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

Time keeping and searching for new physics using metastable states of Cu, Ag and Au

V. A. Dzuba1, Saleh O. Allehabi1, V. V. Flambaum1, Jiguang Li2 and S. Schiller3 1School of Physics, University of New South Wales, Sydney 2052, Australia 2Institute of Applied Physics and Computational Mathematics, 6 Huayuan Road, Haidian District, Beijing, China 3Institut für Experimentalphysik, Heinrich-Heine-Universität Düsseldorf, 40225 Düsseldorf, Germany
Abstract

We study the prospects of using the electric quadrupole transitions from the ground states of Cu, Ag and Au to the metastable state D5/22{}^{2}{\rm D}_{5/2} as clock transitions in optical lattice clocks. We calculate lifetimes, transition rates, systematic shifts, and demonstrate that the fractional uncertainty of the clocks can be similar to what is achieved in the best current optical clocks. The use of these proposed clocks for the search of new physics, such as time variation of the fine structure constant, search for low-mass scalar dark matter, violation of Local Position Invariance and violation of Lorenz Invariance is discussed.

I Introduction

Using optical clock transitions for searching for new physics beyond the standard model is a promising area of research. A hypothetical manifestation of new physics at low energy is expected to be very small. Therefore, the highest possible accuracy of the measurements is needed. Fractional uncertainty of the best optical clocks currently is around 1×1\times 101810^{-18} Ludlow ; Chou ; Beloy1 ; Beloy2 ; Ushijima ; Nicholson ; Katori , the highest accuracy so far achieved in the history of measurements. However, apart from a few exceptions (Hg+, Yb+) the best optical clocks are not sensitive to new physics such as time variation of the fine structure constant, violation of Local Position Invarance (LPI) and violation of Local Lorentz Invariance (LLI), etc. qdot ; EEP ; Lia ; Yb+LLI . LPI, LLI and the Weak Equivalence Principle form the Einstein Equivalence Principle, the foundation of General Relativity. Several ideas were proposed to combine high accuracy of optical clocks with high sensitivity to new physics. These include the use of the highly charged ions (HCI) H-likeHCI ; HCIa ; HCIb ; HCIc , nuclear clocks th-clock , and metastable atomic states with large value of the total angular momentum JJ (J>1J>1qdot ; Hg+clock0 ; Hg+clock ; Yb-DFS ; Yb+clock . These states are connected to the ground state via transitions which correspond to single-electron transitions with large change of the single-electron total angular momentum jj. The large Δj\Delta j is what makes the transition to be sensitive to the variation of the fine structure constant (see e.g. DFW99 ). For example, in the present work we consider transitions between the nd10(n+1)s2nd^{10}(n+1)s\ ^{2}S1/2 ground state and the nd9(n+1)s22nd^{9}(n+1)s^{2}\ {}^{2}D5/2 excited metastable state. This is roughly the s1/2s_{1/2} to d5/2d_{5/2} transition with Δj\Delta j=2.

The energy diagrams for the five lowest states of Cu, Ag and Au studied in this work are presented in Fig. 1. The metastable state of interest (2D5/2) is the first excited state for Cu and Au. In Ag the 2P1/2o{}^{\rm o}_{1/2} state lies below the 2D5/2 clock state. However, this has no significance since the states are very weakly connected (by the E3, M2 or hyperfine-induced E1 transitions). The clock transition in Ag was studied experimentally in Ref. AgZ2 .

Sensitivity of the above metastable states to variation of the fine structure constant was studied before qdot . In this work we further study the states in terms of their suitability for high accuracy measurements and sensitivity to other manifestations of new physics, such as LPI violation and LLI violation.

Refer to caption
Figure 1: Energy diagram (approximately to scale) for the lowest states of Cu (I=3/2I=3/2), Ag (I=1/2I=1/2) and Au (I=3/2I=3/2). Thick red lines indicate the upper clock states. Electric quadrupole (E2) clock transitions are shown as red dashed lines. Cooling transitions are shown as solid blue lines. The presence of leakage transitions (black dotted lines) implies the need for repumping (magenta lines).

II Calculations

We are mostly interested in five lowest states of Cu, Ag and Au shown on Fig. 1. Two out of the five states have excitations from a dd-shell. This means that the dd-shell is open and dd-electrons should be treated as valence ones. The total number of valence electrons, eleven, is too large for most of standard calculational approaches. We use a version of the configuration interaction (CI) method specifically developed for such systems (the CIPT method CIPT ). In this method off-diagonal matrix elements of the CI Hamiltonian between highly excited states are neglected. This allows to reduce the CI matrix to an effective matrix of a small size in which contribution from high states is included perturbatively (see Ref. CIPT for details).

We perform the calculations in the VN1V^{N-1} approximation, with one electron removed from initial relativistic Hartree-Fock (HF) calculations to obtain the potential for calculating single-electron basis states. Given the ground electronic configurations of the atoms here discussed, at first sight the best choice is the removal of the valence ss electron. It turns out, however, that much better accuracy is achieved if a dd electron is removed instead. Thus, we perform the HF calculations for the [Ca]3d94s3d^{9}4s configuration of Cu, the [Sr]4d95s4d^{9}5s configuration of Ag, and the [Yb]5d96s5d^{9}6s configuration of Au. The B-spline technique B-spline is used to construct single-electron basis states above the core. Many-electron states for the CIPT calculations are constructed by exciting one or two electrons from a reference configuration and then using the resulting configurations to build all corresponding many-electron states of definite value of the total angular momentum JJ and its projection JzJ_{z}. States corresponding to about a hundred lowest non-relativistic configurations go into the effective CI matrix, while higher states are treated perturbatively. Note that our calculations are completely relativistic. We only use non-relativistic configurations to simplify the procedure of generating many-electron basis states. In the list of non-relativistic configurations each of them is subsequently replaced by a corresponding set of relativistic configurations. E.g., the 5d96s6p5d^{9}6s6p configuration is replaced by four relativistic ones, the 5d3/245d5/256s6p1/25d_{3/2}^{4}5d_{5/2}^{5}6s6p_{1/2}, 5d3/235d5/266s6p1/25d_{3/2}^{3}5d_{5/2}^{6}6s6p_{1/2}, 5d3/245d5/256s6p3/25d_{3/2}^{4}5d_{5/2}^{5}6s6p_{3/2}, and 5d3/235d5/266s6p3/25d_{3/2}^{3}5d_{5/2}^{6}6s6p_{3/2} configurations.

