This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Determination of hyperfine splittings and Landé gJg_{J} factors of 5sS1/225s~{}^{2}S_{1/2} and 5pP1/2,3/225p~{}^{2}P_{1/2,3/2} states of 111,113Cd+ for a microwave frequency standard

J. Z. Han These authors contributed equally to this work. State Key Laboratory of Precision Measurement Technology and Instruments, Key Laboratory of Photon Measurement and Control Technology of Ministry of Education, Department of Precision Instrument, Tsinghua University, Beijing 100084, China    R. Si These authors contributed equally to this work. Shanghai EBIT Lab, Key Laboratory of Nuclear Physics and Ion-beam Application, Institute of Modern Physics, Department of Nuclear Science and Technology, Fudan University, Shanghai 200433, China    H. R. Qin Department of Physics, Tsinghua University, Beijing 100084, China    N. C. Xin State Key Laboratory of Precision Measurement Technology and Instruments, Key Laboratory of Photon Measurement and Control Technology of Ministry of Education, Department of Precision Instrument, Tsinghua University, Beijing 100084, China    Y. T. Chen State Key Laboratory of Precision Measurement Technology and Instruments, Key Laboratory of Photon Measurement and Control Technology of Ministry of Education, Department of Precision Instrument, Tsinghua University, Beijing 100084, China    S. N. Miao State Key Laboratory of Precision Measurement Technology and Instruments, Key Laboratory of Photon Measurement and Control Technology of Ministry of Education, Department of Precision Instrument, Tsinghua University, Beijing 100084, China    C. Y. Chen [email protected] Shanghai EBIT Lab, Key Laboratory of Nuclear Physics and Ion-beam Application, Institute of Modern Physics, Department of Nuclear Science and Technology, Fudan University, Shanghai 200433, China    J. W. Zhang [email protected] State Key Laboratory of Precision Measurement Technology and Instruments, Key Laboratory of Photon Measurement and Control Technology of Ministry of Education, Department of Precision Instrument, Tsinghua University, Beijing 100084, China    L. J. Wang [email protected] State Key Laboratory of Precision Measurement Technology and Instruments, Key Laboratory of Photon Measurement and Control Technology of Ministry of Education, Department of Precision Instrument, Tsinghua University, Beijing 100084, China Department of Physics, Tsinghua University, Beijing 100084, China
Abstract

Regarding trapped-ion microwave-frequency standards, we report on the determination of hyperfine splittings and Landé gJg_{J} factors of 111,113Cd+. The hyperfine splittings of the 5pP3/225p~{}^{2}P_{3/2} state of 111,113Cd+ ions were measured using laser-induced fluorescence spectroscopy. The Cd+ ions were confined in a linear Paul trap and sympathetically cooled by Ca+ ions. Furthermore, the hyperfine splittings and Landé gJg_{J} factors of the 5sS1/225s~{}^{2}S_{1/2} and 5pP1/2,3/225p~{}^{2}P_{1/2,3/2} levels of 111,113Cd+ were calculated with greater accuracy using the multiconfiguration Dirac–Hartree–Fock scheme. The measured hyperfine splittings and the Dirac–Hartree–Fock calculation values were cross-checked, thereby further guaranteeing the reliability of our results. The results provided in this work can improve the signal-to-noise ratio of the clock transition and the accuracy of the second-order Zeeman shift correction, and subsequently the stability and accuracy of the microwave frequency standard based on trapped Cd+ ions.

I Introduction

With their improvements in accuracy over time, atomic clocks have played an important role in practical applications Hinkley et al. (2013); Burt et al. (2021) and testing the fundamental physics Dzuba et al. (2016); Wcisło et al. (2016); Safronova et al. (2018). Indeed, the microwave-frequency atomic clock plays a vital role in satellite navigation Bandi et al. (2011), deep space exploration Prestage and Weaver (2007); Burt et al. (2016), and timekeeping Diddams et al. (2004). Among the many clock proposals, trapped-ion microwave-frequency clocks have attracted wide attention from researchers because the ions are well isolated from the external environment in an ion trap. The setup is conducive to improvements in the transportability of atomic clocksSchwindt et al. (2016); Mulholland et al. (2019a, b); Hoang et al. (2021). Such clocks are also considered the next generation of practical microwave clocks Schmittberger and Scherer (2020).

Cadmium ions (Cd+) benefit from a simple and distinct electronic structure, which is easily controlled, manipulated, and measurable with high precision. The microwave-frequency standard based on laser-cooled 113Cd+ has achieved an accuracy of 1.8×10141.8\times 10^{-14} and a short-term stability of 4.2×1013/τ4.2\times 10^{-13}/\sqrt{\tau} Miao et al. (2021). The high performance and potential for miniaturization make this frequency standard suitable in establishing a ground-based transportable frequency reference for navigation systems and for comparing atomic clocks between remote sites Zhang et al. (2012); Wang S. G (2013); Miao et al. (2015, 2021). Moreover, it has been proposed as a means to achieve an ultra-high level of accuracy down to 101510^{-15} Han et al. (2021), highlighting the importance of accurately evaluating systematic frequency shifts.

Optical pumping is a fundamental process in operating a trapped-ion microwave frequency standard. The optical pumping efficiency determines directly the signal-to-noise ratio of the “clock signal,” which affects the short-term stability and measurement accuracy of the ground-state hyperfine splitting (HFS) for such frequency standards. Realizing optical pumping for the 113Cd+ microwave-frequency standard requires a blueshift in the laser frequency of the Doppler-cooling transition 5sS1/22F=1,mF=15s~{}^{2}S_{1/2}~{}F=1,~{}m_{F}=15pP3/22F=2,mF=25p~{}^{2}P_{3/2}~{}F=2,~{}m_{F}=2 to reach the 5pP3/22F=15p~{}^{2}P_{3/2}~{}F=1 hyperfine level. However, there are no precise measurements available of the HFSs for other excited states Li et al. (2018). A preliminary measurement of the HFS for the 5pP3/225p~{}^{2}P_{3/2} level of the 113Cd+ ion is approximately 800 MHz Tanaka et al. (1996). Therefore, to improve the optical pumping efficiency and hence the performance of the 113Cd+ microwave frequency standard, the HFSs of the 5pP3/225p~{}^{2}P_{3/2} level of the 113Cd+ ion need to be determined with greater accuracy. From the perspective of atomic structure calculations, the high-precision measurements of the HFSs for the 5pP3/225p~{}^{2}P_{3/2} level of 111,113Cd+ can also be used for testing and developing calculation models of the atomic structure.

