifetime for HfF+, Phys. Rev. A 96, 040502 (2017a).
- Skripnikov [2017] L. V. Skripnikov, Communication: Theoretical study of HfF+ cation to search for the T,P-odd interactions, J. Chem. Phys. 147, 021101 (2017).
- Branco et al. [2012] G. C. Branco, R. G. Felipe, and F. R. Joaquim, Leptonic CP violation, Rev. Mod. Phys. 84, 515 (2012).
- Shindler [2021] A. Shindler, Flavor-diagonal CP violation: the electric dipole moment, Eur. Phys. J. D 57, 128 (2021).
- Ng et al. [2022a] K. B. Ng, Y. Zhou, L. Cheng, N. Schlossberger, S. Y. Park, T. S. Roussy, L. Caldwell, Y. Shagam, A. J. Vigil, E. A. Cornell, and J. Ye, Spectroscopy on the electron-electric-dipole-moment–sensitive states of ThF+, Phys. Rev. A 105, 022823 (2022a).
- Gresh et al. [2016] D. N. Gresh, K. C. Cossel, Y. Zhou, J. Ye, and E. A. Cornell, Broadband velocity modulation spectroscopy of ThF+ for use in a measurement of the electron electric dipole moment, J. Mol. Spectrosc. 319, 1 (2016).
- Denis et al. [2015a] M. Denis, M. Nørby, H. J. Aa. Jensen, A. S. P. Gomes, M. K. Nayak, S. Knecht, and T. Fleig, Theoretical study on ThF+, a prospective system in search of time-reversal violation, New J. Phys. 17, 043005 (2015a).
- Skripnikov and Titov [2015] L. V. Skripnikov and A. V. Titov, Theoretical study of ThF+ in the search for ,-violation effects: Effective state of a Th atom in ThF+ and ThO compounds, Phys. Rev. A 91, 042504 (2015).
- Barker et al. [2012] B. J. Barker, I. O. Antonov, M. C. Heaven, and K. A. Peterson, Spectroscopic investigations of ThF and ThF+, J. Chem. Phys. 136, 104305 (2012).
- Ng et al. [2025] K. B. Ng, S. Y. Park, A. Wang, A. Hartman, P. H. Hernandez, R. Kompella, L. Cheng, S. Malbrunot-Ettenauer, J. Ye, and E. A. Cornell, High-efficiency quantum-state detection of ThF+ with resonance-enhanced multiphoton asymmetric dissociation, arXiv [physics.atom-ph] (2025), arXiv:2508.04949 [physics.atom-ph] .
- Fleig [2017b] T. Fleig, TaO+ as a candidate molecular ion for searches of physics beyond the standard model, Phys. Rev. A 95, 022504 (2017b).
- Sunaga and Fleig [2022] A. Sunaga and T. Fleig, Spectroscopic and electric properties of the TaO+ molecule ion for the search of new physics: A platform for identification and state control, J. Quant. Spectrosc. and Rad. Transfer 288, 108229 (2022).
- Zhou et al. [2024] Y. Zhou, J. O. Island, and M. Grau, Quantum logic control and precision measurements of molecular ions in a ring trap: An approach for testing fundamental symmetries, Phys. Rev. A 109, 033107 (2024).
- Penyazkov et al. [2022] G. Penyazkov, L. V. Skripnikov, A. V. Oleynichenko, and A. V. Zaitsevskii, Effect of the neuron quadrupole distribution in the TaO+ cation, Chem. Phys. Lett. 793, 139448 (2022).
- Flambaum et al. [2014] V. V. Flambaum, D. DeMille, and M. G. Kozlov, Time-reversal symmetry violation in molecules induced by nuclear magnetic quadrupole moments, Phys. Rev. Lett. 113, 103003 (2014).
- Lackenby and Flambaum [2018] B. G. C. Lackenby and V. V. Flambaum, Time reversal violating magnetic quadrupole moment in heavy deformed nuclei, Phys. Rev. D 98, 115019 (2018).
- ACM [2018] Improved limit on the electric dipole moment of the electron, Nature 562, 355 (2018), ACME Collaboration.
