An experiment to search for the electron electric dipole moment (eEDM) on the metastable H3Δ1 state of ThO molecule was proposed and now prepared by the ACME Collaboration [http://www.electronedm.org]. To interpret the experiment in terms of eEDM and dimensionless constant kT, P characterizing the strength of the T,P-odd pseudoscalar–scalar electron–nucleus neutral current interaction, an accurate theoretical study of an effective electric field on electron, Eeff, and a parameter of the T,P-odd pseudoscalar–scalar interaction, WT, P, in ThO is required. We report our results for Eeff (84 GV/cm) and WT, P (116 kHz) together with the hyperfine structure constant, molecule frame dipole moment, and H3Δ1X1Σ+ transition energy, which can serve as a measure of reliability of the obtained Eeff and WT, P values. Besides, our results include a parity assignment and evaluation of the electric-field dependence for the magnetic g factors in the Ω-doublets of H3Δ1.

One of the most intriguing fundamental problems of modern physics is the search for a permanent electric dipole moment (EDM) of elementary particles. A nonzero value of EDMs implies manifestation of interactions which are not symmetric with respect to both time (T) and spatial (P) inversions (T,P-odd interactions). Particularly, the observation of electron EDM (eEDM) at the level significantly greater than 10−38e cm would indicate the presence of a “new physics” beyond the Standard model. Popular extensions of the Standard model predict the magnitude of the eEDM at the level of 10−26 to 10−29e cm.1 The most rigid upper bound on the eEDM is attained in the experiments on a beam of YbF molecular radicals2 (1.05 × 10−27e cm) and in the measurements with atomic Tl beam3 (1.6 × 10−27e cm).

Nowadays a number of other prospective experiments are suggested and in part prepared4–6 which promise to achieve a sensitivity to eEDM up to 10−29 to 10−30e cm. One of the most promising experiments towards the measurement of eEDM is proposed and now prepared on the metastable 3Δ1 state of the thorium monoxide (ThO) molecule by ACME Collaboration (groups of DeMille, Gabrielse, and Doyle).7,8 A very high sensitivity to eEDM is expected in the nearest future, up to an order of magnitude and more than that attained in the YbF and Tl experiments, due to some unique combination of experimental advantages of the molecule. Even the value for eEDM compatible with zero will lead to serious consequences for the modern theory of fundamental symmetries.

To interpret the results of the corresponding molecular ThO (or other diatomics) experiment in terms of the eEDM one should know a parameter usually called “the effective electric field on electron,” Eeff, which cannot be measured. To obtain Eeff one can evaluate an expectation value of some T,P-odd operator (discussed in Refs. 9–11):

(1)

where de is the value of eEDM, Ψ is the wave function of the considered state of ThO, and

$\Omega = \langle \Psi |\bm {J}\cdot \bm {n}|\Psi \rangle$
Ω=Ψ|J·n|Ψ⁠, J is the total electronic momentum, n is the unit vector along the molecular axis directed from Th to O (Ω = 1 for the considered 3Δ1 state of ThO),

(2)

E is the inner molecular electric field, and

$\bm {\sigma }$
σ are the Pauli matrices. In these designations Eeff = Wd|Ω|.

In addition to the interaction given by operator (2) there is a T,P-odd pseudoscalar–scalar electron–nucleus neutral currents interaction with the dimensionless constant kT, P. The interaction is given by the following operator (see Ref. 12):

(3)

where G is the Fermi constant, γ0 and γ5 are the Dirac matrices, and

$n(\textbf {r})$
n(r) is the nuclear density normalized to unity. To extract the fundamental kT, P constant from an experiment, one needs to know the factor WT, P that is determined by the electronic structure of a studied molecule on a given nucleus:

(4)

Similarly to Eeff, the WT, P parameter cannot be measured and have to be obtained from a molecular electronic structure calculation.

A commonly used way of verification of the theoretical Eeff and WT, P values is to calculate “on equal footing” (the same approximation for the wave function) those molecular characteristics (properties or effective Hamiltonian parameters) which have comparable to Eeff and WT, P sensitivity to variations of wave function in core region but, in contrast, can be measured. Similar to Eeff and WT, P these parameters should be sensitive to a change of densities of the valence electrons in atomic cores. The hyperfine structure constant, A||, is traditionally used as such a parameter (e.g., see Ref. 13) and this is a valid touchstone for the ThO case as well. To obtain A|| on Th in the ThO molecule theoretically, one can evaluate the following matrix element:

(5)

where μTh is the magnetic moment of an isotope of Th nucleus having spin I.