To calculate transition amplitudes we use the well-known random phase approximation (RPA, see, e.g. CPM ). The RPA equations for a single-electron state have the form

(HHFϵc)δψc=(F^+δVFN1)ψc.(H^{\rm HF}-\epsilon_{c})\delta\psi_{c}=-(\hat{F}+\delta V_{F}^{N-1})\psi_{c}. (1)

Here HHFH^{\rm HF} is the relativistic Hartree-Fock Hamiltonian, index cc numerates single-electron states, F^\hat{F} is the operator of an external field, δψc\delta\psi_{c} is a correction to the state cc due to an external field, δVFN1\delta V_{F}^{N-1} is the correction to the self-consistent Hartree-Fock potential due to the external field. The same VN1V^{N-1} potential is used in RPA and HF calculations. The RPA equations (1) are solved self-consistently for all states cc in the core. Transition amplitudes are found as matrix elements between many-electron states found in the CIPT calculations for the effective operator of an external field

Aab=b|F^+δVFcore|a.A_{ab}=\langle b|\hat{F}+\delta V^{\rm core}_{F}|a\rangle. (2)

Note that valence states are included in VFN1V^{N-1}_{F} in (1) but non included in the effective operator for valence states δVFcore\delta V^{\rm core}_{F} in (2).

The rates of spontaneous emission are given in atomic units by

ΓE1,M1=43(αω)3AE1,M122J+1,\Gamma_{\rm E1,M1}=\frac{4}{3}(\alpha\omega)^{3}\frac{A^{2}_{\rm E1,M1}}{2J+1}, (3)

for electric dipole (E1) and magnetic dipole (M1) transitions, and by

ΓE2,M2=115(αω)5AE2,M222J+1,\Gamma_{\rm E2,M2}=\frac{1}{15}(\alpha\omega)^{5}\frac{A^{2}_{\rm E2,M2}}{2J+1}, (4)

for electric quadrupole (E2) and magnetic quadrupole (M1) transitions. In these formulas, α\alpha is the fine structure constant, ω\omega is the energy difference between the lower and upper states, AA is the amplitude of the transition (2), JJ is the total angular momentum of the upper state. The magnetic amplitudes AM1A_{\rm M1} and AM2A_{\rm M2} are proportional to the Bohr magneton μB=|e|/2mc\mu_{B}=|e|\hbar/2mc. Its numerical value in Gaussian-based atomic units is μB=α/23.65×103\mu_{B}=\alpha/2\approx 3.65\times 10^{-3}. The lifetimes of excited states are calculated by τa=2.4189×1017/bΓab\tau_{a}=2.4189\times 10^{-17}/\sum_{b}\Gamma_{ab}, where τa\tau_{a} is the lifetime of atomic state aa in seconds, the summation goes over all possible transitions to lower states bb, the transition probabilities Γab\Gamma_{ab} are given by (3) or (4).

Energy levels and lifetimes of the five lowest states of Cu, Ag and Au are presented in Table 1. Lifetimes were calculated using transition amplitudes and probabilities from Table 2. Given the complexity of the considered systems the agreement between theory and experiment for the energies can be regarded as very good. However, the theoretical values for the lifetimes and transition rates are less accurate. This is partly due to the fact that the current version of the computer code includes perturbation correction to the energy but does not include a corresponding correction to the wave function CIPT . States with no excitation from the upper dd-shell (e.g., the 5f106p1/2,3/25f^{10}6p_{1/2,3/2} states of Au) can be treated more accurately within different approaches, for example with the use of the correlation potential method CPM . The main advantage of the current approach is that it can be used for any state of considered atoms, including states with excitations from the upper dd-shell where most of other methods would not work.