In a trapped-ion microwave frequency standard, an external magnetic field is applied to provide the quantization axis to break the degeneracy of the ground-state magnetic level. Among all the systematic frequency shifts of a frequency standard, one dominant shift is the second-order Zeeman shift (SOZS) induced by the external magnetic field Berkeland et al. (1998); Phoonthong et al. (2014); Miao et al. (2021). The precise estimation of this SOZS and the calibration of the external magnetic field require accurate knowledge of the ground-state Landé gjg_{j} factor Han et al. (2019). The external magnetic field in our latest laser-cooled microwave-frequency standard based on trapped 113Cd+ ions is approximately 80008000 nT Miao et al. (2021). However, only two theoretical studies have provided a value of the ground-state Landé gJg_{J} factor of Cd+, one giving 2.00286(53)2.00286(53), calculated using the relativistic-coupled-cluster (RCC) theory Han et al. (2019), and the other giving 2.002291(4)2.002291(4), calculated by the Λ\Lambda-approach RCC (Λ\Lambda-RCC) theory Yu et al. (2020a). The two Landé gJg_{J} factors have a difference of 0.00060.0006 that generates a relative frequency shift of 6.6×10146.6\times 10^{-14}. This large systematic shift obviously falls short inaccuracy of our latest 113Cd+ microwave-frequency standard (1.8×10141.8\times 10^{-14}) Miao et al. (2021). Therefore, re-determining the ground-state gJg_{J} factor of 113Cd+ is imperative if further improvements in accuracy for this microwave-frequency standard are to be attained.

In this work, the HFSs of the 5pP3/225p~{}^{2}P_{3/2} level of the 113Cd+ ion is measured using the laser-induced fluorescence (LIF) technique. To maintain a low-temperature environment, the 113Cd+ ions are sympathetic-cooled by laser-cooled 40Ca+ ions, a technique that improves the accuracy of measurements. Furthermore, the HFSs and Landé gJg_{J} factors of both the 5sS1/225s~{}^{2}S_{1/2} and 5pP1/2,3/225p~{}^{2}P_{1/2,3/2} levels were calculated using the multiconfiguration Dirac–Hartree–Fock (MCDHF) method. Electron correlation effects are carefully investigated and taken into account. Off-diagonal terms are also included to improve the calculation accuracy of the HFSs for the 5pP1/2,3/225p~{}^{2}P_{1/2,3/2} level in Cd+. Cross-checking the measured and calculated HFS results ensures the reliability and accuracy of the results provided in this work. Our results are of great importance for further improving the performance of the Cd+ microwave-frequency standard.

II Experiment

To obtain the HFSs of the 5pP3/225p~{}^{2}P_{3/2} level for 111,113Cd+, we first measure the frequency shifts from the 5sS1/22F=15s~{}^{2}S_{1/2}~{}F=15pP3/22F=25p~{}^{2}P_{3/2}~{}F=2 transition of 111,113Cd+ to the 5sS1/225s~{}^{2}S_{1/2}5pP3/225p~{}^{2}P_{3/2} transition of 114Cd+. Briefly, for the experimental setup (see Ref. Han et al. (2021) for details), we prepared crystals of two ion species consisting of approximately 10510^{5} Ca+ and Cd+ ions in a linear Paul trap. The Ca+ and Cd+ ions are produced by selected-photoionization using laser beams of wavelength 423-nm (Ca 4s2S014s^{2}~{}^{1}S_{0}4s4pP114s4p~{}^{1}P_{1}) / 374-nm (Ca 4s4pP114s4p~{}^{1}P_{1}– Continuum), and 228-nm (Cd 5s2S015s^{2}~{}^{1}S_{0}5s5pP115s5p~{}^{1}P_{1}). The Ca+ are used as coolant ions that are Doppler-cooled using lasers beams of wavelength 397-nm (Ca+ 4sS1/224s~{}^{2}S_{1/2}4pP1/224p~{}^{2}P_{1/2}) and 866-nm (Ca+ 3dD3/223d~{}^{2}D_{3/2}4pP1/224p~{}^{2}P_{1/2}). The Cd+ ions are sympathetically-cooled to less than 0.5 K through Coulomb interactions with the Ca+ ions. The frequency shifts of the 5sS1/22(F=1)5s~{}^{2}S_{1/2}~{}(F=1)5pP3/22(F=2)5p~{}^{2}P_{3/2}~{}(F=2) transition of 111,113Cd+ and the 5sS1/225s~{}^{2}S_{1/2}5pP3/225p~{}^{2}P_{3/2} transitions of 114Cd+ were measured using scanning frequencies in a weak 214.5-nm probe laser beam. The 5sS1/22(F=1,mF=1)5s~{}^{2}S_{1/2}~{}(F=1,m_{F}=1)5pP3/22(F=2,mF=2)5p~{}^{2}P_{3/2}~{}(F=2,m_{F}=2) transition is a cycling transition that was used to cool and detect the 111,113Cd+ ions. Although the circularly polarized cooling laser beam excites a cycling transition, ions may, as a result of the polarization impurity, still, leak to the 5sS1/22F=05s~{}^{2}S_{1/2}~{}F=0 state via 5pP3/22F=15p~{}^{2}P_{3/2}~{}F=1 state. To increase detection efficiency, 20-dBm microwave radiation resonant with the ground-state hyperfine transition (15.2-GHz for 113Cd+ and 14.5-GHz for 111Cd+) is applied during LIF detection. The frequency of each laser beam is measured using a high-precision wavemeter (HighFinesse WS8-2).

Refer to caption
Figure 1: Schematic of the experiment setup for the 111,113Cd+ HFS measurements. Sympathetic cooling technology is used to maintain the Cd+ cloud at low temperatures and improve measurement accuracy. The energy level scheme (not to scale) for 111,113Cd+ is also presented.