To validate our present study of Eeff, WT, P, and the hyperfine structure constant A|| for the 3Δ1 state of 229ThO, we have also performed calculations of the H3Δ1X1Σ transition energy and the molecule-frame dipole moment for the 3Δ1 state.

Recently, spectroscopy of ThO was investigated both theoretically and experimentally in a series of papers.5,14–16 First estimates of Eeff in the 3Δ1 state of ThO were performed in Ref. 5. The experimental measurements of molecule-frame dipole moments of the ground X1Σ+ and excited E states were performed in Ref. 14. In Ref. 15 a theoretical study of the potential energy curve and electric properties of the ground ThO state were performed. A series of relativistic calculations of the spectroscopic parameters of the ground and excited states were performed in Ref. 16. In the present paper we concentrate on the precise two-component relativistic coupled clusters calculations of the complementary data (Eeff, WT, P) which are most relevant for the direct interpretation of ThO experiment in terms of the electron EDM and constant of T,P-odd pseudoscalar–scalar electron–nucleus interaction.

The evaluation of Eeff, WT, P, and A|| is usually a challenging problem for modern ab initio methods when studying systems containing heavy transition metals, lanthanides and, particularly, actinides (such as Th in the present consideration). An accurate theoretical investigation of such systems should take into account both the relativistic and correlation effects with the best to-date accuracy. It follows from Eqs. (2)–(5) that the operators related to Eeff, WT, P, and A|| are essentially localized in the atomic core region. On the other hand, the main contribution to the corresponding matrix elements is due to the valence electrons since contributions from the closed inner-core shells compensate each other in most cases of practical interest for the operators dependent on the total angular momentum and spin. It was shown by our group (see Ref. 11 and references therein) that the problem of computation of such characteristics can be significantly simplified by splitting the calculation into two steps. In the first step the electron correlation for valence electrons (and outer-core electrons for better accuracy) is taken into account in a molecular calculation using some method of electron correlation treatment (such as coupled clusters, configuration interaction, etc.), whereas the core (inner-core) electrons are excluded from this calculation using the generalized relativistic effective core potential (GRECP),17,18 which yields an accurate valence region wave function in the most economical way. Second, since the inner-core parts of the valence one-electron “pseudo-wavefunctions” are smoothed within the GRECP method, they have to be recovered using some core-restoration method.11 The non-variational restoration is based on a proportionality (scaling) of valence and virtual spinors in the inner-core region of heavy atoms (e.g., see Ref. 18 for details). The two-step approach has been recently used in Refs. 19–23 for calculation of a number of characteristics, such as hyperfine structure constants, electron electric dipole moment enhancement factor, etc., in molecules and atoms. Besides it has been extended to the case of crystals in Ref. 24.

Recently24 we have developed a code of nonvariational restoration which has been interfaced to the DIRAC1225 and the MRCC26 codes. These codes are used in the present paper. Scalar-relativistic calculations (i.e., without spin-orbit terms in the GRECP operator) were performed using CFOUR code.27 

The 1s–4f inner-core electrons of Th were excluded from molecular correlation calculations using the valence (semi-local) version of GRECP17 operator. Thus, the outermost 38 electrons were treated explicitly. The basis set for Th was constructed using the generalized correlated scheme.28 It can be written as (30,20,17,11,4,1)/[30,8,6,4,4,1], so the only p, d, and f Gaussians are contracted. For oxygen the aug-ccpVQZ basis set29 with removed two g-type basis functions was employed, i.e., we used the (13,7,4,3)/[6,5,4,3] basis set.

Calculations of the transition energy between the ground X1Σ+ and excited H3Δ1 states as well as the molecule-frame dipole moment, Eeff, WT, P, and A|| constants for the 3Δ1 state of ThO were performed using the single reference two-component relativistic coupled clusters with single, double, and perturbative treatment of triple cluster amplitudes (CCSD(T)).30,39 In addition, the basis set enlargement corrections to the considered parameters were also calculated. For this we have performed (i) scalar-relativistic CCSD(T) calculation using the same basis set as used for the two-component calculation; (ii) scalar-relativistic CCSD(T) calculation using the extended basis set on Th (with added f, g, h, and i Gaussians). Corrections were estimated as a difference between the values of the corresponding parameters. The experimental equilibrium internuclear distances31,32 (3.478 a.u. for X1Σ+ state and 3.511 a.u. for H3Δ1 state) were used in these calculations. The results are given in Table I. Optimized (at the scalar-relativistic CCSD(T) level and the enlarged basis set on Th) internuclear distances are close to the experimental ones: 3.482 a.u. for X1Σ+ and 3.515 a.u. for H3Δ1. The harmonic frequencies are also in a very good agreement with experiment:31,32 for X1Σ+ the calculated value is 895 cm−1 and experimental is 896 cm−1, while for H3Δ1 the calculated value is 857 cm−1 and experimental is 864 cm−1.