Table 1: Excitation energies (EE, cm-1) and lifetime for the lowest states of Cu, Ag and Au.
Energy [cm-1] Lifetime
NN Conf. Term NIST NIST Present work Present work
Cu
1 3d104s3d^{10}4s 2S1/2 0 0 \infty
2 3d94s23d^{9}4s^{2} 2D5/2 11203 10521 44.9 s
3 3d94s23d^{9}4s^{2} 2D3/2 13245 12270 7.3 s
4 3d104p3d^{10}4p 2Po1/2 30535 29489 7.1 ns
5 3d104p3d^{10}4p 2Po3/2 30784 31115 6.9 ns
Ag
1 4d105s4d^{10}5s 2S1/2 0 0 \infty
2 4d105p4d^{10}5p 2Po1/2 29552 29495 6.6 ns
3 4d105p4d^{10}5p 2Po3/2 30473 30451 6.1 ns
4 4d95s24d^{9}5s^{2} 2D5/2 30242 32480 192 ms
5 4d95s24d^{9}5s^{2} 2D3/2 34714 36430 174 μ\mus
Au
1 5d106s5d^{10}6s 2S1/2 0 0 \infty
2 5d96s25d^{9}6s^{2} 2D5/2 9161 10671 43.7 s
3 5d96s25d^{9}6s^{2} 2D3/2 21435 22096 33 ms
4 5d106p5d^{10}6p 2Po1/2 37359 38853 4.1111Experimental values are 6.0(1) ns for the 2P1/2o{}^{\rm o}_{1/2} state and 4.6(2) ns for the 2P3/2o{}^{\rm o}_{3/2} state Hannaford . ns
5 5d106p5d^{10}6p 2Po3/2 41175 43028 3.3111Experimental values are 6.0(1) ns for the 2P1/2o{}^{\rm o}_{1/2} state and 4.6(2) ns for the 2P3/2o{}^{\rm o}_{3/2} state Hannaford . ns
Table 2: Transition amplitudes and probabilities for all possible transitions between the five lowest states of Cu, Ag and Au.
Transition Type ω\hbar\omega [cm-1] |A||A| [a.u] Γ\Gamma [s-1]
Cu
2-1 E2 11203 2.6032.603 2.23\cdot10-2
3-1 M1 13245 0.0002μB0.0002\mu_{B} 6.27\cdot10-7
4-1 E1 30535 2.2172.217 1.42\cdot10+8
5-1 E1 30784 3.1403.140 1.45\cdot10+8
3-2 M1 2043 1.549μB1.549\mu_{B} 0.138
5-2 E1 19581 0.5460.546 1.13\cdot10+6
4-3 E1 17290 0.4040.404 8.48\cdot10+5
5-3 E1 17538 0.1740.174 8.23\cdot10+4
5-4 M1 248 1.154μB1.154\mu_{B} 1.38\cdot10-4
Ag
2-1 E1 29552 2.4102.410 1.52\cdot10+8
3-1 E1 30473 3.3823.382 1.64\cdot10+8
4-1 E2 30242 3.3253.325 5.22
5-1 M1 34714 0.0002μB0.0002\mu_{B} 1.13\cdot10-5
3-2 M1 921 1.154μB1.154\mu_{B} 7.01\cdot10-3
5-2 E1 5162 0.2740.274 1.04\cdot10+4
4-3 E1 230 0.3680.368 0.839111Experimental value is 1.6(6) s-1 Aglasercooling .
5-4 M1 4472 1.549μB1.549\mu_{B} 1.45
5-3 E1 4242 0.1190.119 5.45\cdot10+2
Au
2-1 E2 9161 4.3594.359 2.29\cdot10-2
3-1 M1 21435 0.0008μB0.0008\mu_{B} 4.25\cdot10-5
4-1 E1 37359 2.1532.153 2.45\cdot10+8
5-1 E1 41175 2.9232.923 3.02\cdot10+8
3-2 M1 12274 1.549μB1.549\mu_{B} 29.9
5-2 E1 32013 0.9830.983 1.61\cdot10+6
4-3 E1 15924 0.5040.504 1.04\cdot10+6
5-3 E1 19739 0.2430.243 2.30\cdot10+5
5-4 M1 3816 1.141μB1.141\mu_{B} 0.488

III Analysis

III.1 Clock transitions

Cu has two long-lived metastable states (N=2,3N=2,3 in Tab. 1), Ag has one (N=4N=4), and Au has one (N=2N=2). Only the states of Cu and Au have lifetimes substantially larger than 1 s, comparable to those of the currently used Sr and Yb lattice clocks. The 0.2 s lifetime of the Ag upper clock state may turn out to be a competitive disadvantage for achieving top performance. In the following we consider the clock transitions between these four states and their respective ground states. Note that the two clocks states of Cu are very similar and therefore in most of cases we present the data for only the 2D5/2 state.

III.2 Hyperfine structure

The atoms considered here all exhibit hyperfine structure in the ground state, in the clock state and in the excited state addressed in laser cooling. The nuclear spins are given in Tab. 3. The hfs splitting is given by Landau

EHFS(F)\displaystyle E_{\rm HFS}(F) =\displaystyle= A2F(F+1)+B2[F2(F+1)2+\displaystyle\frac{A}{2}F(F+1)+\frac{B}{2}\left[F^{2}(F+1)^{2}+\right.
+\displaystyle+ F(F+1)[12J(J+1)2I(I+1)]].\displaystyle\left.F(F+1)[1-2J(J+1)-2I(I+1)]\right].

The total angular momentum is 𝐅=𝐉+𝐈\mathbf{F}=\mathbf{J}+\mathbf{I}, where II is nuclear spin. AA and BB are magnetic dipole and electric quadrupole hfs constants, respectively. They are reported in Tab. 4.

Table 3: Stable isotopes with non-zero nuclear spin (II) and possible values of total angular momentum FF (𝐅=𝐈+𝐉\mathbf{F}=\mathbf{I}+\mathbf{J}) for ground and clock states of Cu, Ag, Au.
Isotopes Transition II FF for GS FF for CS
63,65Cu, 197Au 2S21/2{}_{1/2}-^{2}D5/2 3/2 1,2 1,2,3,4
63,65Cu 2S21/2{}_{1/2}-^{2}D3/2 3/2 1,2 0,1,2,3
107,109Ag 2S21/2{}_{1/2}-^{2}D5/2 1/2 0,1 2,3
Table 4: Magnetic dipole (AA) and electric quadrupole (BB) hfs constants (MHz) used in the calculation of the second-order Zeeman shift.
Atom Ground state Clock state Reference
AA BB AA BB
63Cu 2S1/2 5863 0 2D5/2 749.1 186.0 Cu-hfs ; Cu-hfs1
63Cu 2S1/2 5863 0 2D3/2 1851.0 137.4 Cu-hfs ; Cu-hfs1
107Ag 2S1/2 -1713 0 2D5/2 -126 0 AgZ2
197Au 2S1/2 3050 0 2D5/2 80.2 -1049 Au-hfs