In obtaining the measured LIF spectrum (Fig. 2), the beam intensity is maintained below 5 μ\muW/mm2 (saturation parameter 0.0006) to reduce the cooling and heating effects of the probe beam. The fitted curve is the Voigt profile Zuo et al. (2019); Han et al. (2021), expressed as

F\displaystyle F =\displaystyle= F0+(FLFG)(ν),\displaystyle F_{0}+(F_{L}\ast F_{G})(\nu),
FL(ν)\displaystyle F_{L}(\nu) =\displaystyle= 2AπνL4(ννc)2+νL2,\displaystyle\frac{2A}{\pi}\frac{\nu_{L}}{4(\nu-\nu_{c})^{2}+\nu_{L}^{2}},
FG(ν)\displaystyle F_{G}(\nu) =\displaystyle= 4ln2πe4ln2νG2ν2νG,\displaystyle\sqrt{\frac{4\ln 2}{\pi}}\frac{e^{-\frac{4\ln 2}{\nu^{2}_{G}}\nu^{2}}}{\nu_{G}}, (1)

where F0F_{0} is the offset, ν\nu is the laser beam frequency, νc\nu_{c} is the ion resonance frequency, AA is the area, νL\nu_{L} is the Lorentzian width, νG\nu_{G} is the Gaussian width of Doppler broadening. The line profile is slightly asymmetrically because of the heating and cooling effects of the probe beam, which lead to a slight increase in the uncertainty of the estimated transition frequency.

Refer to caption
Figure 2: Typical LIF spectrum of Cd+, using 113Cd+ as an example. The frequency of the probe laser beam (214 nm wavelength) is scanned around the resonance frequency over a range of 600 MHz. The measured line profiles are fitted with a Voigt function. The inset is an image of the two-species ion cloud of Cd+ and Ca+ captured by an EMCCD camera; aberration has blurred the image.

Measurements present three sources of uncertainty:

  • i)

    Statistical uncertainties. For the Cd+ 5sS1/225s~{}^{2}S_{1/2}5pP3/225p~{}^{2}P_{3/2} transition, νL\nu_{L} is 60.13 MHz which represent the natural width, the fitted νG\nu_{G} is approximately 30 MHz, and the ion temperature is estimated to be approximately 100 mK. The statistical uncertainty associated with the transition frequencies of 114Cd+ and the 111,113Cd+ is approximately 1 MHz, and thus the statistical uncertainty in their differences is approximately 1.4 MHz;

  • ii)

    Instrument uncertainties. The uncertainty arising from the drift in the wavemeter is less than 0.5 MHz in a laboratory environment Liu et al. (2018);

  • iii)

    Systematic uncertainties. Most systematic shifts are common to the 5sS1/225s~{}^{2}S_{1/2}5pP3/225p~{}^{2}P_{3/2} transitions of both 114Cd+ and 111/113Cd+ and thus cancel each other out. Because the 5sS1/22(F=1)5s~{}^{2}S_{1/2}~{}(F=1)5pP3/22(F=2)5p~{}^{2}P_{3/2}~{}(F=2) transition of Cd+ is sensitive to magnetic fields, the Zeeman shift becomes the dominant contributor to systematic uncertainties.

In a weak field (μBB0.14\mu_{B}B\sim 0.14 MHz \ll hyperfine constant AA), the Zeeman shift for a specific energy level is expressed as

EZeeman=gFMFμBB,E_{Zeeman}=g_{F}M_{F}\mu_{B}\cdot B, (2)

where gFg_{F} is given by

gF=F(F+1)+J(J+1)I(I+1)2F(F+1)gJ,g_{F}=\frac{F(F+1)+J(J+1)-I(I+1)}{2F(F+1)}g_{J}, (3)

where J=L+SJ=L+S the total electron angular momentum (SS and LL the spin and orbital angular momenta), and F=I+JF=I+J the total angular momentum with II denoting the nuclear spin. In typical conditions of our experiment, B8000B\sim 8000 nT Miao et al. (2015, 2021). By introducing values gJ=2.002257g_{J}=2.002257 and 1.3340561.334056 for the levels 5sS1/225s~{}^{2}S_{1/2} and 5pP3/225p~{}^{2}P_{3/2} calculated in this work (see text below), the Zeeman shift for the 5sS1/22(F=1)5s~{}^{2}S_{1/2}~{}(F=1)5pP3/22(F=2)5p~{}^{2}P_{3/2}~{}(F=2) transition of Cd+ is estimated to be 0.11 MHz. Therefore, the total systematic shifts for the 5sS1/22(F=1)5s~{}^{2}S_{1/2}~{}(F=1)5pP3/22(F=2)5p~{}^{2}P_{3/2}~{}(F=2) transitions of Cd+ are estimated to be below 0.5 MHz.

The final frequencies for the 5sS1/22(F=1)5s~{}^{2}S_{1/2}~{}(F=1)5pP3/22(F=2)5p~{}^{2}P_{3/2}~{}(F=2) transitions of 111,113Cd+ and that for 5sS1/225s~{}^{2}S_{1/2}5pP3/225p~{}^{2}P_{3/2} of 114Cd+ are determined to be 4649.0(1.6) MHz and 4041.8(1.6) MHz, respectively.

In LS-coupling, the energy shifts after the hyperfine interaction are expressed as Foot et al. (2005)

EHFS=AIJ=A2[F(F+1)I(I+1)J(J+1)].E_{HFS}=A\langle I\cdot J\rangle=\frac{A}{2}[F(F+1)-I(I+1)-J(J+1)]. (4)

Therefore, for 111,113Cd+, we have

νs,F=1p,F=2111/113=νsp111/113+14νHFS,s38νHFS,p,\nu_{s,F=1\rightarrow p,F=2}^{111/113}=\nu_{s\rightarrow p}^{111/113}+\frac{1}{4}\nu_{HFS,s}-\frac{3}{8}\nu_{HFS,p}, (5)

where νs,F=1p,F=2111/113\nu_{s,F=1\rightarrow p,F=2}^{111/113} is the transition frequency of 5sS1/22(F=1)5s~{}^{2}S_{1/2}~{}(F=1)5pP3/22(F=2)5p~{}^{2}P_{3/2}~{}(F=2) of 111,113Cd+; νsp111/113\nu_{s\rightarrow p}^{111/113} is the transition frequency of 5sS1/225s~{}^{2}S_{1/2}5pP3/225p~{}^{2}P_{3/2}; νHFS,s\nu_{HFS,s} is the HFS of 5sS1/225s~{}^{2}S_{1/2}; and νHFS,p\nu_{HFS,p} is the HFS of 5pP3/225p~{}^{2}P_{3/2}. In reference to 114Cd+, through a linear transformation, Eq. (5) may be expressed as