Table I.

The calculated values of transition energy (Te), molecule-frame dipole moment (d), effective electric field (Eeff), parameter of the T,P-odd pseudoscalar–scalar electron–nucleus neutral currents interaction (WT, P), and hyperfine structure constant (A||) of the H3Δ1 state of ThO using the coupled clusters methods.

 TedEeffWT, PA||
Method(cm−1)(debye)(GV/cm)(kHz)(
$\frac{\mu _{\rm Th}}{\mu _{\rm N}}\,$
μ Th μN
MHz)
2c-CCSD 5510 4.27 87 118 −2958 
2c-CCSD(T) 6120 4.22 84 116 −2885 
2c-CCSD(T) 5808 4.32 84 116 … 
+ basis corr.           
Experiment 5321 4.24 ± 0.1 … … … 
  31 and 32  8        
 TedEeffWT, PA||
Method(cm−1)(debye)(GV/cm)(kHz)(
$\frac{\mu _{\rm Th}}{\mu _{\rm N}}\,$
μ Th μN
MHz)
2c-CCSD 5510 4.27 87 118 −2958 
2c-CCSD(T) 6120 4.22 84 116 −2885 
2c-CCSD(T) 5808 4.32 84 116 … 
+ basis corr.           
Experiment 5321 4.24 ± 0.1 … … … 
  31 and 32  8        

In the scalar-relativistic CCSD(T) calculations we have found that Eeff only very slightly depends on the internuclear distance: it has a maximum at 3.48 a.u., while at the interatomic distances of 3.4 a.u. and 3.56 a.u. it decreases by 0.1 GV/cm with respect to the maximal value.

The calculated value of transition energy is in a very good agreement with experimental datum, the deviation is on the level of accuracy attained earlier by our group for compounds of transition metals and lanthanides. It is also in a good agreement with previous calculation performed in Ref. 16.

It was recently shown in Ref. 33 that the magnetic moment of 229Th nucleus determined earlier34 is inaccurate. Therefore, A|| is given in Table I in the units of (μThN) MHz (where μN is the nuclear magneton) in Table I to exclude the uncertainty of μTh from our result. One can see from Table I that a good convergence of Eeff with respect to both the basis set enlargement and correlation level is achieved. Taking into account the results from Table I as well as our earlier studies within the two-step procedure (e.g., see Ref. 22) with calculating the Eeff, WT, P, and A|| we expect that the theoretical uncertainty for our final values of the constants is smaller than 15%. Unfortunately, there are no experimental data on A|| up to now. Therefore, the corresponding indirect experimental verification of accuracy of Eeff (see above) cannot be performed to-date and further experimental measurements of A|| are required.

In the eEDM search experiment on the ThO molecule, the eEDM induced Stark splitting between the J = 1, M = ± 1 states of e (parity is ( − 1)J) or f (parity is −( − 1)J) levels of the Ω-doublet is measured. The H3Δ1 state has a very small magnetic moment, μH[ThO] = 8.5(5) × 10−3μB,8 where μB is the Bohr magneton. The latter is a benefit for suppressing systematic effects due to spurious magnetic field. In a polarized molecule the e and f levels have opposite signs of Eeff and almost identical g factors. Therefore, when taking the difference between the splitting for e and f levels further suppression of the systematics is possible.35 The small difference between g factors, Δg, comes from interactions of H3Δ1 with 0+ and 0 electronic states.21 Our calculations show that being presented in the ΛS coupling scheme, the spin-orbit mixed H state of ThO has the main contribution (more than 95%) from the 3Δ1 configurations (see also Ref. 16). Therefore, due to the identity