For example, the hyperfine structure of Au has been studied experimentally with high precision in the 1960s, and has also been calculated Itano ; Bieron . The hyperfine splitting between F=1,2F=1,2 in the ground state amounts to 6.10 GHz Dahmen . The splittings in 2D5/2 are Au-hfs

F=1F=2F=1\leftrightarrow F=2: 1.00 GHz,

F=2F=3F=2\leftrightarrow F=3: 0.71 GHz, and

F=3F=4F=3\leftrightarrow F=4: 0.52 GHz.

The splittings in 2D3/2 are Blachman

F=0F=1F=0\leftrightarrow F=1: 1.11 GHz,

F=1F=2F=1\leftrightarrow F=2: 1.31 GHz, and

F=2F=3F=2\leftrightarrow F=3: 0.31 GHz.

III.3 Laser cooling of Cu, Ag and Au

III.3.1 Silver

Silver has been laser-cooled Aglasercooling . Here, the cooling scheme is straightforward: the cooling transition is between ground and second excited state, 2S21/2{}_{1/2}\rightarrow^{2}P3/2, so that there are no leakage channels to other electronic states. No repumper laser fields are needed.

III.3.2 Gold

A scheme for laser cooling of Au is presented in Fig. 2. The main cooling transition is the electric dipole transition between the ground state and the excited odd-parity 2P1/2o{}^{\rm o}_{1/2} state. Compared to using 2P3/2o{}^{\rm o}_{3/2} as upper level, the advantage is that only one repumper is needed and that the longer cooling wavelength is experimentally advantageous. There is leakage from the 2P1/2o{}^{\rm o}_{1/2} to the 2D3/2 state by another electric dipole transition (434\rightarrow 3). Therefore, without repumping only \sim 250 cycles are possible. With repumping (628 nm) the cooling may go for as long as needed. Another leakage channel is too weak to affect the scheme.

III.3.3 Copper

A cooling scheme similar to silver can be considered for copper: 2S(Fg=2)1/22{}_{1/2}(F_{g}=2)\rightarrow^{2}P(Fe=3)3/2{}_{3/2}(F_{e}=3).

III.3.4 Additional remarks

Optical lattice clocks require the cooling of atoms to the μ\muK level for efficient loading of the optical lattice with the atoms. Therefore, after cooling on the strong E1 transition to a temperature on the order of 1 mK, a second cooling process utilizing a weak transition should follow (”narrow-linewidth cooling”). One option is to cool on the 2S1/2{}_{1/2}\rightarrow 2D3/2 transition (131-3 for Cu, 151-5 for Ag, 131-3 for Au). These are M1 transitions and are very weak. However, the strengths could be increased and the lifetime of the 2D3/2 states shortened by E1 coupling them to the respective 2P states using appropriate waves.

The hyperfine structure in both lower and upper laser cooling levels will typically require additional repumper fields to optimize cooling efficiency. We shall not discuss such experimental details here.

Finally, we note that copper and silver atoms have been cooled using buffer-gas cooling CuAg-cool .

Refer to caption
Figure 2: Details of the level scheme of Au197{}^{197}{\rm Au} (I=3/2I=3/2) (not to scale) with proposed laser cooling. The hyperfine structure is shown schematically. The magenta arrow shows the repumper transition. Narrow-linewidth laser cooling is not shown. The clock transition (dashed red line) is composed of several hyperfine components.

III.4 Polarizabilities, black-body radiation shifts and magic frequencies

Knowledge of the atomic polarizabilities for both states of the clock transition is important for estimation of the frequency shift caused by black-body radiation and for finding the so called magic frequency of the lattice laser field, i.e. the frequency at which the dynamic polarizabilities of both states are equal, causing no frequency shift.

The static scalar polarizability αv(0)\alpha_{v}(0) of an atom in state vv is given by

αv(0)=23(2Jv+1)n|v||D||n|2EnEv,\alpha_{v}(0)=\frac{2}{3(2J_{v}+1)}\sum_{n}\frac{|\langle v||D||n\rangle|^{2}}{E_{n}-E_{v}}, (6)

where DD is the electric dipole operator with the RPA correction (see the previous section), and the summation goes over the complete set of excited many-electron states.

Static scalar polarizabilities of the ground states of Cu, Ag and Au are known from a number of calculations and measurements pol0 . Table 5 presents the recommended values taken from Ref. pol0 . In contrast, to the best of our knowledge, there is no similar data for the upper clock states of Cu, Ag and Au. Therefore, we performed the calculations using two different approaches.

In the first approach we stay within the CIPT method and calculate twenty odd-parity states for each value of the total angular momentum JJ which satisfies the electric dipole selection rules for the transitions from the ground and clock states (J=1/2,3/2,5/2,7/2J=1/2,3/2,5/2,7/2). Then we use the formula (6) to perform the calculations for both states. These calculations show three important things: (a) there is good agreement with other data for the ground state, (b) there is good saturation of the summation in (6), (c) the summation for the clock states is strongly dominated by the transitions to the states of the 5d96s6p5d^{9}6s6p configuration (we use the Au atom as an example).