Δνs,F=1p,F=2111/113,114\displaystyle\Delta\nu_{s,F=1\rightarrow p,F=2}^{111/113,114} =\displaystyle= Δνsp111/113,114+14νHFS,s111/113\displaystyle\Delta\nu_{s\rightarrow p}^{111/113,114}+\frac{1}{4}\nu_{HFS,s}^{111/113}
\displaystyle- 38νHFS,p111/113,\displaystyle\frac{3}{8}\nu_{HFS,p}^{111/113},

where Δνs,F=1p,F=2111/113,114=νs,F=1p,F=2111/113νsp114\Delta\nu_{s,F=1\rightarrow p,F=2}^{111/113,114}=\nu_{s,F=1\rightarrow p,F=2}^{111/113}-\nu_{s\rightarrow p}^{114} and Δνsp111/113,114=νsp111/113νsp114\Delta\nu_{s\rightarrow p}^{111/113,114}=\nu_{s\rightarrow p}^{111/113}-\nu_{s\rightarrow p}^{114}. With our measurements, Δνs,F=1p,F=2111/113,114\Delta\nu_{s,F=1\rightarrow p,F=2}^{111/113,114} are respectively 4649.0(1.6) MHz and 4041.8(1.6) MHz, whereas Δνsp111/113,114\Delta\nu_{s\rightarrow p}^{111/113,114} are 1314.3(22)[023] MHz and 555.2(23)[008] MHz Hammen et al. (2018). From our previous measurements obtained through double-resonance microwave laser spectroscopy, the νHFS,s111/113\nu_{HFS,s}^{111/113} were accurately measured to be 14530507349.9(1.1) Hz Zhang et al. (2012) and 15199862855.02799(27) Hz Miao et al. (2021), from which we derived νHFS,p111/113\nu_{HFS,p}^{111/113} to be 794.6(3.6) MHz and 835.5(2.9) MHz.

III Theory

III.1 Multiconfiguration Dirac–Hartree–Fock approach

The MCDHF method C. Froese Fischer et al. (2016), as implemented in the Grasp computer package Jönsson et al. (2013); C. Froese Fischer et al. (2019), is employed to obtain wave functions referred to as atomic state functions. Specifically, they are approximate eigenfunctions of the Dirac Hamiltonian describing a Coulombic system given by

HDC=i=1N(c𝜶𝒊𝒑𝒊+(βi1)c2+Vi)+i<jN1rij,H_{\rm DC}=\sum_{i=1}^{N}(c~{}\bm{\alpha_{i}}\cdot\bm{p_{i}}+(\beta_{i}-1)c^{2}+V_{i})+\sum_{i<j}^{N}\frac{1}{r_{ij}}, (7)

where ViV_{i} denotes the monopole part of the electron–nucleus interaction for a finite nucleus and rijr_{ij} the distance between electrons ii and jj; 𝜶i\bm{\alpha}_{i} and βi\beta_{i} are the Dirac matrices for electron ii.

Electron correlations are included by expanding |ΓJ|\Gamma J\rangle, an atomic state function, over a linear combination of configuration state functions (CSFs) |γJ|\gamma J\rangle,

|ΓJ=γcγ|γJ,|\Gamma J\rangle=\sum_{\gamma}c_{\gamma}|\gamma J\rangle, (8)

where γ\gamma represents the parity and all the coupling tree quantum numbers needed to define the CSF uniquely. The CSFs are four-component spin-angular coupled, antisymmetric products of Dirac orbitals of the form

ϕ(𝐫)=1r(Pnκ(r)χκm(θ,ϕ)iQnκ(r)χκm(θ,ϕ)).\phi({\bf r})=\frac{1}{r}\left(\begin{array}[]{c}P_{n\kappa}(r)\chi_{\kappa m}(\theta,\phi)\\ iQ_{n\kappa}(r)\chi_{-\kappa m}(\theta,\phi)\end{array}\right). (9)

The radial parts of the one-electron orbitals and the expansion coefficients cγc_{\gamma} of the CSFs are obtained by the self-consistent relativistic field (RSCF) procedure. In the following calculations of the relativistic configuration interaction (RCI), the Dirac orbitals are kept fixed, and only the expansion coefficients of the CSFs are determined for selected eigenvalues and eigenvectors of the complete interaction matrix. This procedure includes the Breit interaction and the leading quantum electrodynamic (QED) effects (vacuum polarization and self-energy).

The restricted active-set method is used in obtaining the CSF expansions by allowing single and double (SD) substitutions from a selected set of reference configurations to an active set (AS) of given orbitals. The configuration space is increased step by step by increasing the number of layers, specifically, a set of virtual orbitals. These virtual orbitals are optimized in the RSCF procedure while all orbitals of the inner layers are fixed.

The interaction between the electrons and the electromagnetic multipole moments of the nucleus splits each fine structure level into multiple hyperfine levels. The interaction couples the nuclear spin 𝑰\bm{I} with the total electronic angular momentum 𝑱\bm{J} to obtain total angular momentum 𝑭=𝑰+𝑱\bm{F=I+J}.

The hyperfine contribution to the Hamiltonian is represented by a multipole expansion

HHFS=k1𝐓(𝐤)𝐌(𝐤),H_{\rm HFS}=\sum_{k\geq 1}\bf{T^{(k)}}\cdot\bf{M^{(k)}}, (10)

where 𝐓(𝐤)\bf{T^{(k)}} and 𝐌(𝐤)\bf{M^{(k)}} are spherical tensor operators of rank kk in the electronic and nuclear spaces, respectively Lindgren and Rosén (1975). The k=1k=1 term represents the magnetic dipole interaction, and the k=2k=2 term the electric quadrupole interaction. Higher-order terms are minor and can often be neglected.