$\langle \Psi _{^3\Delta _1}|S^e_+ |\Psi _{n0^\pm }\rangle \equiv 0$
ΨΔ13|S+e|Ψn0±0⁠, where Se is the electronic spin operator, for pure ΛS state, the inequality
$|\langle \Psi _{H^3\Delta _1}|S^e_+ |\Psi _{n0^\pm }\rangle | \ll 1$
|ΨH3Δ1|S+e|Ψn0±|1
holds with a good accuracy. This inequality gives sufficient condition for Δg to be determined by the energy splitting between the top and bottom levels of the Ω-doublet.21,22 The rotational analysis given in Ref. 32 for P(Ω = 0) − H3Δ1 and O(Ω = 0) − H3Δ1 bands has shown that the Ω-doublet spacing (Δ = |E(J = 1) − E(J = 1+)|) in H3Δ1 is 350–470 kHz. At the moment, the parity assignment for electronic states P and O is yet unclear. According to Ref. 36, P is 0 and O is 0+. The latter, as can be shown, indicates that e states are the top levels whereas the f states are the bottom levels of the Ω-doublets for H3Δ1.

The microwave spectroscopy data37 obtained later confirm our conclusion about the levels ordering and finally give

$E(J=1^-) - E(J=1^+)=\rm {362}\pm 10$
E(J=1)E(J=1+)=362±10 kHz. In Fig. 1 the calculated g factors for the J = 1 levels of ThO H3Δ1 state are given as functions of the laboratory electric field. The lowest value, Δg = 2.7 × 10−6, is attained at the electric field 4.4 V/cm. Note that the molecule is almost completely polarized at the electric field larger than 3 V/cm.

FIG. 1.

Calculated g-factor curves for J = 1 rotational level of 232Th16O.

FIG. 1.

Calculated g-factor curves for J = 1 rotational level of 232Th16O.

Close modal

It should be noted that there is a discrepancy in sign conventions of Eeff in different publications. Therefore, to avoid such an ambiguity, the energy levels of ThO in the eEDM experiment with J = 1 are shown in Fig. 2 for the case of a positive de value. Solid lines are the levels in the external electric field, dotted lines are the levels with eEDM shift added. According to Eqs. (1) and (2), the eEDM shift for a completely polarized molecule is determined by the expression +deEeffΩ, where Ω = Mnz, M is the electron-rotational angular momentum projection on the laboratory axis z. Note that for the opposite direction of n (from O to Th), the signs of all the Ω in Fig. 2 will be opposite. This does not affect the directions of eEDM shift, since the Eeff with n directed from O to Th will have the opposite sign (negative).

FIG. 2.

Energy levels of ThO molecules in eEDM experiment. Solid lines are the levels in external electric field and dotted lines are the levels with EDM shift added. The dashed arrow is unit vector along the molecular axis directed from Th to O.

FIG. 2.

Energy levels of ThO molecules in eEDM experiment. Solid lines are the levels in external electric field and dotted lines are the levels with EDM shift added. The dashed arrow is unit vector along the molecular axis directed from Th to O.

Close modal

The parameters Eeff and WT, P, which are required to interpret experimental measurements on the H3Δ1 state of ThO molecule in terms of fundamental quantities are calculated. Though the previous estimation of Eeff made in Ref. 5 is only 25% more than our final value, the good agreement can be rather considered as “fortunate” since the accuracy of semiempirical estimates for the systems like ThO having a very complicated electronic structure is severely limited, see Ref. 19. In turn, a reliable ab initio calculation of ThO is on the threshold of current possibilities of computational methods and we estimate the accuracy of our calculation of Eeff by 15% only. Nevertheless, even such accuracy is important to establish a reliable eEDM estimate in the ongoing ThO experiment compared to the measured upper bounds on eEDM in Tl3 and YbF2 experiments.

We are grateful to N. S. Mosyagin for providing us with the GRECP for Th.38 This work is supported by the SPbU Fundamental Science Research grant from Federal Budget No. 0.38.652.2013, RFBR Grant No. 13-02-01406, and partially by the RFBR Grant No. 13-03-01307a. L.S. is also grateful to the Dmitry Zimin “Dynasty” Foundation. The molecular calculations were performed on the Supercomputer “Lomonosov.”