The last fact implies that a different approach can be used, previously suggested for atoms with open ff-shells AKozlov . In this second approach we use the fact that the sum (6) is dominated by the 6s6s - 6p6p transitions while the open 5d95d^{9} subshell remains unchanged. Therefore, the open dd-shell is attributed to the core and treated as a closed shell with an occupational number of 0.9. The atom is treated as a system with two external electrons above the closed-shell core and an appropriate CI+MBPT CI+MBPT method is used (see Ref. AKozlov for more details). The advantage of this approach is the efficient completeness of the basis with two electron excitations. The shortcoming is the omission of the transition amplitudes involving excitations from the dd-shell. In contrast, the CIPT approach includes all amplitudes; however the summation in (6) is truncated much earlier.

In the end, both approaches give similar results. The results for the clock states are presented in Table 5 together with estimated uncertainties. For these estimations we used a comparison of the two approaches for the clock states as well as comparison of the CIPT calculation with other data for the ground states.

The results of the calculations indicate that the values of the polarizabilities of the clock states of Cu, Ag and Au are similar to those of the ground state. This is a non-standard situation. More often, the polarizabilities of excited states are larger. Indeed, the higher is the state on the energy scale, the smaller is the energy denominator in (6). Present results can be explained by the fact that summation in (6) is dominated by the states of the 5d10np5d^{10}np configurations for the ground state (we use Au again as an example) and by the states of the 5d96s6p5d^{9}6s6p configuration for the clock state. The later states are higher on the energy scale.

The BBR shift is given by (see, e.g. BBR )

δνBBR=215(απ)3T4(αc(0)αg(0)),\delta\nu_{\rm BBR}=-\frac{2}{15}(\alpha\pi)^{3}T^{4}\left(\alpha_{c}(0)-\alpha_{g}(0)\right), (7)

where α\alpha is the fine structure constant, TT is temperature, αc(0)\alpha_{c}(0) and αg(0)\alpha_{g}(0) are static scalar polarizabilities of the clock and ground states. For simplicity, we do not include the dynamic correction to the BBR shift. For the more complete formula see, e.g. BBR .

The BBR shift is proportional to the difference between the polarizabilities of the two states. The similarity of the polarizabilities implies a substantial cancellation of the black-body radiation (BBR) frequency shift, a very favourable effect. Since the differences are smaller or close to the uncertainty of both numbers we can only give upper limits for the BBR shifts. The results are presented in Table 5. These limits are lower than the shifts in the standard ytterbium and strontium lattice clocks. More accurate estimations might be possible if the polarizabilities are measured or calculated to higher accuracy.

Magic frequencies can be found in the vicinity of every resonance for one of the polarizabilities, i.e. when the frequency of the lattice laser field is approximately equal to the excitation energy (energy denominator in (6)). The first magic frequency is near the first resonance for the ground state polarizability, i.e. ωm30535cm1\hbar\omega_{m}\simeq 30535\,{\rm cm}^{-1} for Cu, ωm29552cm1\hbar\omega_{m}\simeq 29552\,{\rm cm}^{-1} for Ag and ωm37359cm1\hbar\omega_{m}\simeq 37359\,{\rm cm}^{-1} for Au. Note that since the clock states have large values of the total angular momentum (J=5/2J=5/2), the magic frequencies would also depend on the quadrupole contribution to the polarizabilities. The current level of computational accuracy does not allow to find accurate values of the magic frequencies. Having more experimental data may help. In the vicinity of a resonance or a few resonances a semi-empirical formula can be used

αa(ω)αa(0)+23(2Jv+1)bAab2ωΔEab,\alpha_{a}(\omega)\approx\alpha^{\prime}_{a}(0)+\frac{2}{3(2J_{v}+1)}\sum_{b}\frac{A_{ab}^{2}}{\omega-\Delta E_{ab}}, (8)

where αa(0)\alpha^{\prime}_{a}(0) is chosen in such a way that αa(ω=0)\alpha_{a}(\omega=0) is equal to known (e.g. experimental) static polarizability of state aa. Summation in (8) goes over close resonances. If the static polarizability is known to sufficient accuracy and amplitudes AabA_{ab} of E1-transitions are extracted from experimental data or from accurate atomic calculations, then (8) can be used to find magic frequencies.

Table 5: Scalar static polarizabilities (in aB3a_{B}^{3}) and BBR frequency shift for clock transitions of Cu, Ag and Au. Δα\Delta\alpha is the difference between the theoretical value for the upper clock state and the experimental value of the lower clock state.
Atom αg(0)\alpha_{g}(0) αc(0)\alpha_{c}(0) Δα\Delta\alpha BBR (T=300T=300\,K)
Expt. pol0 CIPT CIPT CI+MBPT Final Δν\Delta\nu [Hz] Δν/ν\Delta\nu/\nu
Cu111State cc is the 2D5/2 clock state. 47(1) 54.5 46.8 42.9 45(8) 2(8) <0.09<0.09 <2.61016<2.6\cdot 10^{-16}
Ag 55(8) 51.8 45.9 49.5 47(2) -8(8) <0.14<0.14 <1.51017<1.5\cdot 10^{-17}
Au 36(3) 35.7 38.9 33.2 36(3) 0(4) <0.03<0.03 <5.61017<5.6\cdot 10^{-17}

III.5 Stark, quadrupole and Zeeman shifts

Interaction of atomic electrons with external electric field and its gradient lead to Stark and electric quadrupole shifts of transition frequencies. These shifts are tiny in optical lattice clocks 111On the inner surface of a metallic vacuum chamber there can be spatial variations of the electrostatic potential of order 0.1 V. The typical size of a vacuum chamber may be 10 cm. Thus, the electric field gradient is smaller than 0.1 V/(10 cm)2)^{2}. Assuming a typical quadrupole moment Q1Q\sim 1 a.u. leads to a negligible quadrupole shift 105\sim 10^{-5} Hz. The corresponding Stark shift is 107\sim 10^{-7} Hz.. We consider the shifts in more details in the Appendix.