For the scheme considered in this work (111,113Cd+ with I=1/2I=1/2), only the magnetic dipole interaction is non-zero. To first-order, the fine-structure level γJ\gamma J is then split according to

ΓIJFMF|𝑻(𝟏)𝑴(𝟏)|ΓIJFMF\displaystyle\langle\Gamma IJFM_{F}|\bm{T^{(1)}\cdot M^{(1)}}|\Gamma IJFM_{F}\rangle (11)
=\displaystyle= (1)I+J+F{IJFJI1}ΓJT(1)ΓJΓIM(1)ΓI,\displaystyle(-1)^{I+J+F}\left\{\begin{array}[]{ccc}I&J&F\\ J&I&1\end{array}\right\}\langle\Gamma J||T^{(1)}||\Gamma J\rangle\langle\Gamma I||M^{(1)}||\Gamma I\rangle,

where the coefficient in curly brackets in the 6j symbol of the rotation group. The reduced matrix elements of the nuclear tensor operators are related to the conventional nuclear magnetic dipole moment,

ΓIM(1)ΓI=μI(2I+1)(I+1)I.\langle\Gamma I||M^{(1)}||\Gamma I\rangle=\mu_{I}\sqrt{\frac{(2I+1)(I+1)}{I}}. (12)

The hyperfine interaction energy contribution to a specified hyperfine level is then given by

ΓIJFMF|𝑻(𝟏)𝑴(𝟏)|ΓIJFMF=12AJC,\langle\Gamma IJFM_{F}|\bm{T^{(1)}\cdot M^{(1)}}|\Gamma IJFM_{F}\rangle=\frac{1}{2}A_{J}C, (13)

where AJA_{J} is the magnetic dipole hyperfine constants

AJ=μII1J(J+1)γJT(1)γJ,A_{J}=\frac{\mu_{I}}{I}\frac{1}{\sqrt{J(J+1)}}\langle\gamma J||T^{(1)}||\gamma J\rangle, (14)

and C=F(F+1)J(J+1)I(I+1)C=F(F+1)-J(J+1)-I(I+1).

However, this method discards the off-diagonal hyperfine interaction. This is not sufficient for P1/2,3/22{}^{2}P_{1/2,3/2} because the hyperfine interaction between the two F=1F=1 hyperfine levels is non-negligible owing to their minor splitting. To account for the off-diagonal hyperfine effects, we consider the second-order hyperfine interaction between P1/2,3/22{}^{2}P_{1/2,3/2}. The contribution associated with sublevel labelled γIJFMF\gamma IJFM_{F} may be expressed as

|γIJFMF|Hhfs|γIJFMF|2EJEJ.\frac{|\langle\gamma IJFM_{F}|H_{hfs}|\gamma^{\prime}IJ^{\prime}FM_{F}\rangle|^{2}}{E_{J}-E_{J^{\prime}}}. (15)

In the relativistic theory, choosing the direction of the magnetic field as the zz-direction of the interaction and neglecting all diamagnetic contributions, the interaction between the magnetic moment of the atom and an external field may be written as

HM=(N0(1)+ΔN01)B,H_{M}=(N_{0}^{(1)}+\Delta N_{0}^{1})B, (16)

where the last term is the Schwinger QED correction Cheng and Childs (1985). To first order, the fine-structure splitting in the energy level is

ΓJMJ|N0(1)+ΔN01|ΓJMJB\displaystyle\langle\Gamma JM_{J}|N_{0}^{(1)}+\Delta N_{0}^{1}|\Gamma JM_{J}\rangle B (17)
=\displaystyle= MJJ(J+1)ΓJ𝑵(𝟏)+𝚫𝑵(𝟏)ΓJB.\displaystyle\frac{M_{J}}{\sqrt{J(J+1)}}\langle\Gamma J||\bm{N^{(1)}+\Delta N^{(1)}}||\Gamma J\rangle B.

Factoring out the dependence on the MJM_{J} quantum number, the energies are expressed in terms of the Landé gJg_{J} factor

gJ=2ΓJ𝑵(𝟏)+𝚫𝑵(𝟏)ΓJJ(J+1).g_{J}=2\frac{\langle\Gamma J||\bm{N^{(1)}+\Delta N^{(1)}}||\Gamma J\rangle}{\sqrt{J(J+1)}}. (18)

The energy splittings are then given by

gJMJB2.g_{J}M_{J}\frac{B}{2}. (19)
Table 1: Calculated hyperfine splitting (HFS) (in MHz) and Landé gJg_{J} factors for 4d105s2S1/24d^{10}5s\ ^{2}S_{1/2} of 111,113Cd+ obtained through the MCDHF approach.
MCDHF2+RCI MCDHF3+RCI
111HFS 113HFS gJg_{J} 111HFS 113HFS gJg_{J}
5 11841 12386 2.002243 12925 13521 2.002242
6 13421 14040 2.002245 14096 14746 2.002247
7 13671 14301 2.002254 14414 15079 2.002250
8 13879 14518 2.002260 14476 15143 2.002256
9 13884 14524 2.002262 14515 15184 2.002257
10 13925 14567 2.002263 14541 15211 2.002257
11 13919 14561 2.002263 14532 15202 2.002257
12 13921 14562 2.002262 14537 15207 2.002257
Final 13921(7) 14563(7) 2.002262(2) 14536(9) 15206(9) 2.002257(1)

III.2 Computation Strategy

Initially, the MCDHF calculation was performed using the extended optimal-level scheme for the states of the 4d105s4d^{10}5s and 4d105p4d^{10}5p configurations, and these occupied orbitals were determined simultaneously and maintained throughout subsequent calculations. Because the 4f4f subshell in both the 4d105s4d^{10}5s and 4d105p4d^{10}5p configurations is vacant, imaging the strong pair correlations between 4d24d^{2} and 4f24f^{2}, and between 4d24d^{2} and 5d25d^{2} is easy. The strong core-core (CC) correlation of the 4d4d electrons and the single valence 5s/5p5s/5p electron do not allow us to include only the valence correlation; hence, our MCDHF calculation starts from the CC4d mode, in which the 5s/5p5s/5p and 4d4d electrons can be SD-excited to the n8,l6n\leq 8,l\leq 6 level (AS8, where ‘8’ labels the maximum principal quantum number in the corresponding AS). To investigate the correlation effects of the inner core electrons, we performed a few RCI calculations using the MCDHF wavefunctions derived from the CC4d calculation. This calculation method is labelled MCDHF1+RCI in this paper. From the plots of the AJA_{J} and gJg_{J} factors for the S1/22{}^{2}S_{1/2} and P1/2,3/22{}^{2}P_{1/2,3/2} levels of 113Cd+ from this calculation method (Fig. 3), we see that they are sensitive to the core correlations, and even the core–valence (CV) correlation of the 1s1s orbital. To describe the wavefunctions better, we performed a second calculation that includes the strong CV and CC correlations in the RSCF procedure rather than only including the CV/CC correlations in the RCI procedure.