- Baron et al. [2017] J. Baron, W. C. Campbell, D. DeMille, J. M. Doyle, G. Gabrielse, Y. V. Gurevich, P. W. Hess, N. R. Hutzler, E. Kirilov, I. Kozyryev, B. R. O’Leary, C. D. Panda, M. F. Parsons, B. Spaun, A. C. Vutha, A. D. West, E. P. West, and A. Collaboration, Methods, analysis, and the treatment of systematic errors for the electron electric dipole moment search in thorium monoxide, New J. Phys. 19, 073029 (2017).
- Kobayashi and Maskawa [1973] M. Kobayashi and T. Maskawa, CP violation in the renormalizable theory of weak interaction, Prog. Theor. Phys. 49, 652 (1973).
- Fleig and Marian [1994] T. Fleig and C. M. Marian, Relativistic all-electron ab-initio calculations on the platinum hydride molecule, Chem. Phys. Lett. 222, 267 (1994).
- Fleig and Marian [1996] T. Fleig and C. M. Marian, Ab InitioCalculation of -splittings and rovibronic states of the PtH and PtD molecules, J. Mol. Spectrosc. 178, 1 (1996).
- Brown et al. [1975a] J. M. Brown, J. T. Hougen, K. Huber, J. Johns, I. Kopp, H. Lefebvre-Brion, A. J. Merer, D. A. Ramsay, J. Rostas, and R. Zare, The labeling of parity doublet levels in linear molecules, Journal of Molecular Spectroscopy 55, 500 (1975a).
- Brown et al. [1979] J. Brown, E. Colbourn, J. Watson, and F. Wayne, An effective hamiltonian for diatomic molecules . ab initio calculations of parameters of HCl, Journal of Molecular Spectroscopy 74, 294 (1979).
- de Vivie et al. [1988] R. de Vivie, C. M. Marian, and S. D. Peyerimhoff, A general procedure for the theoretical study of the -doubling: Application to the X states of OH and SH, Mol. Phys. 63, 3 (1988).
- Kozlov [2009] M. G. Kozlov, -doublet spectra of diatomic radicals and their dependence on fundamental constants, Phys. Rev. A 80, 022118 (2009).
- Fan and Cheng [2025] X. Fan and L. Cheng, Nuclear-electric-quadrupole-moment-induced parity doubling in molecules for symmetry-violation searches, Phys. Rev. A 112, 022823 (2025).
- Gordon and Field [2025] R. J. Gordon and R. W. Field, The energy denominator effect in lambda-doubling, J. Chem. Phys. 163, 10.1063/5.0277788 (2025).
- Marian [1995] C. M. Marian, An approach to the calculation of ‐splittings in diatomic molecules with strongly coupled electronic states and its application to NiH and NiD, Ber. Bunsenges. Phys. Chem. 99, 254 (1995).
- Lefbvre-Brion and Field [1986] H. Lefbvre-Brion and R. W. Field, Perturbations in the Spectra of Diatomic Molecules (Academic Press Inc. (London) Ltd., 1986).
- Meerts and Dymanus [1972] W. L. Meerts and A. Dymanus, The hyperfine -doubling spectrum of 14N16O and 15N16O, J. Mol. Spectrosc. 44, 320 (1972).
- Meerts [1976] W. L. Meerts, A theoretical reinvestigation of the rotational and hyperfine lambda doubling spectra of diatomic molecules with a state: the spectrum of no, Chem. Phys. 14, 421 (1976).
- Brown et al. [1987] J. M. Brown, A. S. C. Cheung, and A. J. Merer, -type doubling parameters for molecules in electronic states, 124, 464 (1987).
- Brown et al. [1975b] J. M. Brown, J. T. Hougen, K. Huber, J. Johns, I. Kopp, H. Lefebvre-Brion, A. J. Merer, D. A. Ramsay, J. Rostas, and R. Zare, The labeling of parity doublet levels in linear molecules, Journal of Molecular Spectroscopy 55, 500 (1975b).
- Hougen [1970] J. T. Hougen, The calculation of rotational energy levels and rotational line intensities in diatomic molecules, Vol. 115 (US National Bureau of Standards, 1970).
- Born and Oppenheimer [1927] M. Born and R. Oppenheimer, Ann. Phys. 84, 457 (1927).
- Bransden and Joachain [2003] B. H. Bransden and C. J. Joachain, Physics of Atoms and Molecules (Pearson Education Limited, Edinburgh Gate, Harlow, Essex CM20 2JE, England, 2003).