The linear Zeeman shift is given by the expression

ΔEF,Fz=gFμBBFz,\Delta E_{F,F_{z}}=g_{F}\mu_{B}BF_{z}, (9)

where gFg_{F} is the gg-factor of a particular hfs state. It is related to the electron gJg_{J}-factor by

gF=gJF,Fz=F,I,J|J^z|F,Fz=F,I,J/F.g_{F}=g_{J}\langle F,F_{z}=F,I,J|\hat{J}_{z}|F,F_{z}=F,I,J\rangle/F. (10)

Electron gg-factors have approximate values g1/22g_{1/2}\approx 2, g3/20.8g_{3/2}\approx 0.8, g5/21.2g_{5/2}\approx 1.2. More accurate values for Cu, Ag and Au can be found in NIST tables NIST . For a clock state with J=5/2J=5/2 and F=2F=2 we have g2=(11/12)g5/2=1.1g_{2}=(11/12)g_{5/2}=1.1. For a clock state with J=3/2J=3/2 and F=2F=2 we have g2=(1/2)g3/2=0.4g_{2}=(1/2)g_{3/2}=0.4.

The linear Zeeman shift can be avoided if only transitions between states with Fz=0F_{z}=0 are considered, as sug- gested in the past for clock operation. Alternatively, one can average over the transition frequencies with positive and negative FzF_{z} in order to cancel the linear shift. However, the large individual shifts will make it difficult to achieve an accurate cancellation.

A second-order Zeeman shift is unavoidable. Therefore, it is important to know its value. If we consider transitions between definite hfs components, then the shift is strongly dominated by transitions within the same hfs multiplet. The total shift is the difference between the second-order shifts in the clock and in the ground state. Both shifts are given by

δEF,Fz=\displaystyle\delta E_{F,F_{z}}= (11)
F=F±1,Fz|FFzIJ|J^z|FFzIJgJμBB|2ΔEHFS(F,F).\displaystyle\sum_{F^{\prime}=F\pm 1,F^{\prime}_{z}}\frac{|\langle F^{\prime}F^{\prime}_{z}IJ|\hat{J}_{z}|FF_{z}IJ\rangle g_{J}\mu_{B}B|^{2}}{\Delta E_{\rm{HFS}}(F,F^{\prime})}.

Here ΔEHFS(F,F)=E(FIJ)E(FIJ)\Delta E_{\rm{HFS}}(F,F^{\prime})=E(F\,I\,J)-E(F^{\prime}\,I\,J) is the hfs interval. It has a different sign depending on whether this is an up or down transition.

It follows from (III.2) that

ΔEHFS(F,F+1)=A(F+1)\displaystyle\Delta E_{\rm HFS}(F,F+1)=-A(F+1)-
B(2(F+1)2+12J(J+1)2I(I+1)),\displaystyle B\left(2(F+1)^{2}+1-2J(J+1)-2I(I+1)\right),

and

ΔEHFS(F,F1)=\displaystyle\Delta E_{\rm HFS}(F,F-1)=
AF+B(2F2+12J(J+1)2I(I+1)).\displaystyle AF+B\left(2F^{2}+1-2J(J+1)-2I(I+1)\right).

Using experimental values for AA and BB (see Table 4) we calculate the second-order Zeeman shift for Cu, Ag and Au. The results are presented in Tables 6 and 7. The shift for Ag was studied before AgZ2 . Our result differs from theirs; this may be due to a simple calculational error. Table 6 presents separate contributions from the shifts in the ground and excited states. One can see that the disagreement may come from the sing error in a particular contribution. Different signs are caused by energy denominators. For example, when we move from first to the second line of the table, the sign of the energy denominator for the ground state contribution changes and so does the contribution itself. Since other contribution remains the same, the total shift must change.

Table 7 shows the second-order Zeeman shift for 63Cu and 197Au. As in the case of 107Ag the shift is small. Note that Cu has one clock transition with both tiny quadratic shift coefficient and no linear shift. By measuring two or more Fgz=0Fez=0F_{gz}=0\rightarrow F_{ez}=0 Zeeman components and taking appropriate combinations of the corresponding transition frequencies the second-order shift may be substantially reduced.

The quadratic shift vanishes in the considered approximation for transitions between states with maximum value of FF and its projection FzF_{z} (see bottom lines of Table 7). This is because there are no terms in (11) which would satisfy the selection rules. Note also that the (non-zero) numbers in Table 7 should be considered as rough estimations only. This is because of uncertainties of the experimental data for the electric quadrupole hfs constant BB, in particular for Cu Cu-hfs1 . The numbers can change several times depending on which set of data is used.