Refer to caption
Figure 3: Hyperfine structure constants AJA_{J} and Landé gJg_{J} factors of the 4d105s2S1/24d^{10}5s\ ^{2}S_{1/2} and 4d105p2P1/2,3/24d^{10}5p\ ^{2}P_{1/2,3/2} levels for 113Cd+, from the various correlation models used in our MCDHF1+RCI calculation.
Table 2: Hyperfine splitting (HFS) (in MHz) and Landé gJg_{J} factors for the 4d105p2P1/2,3/24d^{10}5p\ ^{2}P_{1/2,3/2} level of 111,113Cd+. HFSn.o indicate that no off-diagonal contributions were included, whereas HFSw.o refers to HFS calculations with off-diagonal contributions.
MCDHF2+RCI MCDHF3+RCI
111HFSn.o 113HFSn.o gJg_{J} 111HFSn.o 113HFSn.o 111HFSw.o 113HFSw.o gJg_{J}
4d105p2P1/24d^{10}5p\ ^{2}P_{1/2}
5 1816 1900 0.665833 2061 2156 2067 2150 0.665829
6 2099 2196 0.665825 2267 2371 2270 2368 0.665821
7 2054 2148 0.665825 2255 2359 2259 2355 0.665820
8 2142 2241 0.665820 2310 2416 2314 2413 0.665814
9 2123 2220 0.665822 2300 2406 2304 2403 0.665818
10 2146 2245 0.665821 2324 2432 2328 2428 0.665818
11 2144 2243 0.665820 2316 2422 2319 2419 0.665816
12 2145 2244 0.665820 2319 2425 2322 2422 0.665816
Final 2140(24) 2239(25) 0.665821(2) 2314(25) 2420(26) 2317(25) 2417(26) 0.665817(4)
4d105p2P3/24d^{10}5p\ ^{2}P_{3/2}
5 546 571 1.334062 644 674 650 680 1.334056
6 699 732 1.334057 787 824 791 827 1.334052
7 699 731 1.334060 768 803 772 807 1.334056
8 727 761 1.334060 794 830 797 834 1.334056
9 722 755 1.334062 786 822 789 826 1.334058
10 729 763 1.334060 789 828 793 832 1.334055
11 728 762 1.334059 789 824 792 828 1.334057
12 729 762 1.334060 789 825 793 829 1.334056
Final 727(8) 760(8) 1.334061(2) 789(8) 823(9) 792(8) 830(9) 1.334056(3)

In our second calculation, based on the above investigations, we included CV and CC correlations for the 4d4d, 4p4p, 4s4s, 3d3d, and 3p3p electrons, as well as the CV correlation down to the 1s1s subshell by allowing restricted SD excitations to n12,l6n\leq 12,l\leq 6 (AS12), in the RSCF calculation. RCI calculations were also performed to include the Breit and QED effects. Note that this calculation also started from the MCDHF calculation for the 4d105s4d^{10}5s and 4d105p4d^{10}5p configurations and hence is labelled MCDHF2+RCI.

In analyzing the wavefunction compositions from the MCDHF2+RCI calculation, we noticed strong correlations between 4d105s4d^{10}5s, 4d84f25s4d^{8}4f^{2}5s, and 4d94f5p4d^{9}4f5p configurations, and between 4d105p4d^{10}5p, 4d84f25p4d^{8}4f^{2}5p, and 4d85p5d24d^{8}5p5d^{2} configurations. Therefore, instead of starting from the DHF calculation where only 4d105s4d^{10}5s and 4d105p4d^{10}5p were included in the CSF list, we allowed the 4d4d and 5s/5p5s/5p electrons to be SD-excited to {5s,5p,5d,4f5s,5p,5d,4f} to generate the CSF list as a starting point of our third calculation approach. In this way, the spectroscopic orbitals together with the 5d5d and 4f4f orbitals are optimized together, and the correlation effect between the essential CSFs is included in the beginning. The CV and CC correlation effects are included by systematically increasing the virtual excitations to AS12; this calculation method is labelled MCDHF3+RCI.

For 4d105s2S1/24d^{10}5s\ ^{2}S_{1/2}, because there are no other levels with which to have strong hyperfine interactions, we only included the diagonal contributions to its HFS. The calculated HFSs and gJg_{J} factors with an increasing AS size from MCDHF2+RCI and MCDHF3+RCI calculations are listed in Table 1. We find some fluctuations in our calculated HFSs with increasing AS size, but the values from the last few ASs generally tend to some specific value. We, therefore, took the average of the last three values (AS10, AS11, and AS12) as our final calculated result, with the maximum difference between them taken as the calculation error. Although the final splitting for 113Cd+ from the MCDHF2+RCI calculation (i.e., 14563(7) MHz) is much smaller than the experimental measurement (i.e., 15199 MHz), the MCDHF3+RCI calculation, (15206(9) MHz) shows a significant improvement with the experimental value being within the estimated uncertainty of the latter calculation. Following a similar method, the gJg_{J} factors of 2S1/22S_{1/2} from MCDHF2+RCI and MCDHF3+RCI calculations were 2.002262(2) and 2.002257(1), respectively. The HFSs and gJg_{J} factors for 4d105s2P1/2,3/24d^{10}5s\ ^{2}P_{1/2,3/2} with an increasing AS are listed in Table 2. Following the same method as used in determining our final calculation results and their uncertainties, the HFSs for the P3/22{}^{2}P_{3/2} level of 111Cd+/113Cd+ from MCDHF2+RCI and MCDHF3+RCI calculations when not including the off-diagonal contributions are 727(8)/760(8) MHz and 789(8)/823(9) MHz, respectively. With off-diagonal contributions included, the MCDHF3+RCI results increase to 792(8)/830(9) MHz.

IV Results and discussions

The measured HFSs for the 5pP3/225p~{}^{2}P_{3/2} level and the calculated HFSs and Landé gJg_{J} factors for the 5sS1/225s~{}^{2}S_{1/2} and 5pP1/3,3/225p~{}^{2}P_{1/3,3/2} levels in this work are listed in Table 3; other experimental and calculated results are also listed for comparison. For the HFSs, our group’s previous high-accuracy measurements for the 111,113Cd+ ground state provided an excellent benchmark for the atomic structure calculation of Cd+. The present HFSs for the 5sS1/225s~{}^{2}S_{1/2} state calculated using the MCDHF method show stronger agreement with our previous experimental results than those of previous theoretical results Dixit et al. (2008); Li et al. (2018). The present measured HFSs for the 5pP3/225p~{}^{2}P_{3/2} level is also in agreement with the present theoretical results. The cross-checking between experiment and theory ensures the reliability of the Cd+ 5pP3/225p~{}^{2}P_{3/2} HFSs determined in this work. We recommend the adoption of 794.6(3.6) and 835.5(2.9) as the blue-shifted frequencies for optical pumping in the microwave-frequency standard based on 111/113Cd+.