- [39] The same molecular spectroscopic terms ( and ) can be obtained from the combinations of Pt’s and H’s . However, to obtain we have to incorporate , whose excitation energy is 6567 [56]. It is located on much higher energy level than , whose excitation energy is 776 [56].
- Mccarthy et al. [1993] M. C. Mccarthy, R. W. Field, R. Engleman, and P. F. Bernath, Laser and fourier transform spectroscopy of PtH and PtD, J. Mol. Spectrosc. 158, 208 (1993).
- Irikura [2023] K. K. Irikura, Ab initio spectroscopy and thermochemistry of diatomic platinum hydride, PtH, J. Chem. Phys. 158, 174312 (2023).
- Dyall [2007] K. G. Dyall, Relativistic double-zeta, triple-zeta, and quadruple-zeta basis sets for the actinides Ac-Lr, Theoret. Chem. Acc. 117, 491 (2007).
- Ng et al. [2022b] K. B. Ng, Y. Zhou, L. Cheng, N. Schlossberger, S. Y. Park, T. S. Roussy, L. Caldwell, Y. Shagam, A. J. Vigil, E. A. Cornell, and J. Ye, Spectroscopy on the electron-electric-dipole-moment–sensitive states of ThF+, Phys. Rev. A 105, 022823 (2022b).
- Ng [2023] K. Ng, The eEDM experiment: concept, design, and characterization, Ph.D. thesis, Boulder (2023).
- [45] J. E. Sansonetti, W. C. Martin, and S. L. Young, Handbook of Basic Atomic Spectroscopic Data. Available: http://physics.nist.gov/PhysRefData/Handbook/index.html. National Institute of Standards and Technology, Gaithersburg, MD.
- [46] Yan Zhou (Las Vegas), private communication.
- Skripnikov et al. [2015] L. V. Skripnikov, A. N. Petrov, N. S. Mosyagin, A. V. Titov, and V. V. Flambaum, TaN molecule as a candidate for the search for a T, P -violating nuclear magnetic quadrupole moment, Phys. Rev. A 92, 012521 (2015).
- Fleig et al. [2016] T. Fleig, M. K. Nayak, and M. G. Kozlov, TaN, a molecular system for probing P , T -violating hadron physics, Phys. Rev. A 93, 012505 (2016).
- Lee et al. [2009] J. Lee, E. R. Meyer, R. Paudel, J. L. Bohn, and A. E. Leanhardt, An electron electric dipole moment search in the X ground state of tungsten carbide molecules, J. Mod. Opt. 56, 2005 (2009).
- DeMille et al. [2000] D. DeMille, F. Bay, S. Bickman, D. Kawall, D. Krause, S. E. Maxwell, and L. R. Hunter, Investigation of PbO as a system for measuring the electric dipole moment of the electron, Phys. Rev. A 61, 052507 (2000).
- Kozlov and DeMille [2002] M. G. Kozlov and D. DeMille, Enhancement of the electric dipole moment of the electron in PbO, Phys. Rev. Lett. 89, 133001 (2002).
- Eckel et al. [2013] S. Eckel, P. Hamilton, E. Kirilov, H. W. Smith, and D. Demille, Search for the electron electric dipole moment using -doublet levels in PbO, Phys. Rev. A 87, 052130 (2013).
- Veseth [1973a] L. Veseth, Hund’s coupling case (c) in diatomic molecules. I. theory, J. Phys. B: Atom. Mol. 6, 1473 (1973a).
- Veseth [1973b] L. Veseth, Hund’s coupling case (c) in diatomic molecules. II. examples, J. Phys. B: Atom. Mol. 6, 1484 (1973b).
- Denis et al. [2015b] M. Denis, M. S. Norby, H. J. A. Jensen, A. S. P. Gomes, M. K. Nayak, S. Knecht, and T. Fleig, Theoretical study on ThF+, a prospective system in search of time-reversal violation, New J. Phys. 17, 043005 (2015b).
- [56] A. Kramida, Y. Ralchenko, and N. A. T. J. Reader and, NIST Atomic Spectra Database (version 5.1). Available: http://physics.nist.gov/asd [accessed: Friday, 07-Nov-2025 12:38:20 EST]. National Institute of Standards and Technology, Gaithersburg, MD.