Table 6: Second-order Zeeman shift (mHz/(μT)2(\mu T)^{2}) for 107Ag and comparison with other calculations. Index gg is for the ground state, index cc is for the excited (clock) state. It is assumed that Fz=0F_{z}=0.
FcF_{c} FgF_{g} ΔEc\Delta E_{c} ΔEg\Delta E_{g} ΔEcΔEg\Delta E_{c}-\Delta E_{g} Ref. AgZ2
2 0 0.1860.186 0.1140.114 0.0720.072 0.070.07
2 1 0.1860.186 0.114-0.114 0.3010.301 0.070.07
3 1 0.186-0.186 0.114-0.114 0.072-0.072 0.3-0.3
Table 7: Second-order Zeeman shift (mHz/(μT)2(\mu T)^{2}) for 63Cu and 197Au. Gaps in the data mean that corresponding set of quantum numbers is not possible for the transition.
FgF_{g} FgzF_{gz} FcF_{c} FczF_{cz} ΔEcΔEg\Delta E_{c}-\Delta E_{g}
63Cu 63Cu 197Au
2D5/2 2D3/2 2D5/2
1 0 0 0 0.759-0.759
1 0 1 0 0.0870.087 0.7430.743 0.0230.023
1 0 2 0 0.1930.193 0.0580.058 0.0270.027
1 0 3 0 0.247-0.247 0.0250.025 0.0500.050
2 0 0 0 0.792-0.792
2 0 1 0 0.0530.053 0.7100.710 0.041-0.041
2 0 2 0 0.1600.160 0.0240.024 0.037-0.037
2 0 3 0 0.281-0.281 0.009-0.009 0.014-0.014
2 0 4 0 0.0010.001 0.037-0.037
2 ±2\pm 2 3 ±2\pm 2 0.044-0.044 0.0040.004 0.0020.002
2 ±2\pm 2 3 ±3\pm 3 0.017-0.017 0.00.0 0.0040.004
2 ±2\pm 2 4 ±4\pm 4 0.00.0 0.00.0

IV Search for new physics

An exceptionally high accuracy of atomic clocks is a great advantage for using them in a search for new physics. The search is conducted by monitoring relative values of different atomic frequencies over a significant time interval. Hypothetical time-variation of the frequency ratio allows multiple interpretations. E.g., the interaction between low-mass scalar dark matter and ordinary matter may lead to oscillation of the fine structure constant and transient variation effect DM1 ; DM2 ; DM3 . In this section we consider time variation of the fine structure constant α\alpha (α=e2/c\alpha=e^{2}/\hbar c), Local Position Invariance (LPI) violation and Local Lorentz Invariance (LLI) violation.

IV.1 Time variation of the fine structure constant

It is convenient to present to parametrise the α\alpha-dependence of atomic frequencies by the formula ω=ω0+q[(αα0)21]\omega=\omega_{0}+q[(\frac{\alpha}{\alpha_{0}})^{2}-1] CJP , where α0\alpha_{0} and ω0\omega_{0} are present-day values of the fine structure constant and the frequency of the transition, qq is the sensitivity coefficient which comes from the calculations. To monitor possible frequency change one atomic frequency is measured against another. Then

tlnω1ω2=ω˙1ω1ω˙2ω2=(2q1ω12q2ω2)α˙α.\frac{\partial}{\partial t}\ln\frac{\omega_{1}}{\omega_{2}}=\frac{\dot{\omega}_{1}}{\omega_{1}}-\frac{\dot{\omega}_{2}}{\omega_{2}}=\left(\frac{2q_{1}}{\omega_{1}}-\frac{2q_{2}}{\omega_{2}}\right)\frac{\dot{\alpha}}{\alpha}. (12)

The value K=2q/ωK=2q/\omega is called an enhancement factor. It shows that if α\alpha changes in time then ω\omega changes KK times faster. Calculated values of qq and KK for different optical clock transitions are presented in Table 8. Note that we include not all known clock transitions but only those which are sensitive for α\alpha-variation searches. There are six transitions where |K|>1|K|>1. The largest values of |K||K| correspond to the smallest values of transition frequency ω\omega. It would be wrong to say that all these transitions are good for searching for α\alpha-variation. This is because the accuracy of the measurements is equally important (see also discussion in Ref. Yb-DFS ). The true figure of merit is the ratio of the relative frequency shift due to variation of α\alpha and the fractional uncertainty of the measurements, (q/ω)/(δω/ω)=q/δω(q/\omega)/(\delta\omega/\omega)=q/\delta\omega. This ratio does not depend on ω\omega. Therefore, looking for a large value of KK caused by the small value of ω\omega brings no benefit. The value of the relativistic energy shift qq is more important. Comparing the values of qq for different clock transitions (see Table 8) we see that the E2 clock transition in Au is essentially as good as the recently proposed new transitions in neutral ytterbium and only 30% smaller than the octupole transition in the ytterbium ion (Yb II).