1.
E. D.
Commins
,
Adv. At., Mol., Opt. Phys.
40
,
1
(
1998
).
2.
J. J.
Hudson
,
D. M.
Kara
,
I. J.
Smallman
,
B. E.
Sauer
,
M. R.
Tarbutt
, and
E. A.
Hinds
,
Nature (London)
473
,
493
(
2011
).
3.
B. C.
Regan
,
E. D.
Commins
,
C. J.
Schmidt
, and
D.
DeMille
,
Phys. Rev. Lett.
88
,
071805
(
2002
).
4.
A.
Leanhardt
,
J.
Bohn
,
H.
Loh
,
P.
Maletinsky
,
E.
Meyer
,
L.
Sinclair
,
R.
Stutz
, and
E.
Cornell
,
J. Mol. Spectrosc.
270
,
1
(
2011
).
5.
E. R.
Meyer
and
J. L.
Bohn
,
Phys. Rev. A
78
,
010502
(R) (
2008
).
6.
K. C.
Cossel
,
D. N.
Gresh
,
L. C.
Sinclair
,
T.
Coffey
,
L. V.
Skripnikov
,
A. N.
Petrov
,
N. S.
Mosyagin
,
A. V.
Titov
,
R. W.
Field
,
E. R.
Meyer
 et al.,
Chem. Phys. Lett.
546
,
1
(
2012
).
7.
A. C.
Vutha
,
W. C.
Campbell
,
Y. V.
Gurevich
,
N. R.
Hutzler
,
M.
Parsons
,
D.
Patterson
,
E.
Petrik
,
B.
Spaun
,
J. M.
Doyle
,
G.
Gabrielse
 et al.,
J. Phys. B
43
,
074007
(
2010
).
8.
A. C.
Vutha
,
B.
Spaun
,
Y. V.
Gurevich
,
N. R.
Hutzler
,
E.
Kirilov
,
J. M.
Doyle
,
G.
Gabrielse
, and
D.
DeMille
,
Phys. Rev. A
84
,
034502
(
2011
).
9.
M. G.
Kozlov
,
V. I.
Fomichev
,
Yu. Yu.
Dmitriev
,
L. N.
Labzovsky
, and
A. V.
Titov
,
J. Phys. B
20
,
4939
(
1987
).
10.
M.
Kozlov
and
L.
Labzowsky
,
J. Phys. B
28
,
1933
(
1995
).
11.
A. V.
Titov
,
N. S.
Mosyagin
,
A. N.
Petrov
,
T. A.
Isaev
, and
D. P.
DeMille
,
Prog. Theor. Chem. Phys.
B15
,
253
(
2006
).
13.
M. G.
Kozlov
,
A. V.
Titov
,
N. S.
Mosyagin
, and
P. V.
Souchko
,
Phys. Rev. A
56
,
R3326
(
1997
).
14.
F.
Wang
,
A.
Le
,
T. C.
Steimle
, and
M. C.
Heaven
,
J. Chem. Phys.
134
,
031102
(
2011
).
15.
A. A.
Buchachenko
,
J. Chem. Phys.
133
,
041102
(
2010
).
16.
J.
Paulovič
,
T.
Nakajima
,
K.
Hirao
,
R.
Lindh
, and
P. A.
Malmqvist
,
J. Chem. Phys.
119
,
798
(
2003
).
17.
N. S.
Mosyagin
,
A. V.
Zaitsevskii
, and
A. V.
Titov
,
Int. Rev. At. Mol. Phys.
1
,
63
(
2010
).
18.
A. V.
Titov
and
N. S.
Mosyagin
,
Int. J. Quantum Chem.
71
,
359
(
1999
).
19.
L. V.
Skripnikov
,
A. N.
Petrov
,
A. V.
Titov
, and
N. S.
Mosyagin
,
Phys. Rev. A
80
,
060501
(R) (
2009
).
20.
L.
Skripnikov
,
A.
Titov
,
A. N.
Petrov
,
N. S.
Mosyagin
, and
O. P.
Sushkov
,
Phys. Rev. A
84
,
022505
(
2011
).
21.
A. N.
Petrov
,
Phys. Rev. A
83
,
024502
(
2011
).
22.
J.
Lee
,
J.
Chen
,
L. V.
Skripnikov
,
A. N.
Petrov
,
A. V.
Titov
,
N. S.
Mosyagin
, and
A. E.
Leanhardt
,
Phys. Rev. A
87
,
022516
(
2013
).
23.
A. N.
Petrov
,
L. V.
Skripnikov
,
A. V.
Titov
, and
R. J.
Mawhorter
,
Phys. Rev. A
88
,
010501
(R) (
2013
).
24.
L. V.
Skripnikov
and
A. V.
Titov
, “
Theoretical study of the parity and time reversal violating interaction in solids
,”
Phys. Chem. Phys.
(to be published); e-print: arXiv:1308.0163.
25.
DIRAC, a relativistic ab initio electronic structure program, Release DIRAC12 (2012), written by
H. J. Aa.
Jensen
,
R.
Bast
,
T.
Saue
, and
L.
Visscher
, with contributions from
V.
Bakken
,
K. G.
Dyall
,
S.
Dubillard
,
U.