Regarding the Landé gJg_{J} factors, there are no experimental results for Cd+. Accurate calculations of Landé gJg_{J} factors has proven complicated even for alkali atoms and alkali-like ions because they are sensitive to electron correlations. Those calculated in this work using the MCDHF method show strong deviations from previous RCC results Han et al. (2019). The ground state Landé gJg_{J} factor calculated in this work (2.002257(1)2.002257(1)) agrees with the recent result calculated using the Λ\Lambda-RCC theory (2.002291(4)2.002291(4)) Yu et al. (2020b) to the fourth decimal place, although there is no overlap within their margins of uncertainty. To our knowledge, there also exists a significant difference in results between the Λ\Lambda-RCC calculations with the configuration interaction and the many-body perturbation (CI+MBPT) calculations in Yb+ ground-state Landé gJg_{J} factor Gossel et al. (2013); Yu et al. (2020b). Comparing the results of the same physical quantity from different calculation methods is also of great significance for developing atomic structure calculation models and understanding the role of electronic correlation effects. Therefore, we encourage more experimental and theoretical research on the Landé gJg_{J} factors of Cd+.

For precaution, we recommend the value 2.00226(4) for the Cd+ ground state Landé gJg_{J} factor in the evaluation of the SOZS of the microwave frequency standard of trapped Cd+ ions. The SOZS can be estimated using the Breit–Rabi formula,

ΔνZeem(2)(B)=[gjgI]2μB2B22h2Ahf,\Delta\nu^{(2)}_{Zeem}(B)=-\frac{[g_{j}-g_{I}]^{2}\mu_{B}^{2}B^{2}}{2h^{2}A_{hf}}, (20)

for which B8000B\sim 8000 nT for the Cd+ microwave frequency standard during actual operations. Thus, the fractional frequency shifts incurred when using the value of gJ=2.00226(4)g_{J}=2.00226(4) is 4.4×10154.4\times 10^{-15}. The fractional frequency shifts produced by this gJg_{J} factor for the Cd+ ground-state can meet current accuracy requirements for the best Cd+ microwave frequency standard (1.8×10141.8\times 10^{-14}). However, for further developments of this standard, the ground state gJg_{J} factor of Cd+ also needs to be determined more accurately.

Table 3: Measured HFSs for 5pP3/225p~{}^{2}P_{3/2} and the calculated HFSs and Landé gJg_{J} factors of 5sS1/225s~{}^{2}S_{1/2} and 5pP1/3,3/225p~{}^{2}P_{1/3,3/2} of this work. Results of the Cd+ HFSs and Landé gJg_{J} factors from other works are also listed for comparison.
111HFS 113HFS gJg_{J} Method
5sS1/225s~{}^{2}S_{1/2}
14530.507 15199.863 Exp. Zhang et al. (2012)
14536(9) 15206(9) 2.002257(1) MCDHF (This work)
14478(175) 15146(183) RCC Li et al. (2018)
15280 RCC Dixit et al. (2008)
2.00286(53) RCC Han et al. (2019)
2.002291(4) Λ\Lambda-RCC Yu et al. (2020a)
5pP3/225p~{}^{2}P_{3/2}
794.6(3.6) 835.5(2.9) Exp. (This work)
800 Exp. Tanaka et al. (1996)
792(8) 830(9) 1.334056(3) MCDHF (This work)
794(12) 832(12) RCC Li et al. (2018)
812.04 RCC Dixit et al. (2008)
1.33515(43) RCC Han et al. (2019)
5pP1/225p~{}^{2}P_{1/2}
2317(25) 2417(26) 0.665817(4) MCDHF (This work)
2333(31) 2441(33) RCC Li et al. (2018)
2430 RCC Dixit et al. (2008)
0.66747(83) RCC Han et al. (2019)

V Conclusion

We reported on the determination of HFSs and Landé gJg_{J} factors for the 5sS1/225s~{}^{2}S_{1/2} and 5pP1/2,3/225p~{}^{2}P_{1/2,3/2} levels of 111,113Cd+. The HFSs of the 5pP3/225p~{}^{2}P_{3/2} level was measured using the laser-induced-fluorescence technique. The Cd+ ions were co-trapped with Ca+ ions in the same linear ion trap and sympathetically cooled through the Coulomb interaction with laser-cooled Ca+ ions. Furthermore, the HFSs and Landé gJg_{J} factors for both levels of interest were calculated using the MCDHF calculation. Three computational strategies were followed to account for the electronic correlation effects more comprehensively. The final calculated HFSs were in perfect agreement with the measured HFSs of this work and our previous work, which from cross-checks, demonstrated the reliability of the calculations and the experiments. The HFSs and Landé gJg_{J} factors determined in this work can further improve the efficiency of the optical pumping procedure and the accuracy of the second-order Zeeman correction, and the stability and accuracy of the microwave frequency standard based on trapped Cd+ ions.

Acknowledgements

We thank Z. M. Tang for the helpful discussions. This work is supported by the National Key R&D Program of China (No. 2021YFA1400243), National Natural Science Foundation of China (Nos. 91436210, 12074081, 12104095).