Ekström
,
E.
Eliav
,
T.
Enevoldsen
,
T.
Fleig
,
O.
Fossgaard
,
A. S. P.
Gomes
,
T.
Helgaker
,
J. K.
Lærdahl
,
Y. S.
Lee
,
J.
Henriksson
,
M.
Iliaš
,
Ch. R.
Jacob
,
S.
Knecht
,
S.
Komorovský
,
O.
Kullie
,
C. V.
Larsen
,
H. S.
Nataraj
,
P.
Norman
,
G.
Olejniczak
,
J.
Olsen
,
Y. C.
Park
,
J. K.
Pedersen
,
M.
Pernpointner
,
K.
Ruud
,
P.
Sałek
,
B.
Schimmelpfennig
,
J.
Sikkema
,
A. J.
Thorvaldsen
,
J.
Thyssen
,
J.
van Stralen
,
S.
Villaume
,
O.
Visser
,
T.
Winther
, and
S.
Yamamoto
(see http://www.diracprogram.org).
26.
MRCC, a quantum chemical program suite written by
M.
Kállay
,
Z.
Rolik
,
I.
Ladjánszki
,
L.
Szegedy
,
B.
Ladóczki
,
J.
Csontos
, and
B.
Kornis
.
See also
Z.
Rolik
and
M.
Kállay
,
J. Chem. Phys.
135
,
104111
(
2011
), as well as www.mrcc.hu.
[PubMed]
27.
J. F.
Stanton
,
J.
Gauss
,
M. E.
Harding
,
P. G.
Szalay
 et al. (
2011
), “CFOUR: a program package for performing high-level quantum chemical calculations on atoms and molecules,” see http://www.cfour.de.
28.
N. S.
Mosyagin
,
E.
Eliav
,
A. V.
Titov
, and
U.
Kaldor
,
J. Phys. B
33
,
667
(
2000
).
29.
R. A.
Kendall
,
T. H.
Dunning
 Jr.
, and
R. J.
Harrison
,
J. Chem. Phys.
96
,
6796
(
1992
).
30.
For both the 1Σ+ and 3Δ1 states the same set of one-electron spinors obtained in the two-component Hartree-Fock calculation of the 1Σ+ state were employed. Particularly, the reference determinant for the single-reference coupled clusters calculation of 3Δ1 state was constructed from this set of spinors. Such choice of one-electron functions is justified for the case of treatment by the coupled clusters approaches which include single clusters amplitudes. The latter are required to account for relaxation effects (see, e.g., Ref. 39).
31.
K. P.
Huber
and
G.
Herzberg
,
Constants of Diatomic Molecules
(
Van Nostrand-Reinhold
,
New York
,
1979
).
32.
G.
Edvinsson
and
A.
Lagerqvist
,
Phys. Scr.
30
,
309
(
1984
).
33.
M. S.
Safronova
,
U. I.
Safronova
,
A. G.
Radnaev
,
C. J.
Campbell
, and
A.
Kuzmich
, “
Magnetic dipole and electric quadrupole moments of the 229Th nucleus
,”
Phys. Atom. Phys.
(
2013
); e-print: arXiv:1305.0667.
34.
S.
Gerstenkorn
,
P.
Luc
,
J.
Verges
,
D. W.
Englekemeir
,
J. E.
Gindler
, and
F. S.
Tomkins
,
J. Phys. France
35
,
483
(
1974
).
35.
D.
DeMille
,
F.
Bay
,
S.
Bickman
,
D.
Kawall
,
L.
Hunter
,
D.
Krause
 Jr.
,
S.
Maxwell
, and
K.
Ulmer
,
AIP Conf. Proc.
596
,
72
(
2001
).
36.
G.
Edvinsson
and
A.
Lagerqvist
,
Phys. Scr.
41
,
316
(
1990
).
37.
ACME Collaboration [http://www.electronedm.org], private communication (
2013
).
38.
N. S.
Mosyagin
, “
Generalized relativistic effective core potentials for actinides
” (unpublished), see http://qchem.pnpi.spb.ru/files/Th30.inp.
39.
R. J.
Bartlett
,
Modern Electronic Structure Theory
,
Advanced Series in Physical Chemistry
Vol.
2
, edited by
D. R.
Yarkony
(
World Scientific
,
Singapore
,
1995
), pp.
1047
1131
, Part II.