References

  • Hinkley et al. (2013) N. Hinkley, J. A. Sherman, N. B. Phillips, M. Schioppo, N. D. Lemke, K. Beloy, M. Pizzocaro, C. W. Oates,  and A. D. Ludlow, Science 341, 1215 (2013).
  • Burt et al. (2021) E. Burt, J. Prestage, R. Tjoelker, D. Enzer, D. Kuang, D. Murphy, D. Robison, J. Seubert, R. Wang,  and T. Ely, Nature 595, 43 (2021).
  • Dzuba et al. (2016) V. Dzuba, V. Flambaum, M. Safronova, S. Porsev, T. Pruttivarasin, M. Hohensee,  and H. Häffner, Nature Physics 12, 465 (2016).
  • Wcisło et al. (2016) P. Wcisło, P. Morzyński, M. Bober, A. Cygan, D. Lisak, R. Ciuryło,  and M. Zawada, Nature Astronomy 1, 1 (2016).
  • Safronova et al. (2018) M. Safronova, D. Budker, D. DeMille, D. F. J. Kimball, A. Derevianko,  and C. W. Clark, Reviews of Modern Physics 90, 025008 (2018).
  • Bandi et al. (2011) T. Bandi, C. Affolderbach, C. E. Calosso,  and G. Mileti, Electronics letters 47, 698 (2011).
  • Prestage and Weaver (2007) J. D. Prestage and G. L. Weaver, Proceedings of the IEEE 95, 2235 (2007).
  • Burt et al. (2016) E. A. Burt, L. Yi, B. Tucker, R. Hamell,  and R. L. Tjoelker, IEEE transactions on ultrasonics, ferroelectrics, and frequency control 63, 1013 (2016).
  • Diddams et al. (2004) S. A. Diddams, J. C. Bergquist, S. R. Jefferts,  and C. W. Oates, Science 306, 1318 (2004).
  • Schwindt et al. (2016) P. D. Schwindt, Y.-Y. Jau, H. Partner, A. Casias, A. R. Wagner, M. Moorman, R. P. Manginell, J. R. Kellogg,  and J. D. Prestage, Review of Scientific Instruments 87, 053112 (2016).
  • Mulholland et al. (2019a) S. Mulholland, H. Klein, G. Barwood, S. Donnellan, D. Gentle, G. Huang, G. Walsh, P. Baird,  and P. Gill, Applied Physics B 125, 1 (2019a).
  • Mulholland et al. (2019b) S. Mulholland, H. Klein, G. Barwood, S. Donnellan, P. Nisbet-Jones, G. Huang, G. Walsh, P. Baird,  and P. Gill, Review of Scientific Instruments 90, 033105 (2019b).
  • Hoang et al. (2021) T. M. Hoang, S. K. Chung, T. Le, J. D. Prestage, L. Yi, R. L. Tjoelker, S. Park, S.-J. Park, J. G. Eden, C. Holland, et al., Applied Physics Letters 119, 044001 (2021).
  • Schmittberger and Scherer (2020) B. L. Schmittberger and D. R. Scherer, arXiv preprint arXiv:2004.09987  (2020).
  • Miao et al. (2021) S. Miao, J. Zhang, H. Qin, N. Xin, J. Han,  and L. Wang, arXiv preprint arXiv:2110.12353  (2021).
  • Zhang et al. (2012) J. W. Zhang, Z. B. Wang, S. G. Wang, K. Miao, B. Wang,  and L. J. Wang, Phys. Rev. A 86, 022523 (2012).
  • Wang S. G (2013) M. K. W. Z. B. W. L. J. Wang S. G, Zhang J. W, Opt Express 21, 12434 (2013).
  • Miao et al. (2015) K. Miao, J. W. Zhang, X. L. Sun, S. G. Wang, A. M. Zhang, K. Liang,  and L. J. Wang, Opt. Lett. 40, 4249 (2015).
  • Han et al. (2021) J. Han, H. Qin, N. Xin, Y. Yu, V. Dzuba, J. Zhang,  and L. Wang, Applied Physics Letters 118, 101103 (2021).
  • Li et al. (2018) C.-B. Li, Y.-M. Yu,  and B. Sahoo, Physical Review A 97, 022512 (2018).
  • Tanaka et al. (1996) U. Tanaka, H. Imajo, K. Hayasaka, R. Ohmukai, M. Watanabe,  and S. Urabe, Physical Review A 53, 3982 (1996).
  • Berkeland et al. (1998) D. Berkeland, J. Miller, J. C. Bergquist, W. M. Itano,  and D. J. Wineland, Physical Review Letters 80, 2089 (1998).
  • Phoonthong et al. (2014) P. Phoonthong, M. Mizuno, K. Kido,  and N. Shiga, Applied Physics B 117, 673 (2014).
  • Han et al. (2019) J. Han, Y. Yu, B. Sahoo, J. Zhang,  and L. Wang, Physical Review A 100, 042508 (2019).
  • Yu et al. (2020a) Y. Yu, B. Sahoo,  and B. Suo, Physical Review A 102, 062824 (2020a).
  • Zuo et al. (2019) Y. Zuo, J. Han, J. Zhang,  and L. Wang, Applied Physics Letters 115, 061103 (2019).
  • Liu et al. (2018) H. Liu, W. Yuan, F. Cheng, Z. Wang, Z. Xu, K. Deng,  and Z. Lu, Journal of Physics B: Atomic, Molecular and Optical Physics 51, 225002 (2018).
  • Foot et al. (2005) C. J. Foot et al.Atomic physics, Vol. 7 (Oxford University Press, 2005).
  • Hammen et al. (2018) M. Hammen, W. Nörtershäuser, D. Balabanski, M. Bissell, K. Blaum, I. Budinčević, B. Cheal, K. Flanagan, N. Frömmgen, G. Georgiev, et al., Physical review letters 121, 102501 (2018).
  • C. Froese Fischer et al. (2016) C. Froese Fischer, M. Godefroid, T. Brage, P. Jönsson,  and G. Gaigalas, J. Phys. B: At. Mol. Opt. Phys. 49, 182004 (2016).
  • Jönsson et al. (2013) P. Jönsson, G. Gaigalas, J. Bieroń, C. Froese Fischer,  and I. Grant, Comput. Phys. Commun. 184, 2197 (2013).
  • C. Froese Fischer et al. (2019) C. Froese Fischer, G. Gaigalas, P. Jönsson,  and J. Bieroń, Comput. Phys. Commun. 237, 184 (2019).
  • Lindgren and Rosén (1975) I. Lindgren and A. Rosén, in Case Studies in Atomic Physics, edited by E. McDANIEL and M. McDOWELL (Elsevier, 1975) pp. 93–196.
  • Cheng and Childs (1985) K. T. Cheng and W. J. Childs, Phys. Rev. A 31, 2775 (1985).
  • Dixit et al. (2008) G. Dixit, H. Nataraj, B. Sahoo, R. Chaudhuri,  and S. Majumder, Physical Review A 77, 012718 (2008).
  • Yu et al. (2020b) Y. M. Yu, B. K. Sahoo,  and B. B. Suo, Phys. Rev. A 102, 062824 (2020b).
  • Gossel et al. (2013) G. Gossel, V. Dzuba,  and V. Flambaum, Physical Review A 88, 034501 (2013).