Table 8: Sensitivity of clock transitions to variation of the fine structure constant (q,Kq,K), to LLI violation (reduced matrix element cT0(2)c\langle c||T_{0}^{(2)}||c\rangle of the tensor operator (16) for the upper state cc), and to LPI violation (relativistic factor RR). Note that gT0(2)g\langle g||T_{0}^{(2)}||g\rangle is zero for the ground state due to the small value of the total angular momentum J=1/2J=1/2.
Atom/ Transition ω\hbar\omega111NIST NIST . qq K=2q/ωK=2q/\hbar\omega cT0(2)c\langle c||T_{0}^{(2)}||c\rangle RR
Ion Lower state Upper state (cm-1) (cm-1) (a.u.) Present Other
Cu 3d104s23d^{10}4s\ ^{2}S1/2 - 3d94s223d^{9}4s^{2}\ {}^{2}D5/2 11202.56511202.565 4000-4000 0.71-0.71 48-48 0.980.98
Cu 3d104s23d^{10}4s\ ^{2}S1/2 - 3d94s223d^{9}4s^{2}\ {}^{2}D3/2 13245.44313245.443 1900-1900 0.29-0.29 37-37 0.990.99
Ag 4d105s24d^{10}5s\ ^{2}S1/2 - 4d95s224d^{9}5s^{2}\ {}^{2}D5/2 30242.06130242.061 11300-11300 0.75-0.75 41-41 0.930.93
Au 5d106s25d^{10}6s\ ^{2}S1/2 - 5d96s225d^{9}6s^{2}\ {}^{2}D5/2 9161.1779161.177 38550-38550 8.4-8.4 45-45 0.670.67
Hg II222Ref. CJP ; Yb-DFS . 5d106s25d^{10}6s\ ^{2}S1/2 - 5d96s225d^{9}6s^{2}\ {}^{2}D5/2 35514.62435514.624 52200-52200 2.94-2.94 0.680.68 0.20.2333Ref. EEP .
Yb444Ref. Yb-DFS . 4f146s214f^{14}6s^{2}\ {}^{1}S0 - 4f146s6p34f^{14}6s6p\ ^{3}P0o{}^{\rm o}_{0} 17288.43917288.439 27142714 0.310.31 1.121.12 1.201.20333Ref. EEP .
Yb444Ref. Yb-DFS . 4f146s214f^{14}6s^{2}\ {}^{1}S0 - 4f135d6s2J=24f^{13}5d6s^{2}J=2 23188.51823188.518 44290-44290 3.82-3.82 0.650.65 1.401.40444Ref. Yb-DFS .
Yb555Ref. SafronovaYb . 4f146s6p14f^{14}6s6p\ ^{1}P0o{}^{\rm o}_{0} - 4f135d6s2J=24f^{13}5d6s^{2}J=2 5900.0795900.079 43530-43530 15-15
Yb II666Ref. Yb+LLI ; YbIIq . 4f146s24f^{14}6s\ ^{2}S1/2 - 4f136s224f^{13}6s^{2}\ {}^{2}F7/2o{}^{\rm o}_{7/2} 21418.7521418.75 56737-56737 5.3-5.3 135-135 0.580.58 1.9-1.9333Ref. EEP .
Yb II666Ref. Yb+LLI ; YbIIq . 4f146s24f^{14}6s\ ^{2}S1/2 - 4f145d24f^{14}5d\ ^{2}D3/2 22960.8022960.80 1011810118 0.880.88 1010 1.421.42 1.481.48333Ref. EEP .
Yb II666Ref. Yb+LLI ; YbIIq . 4f136s224f^{13}6s^{2}\ {}^{2}F7/2o{}^{\rm o}_{7/2} - 4f145d24f^{14}5d\ ^{2}D3/2 1542.061542.06 66855-66855 87-87

IV.2 LPI violation

In the standard model extension, the term in the Hamiltonian responsible for the LPI violation can be presented in the form (see, e.g., Ref. EEP )

H^EEP=c0023Uc2K^,\hat{H}_{\rm EEP}=c_{00}\frac{2}{3}\frac{U}{c^{2}}\hat{K}, (13)

where c00c_{00} is the unknown parameter characterizing the magnitude of the LPI violation, UU is the gravitational potential, cc is the speed of light, K^=cγ0γjpj/2\hat{K}=c\gamma_{0}\gamma^{j}p_{j}/2 is the relativistic operator of kinetic energy in which γ0\gamma_{0} and γj\gamma^{j} are Dirac matrices, and 𝐩=i\mathbf{p}=-i\hbar\mathbf{\nabla} is electron momentum operator.

The presence of the term (13) in the Hamiltonian would manifest itself via a dependence of the atomic frequencies on the time in the year, caused by the changing Sun-Earth distance leading to change of the Sun’s gravitational potential UU. As in the case of the α\alpha-variation, at least two clock transitions are needed to measure one clock frequency against the other. The interpretation of the measurements is based on the formula EEP

Δω1ω1Δω2ω2=(R1R2)23c00ΔUc2,\frac{\Delta\omega_{1}}{\omega_{1}}-\frac{\Delta\omega_{2}}{\omega_{2}}=-(R_{1}-R_{2})\frac{2}{3}c_{00}\frac{\Delta U}{c^{2}}, (14)

where Δω\Delta\omega and ΔU\Delta U are the change of atomic frequencies and gravitational potential between the measurements, respectively. RR in (14) is the relativistic factor which describes the deviation of the kinetic energy EKE_{K} from the value given by the non-relativistic virial theorem (which states that EK=EE_{K}=-E, where EE is the total energy),

Rab=EK,aEK,bEaEb.R_{ab}=-\frac{E_{K,a}-E_{K,b}}{E_{a}-E_{b}}. (15)

The values of the factor RR are calculated in computer codes by varying the value of the kinetic energy operator in the Dirac equation (see Ref. EEP for details).

The results are very sensitive to the many-body effects which means that the effects should be treated very accurately or avoided. Otherwise the results are unstable. A good criterion for the reliability of the results is the achievement of the non-relativistic limit R=1R=1. This can be done by setting to zero the value of the fine structure constant α\alpha in the computer codes. It turns out that for complicated systems like those considered in present work the best results are obtained by simple estimations based on single-electron consideration. Namely, all clock transitions in Cu, Ag and Au can be considered as a ns(n1)d5/2ns\rightarrow(n-1)d_{5/2} single-electron transitions (n=4,5,6n=4,5,6 for Cu, Ag, Au respectively). Therefore, we just use single-electron energies of theses states in (15). We use the same approach for Hg+ and Yb+.

The results are presented in Table 8 together with results obtained earlier for other systems. Note, that the results for the transitions involving excitation from the 5d5d shell in Hg+ and the 4f4f shell in Yb+ are different from what was published before. The old calculations were based on a version of the CI method CI1 ; CI2 which contained a fitting parameter responsible for correct energy interval between states of different configurations. It was assumed that this parameter does not change under variation of the kinetic energy operator. We believe that present results are more reliable because they are free from any assumptions and because they reproduce the non-relativistic limit R=1R=1. Note that the values of RR for transitions in Yb and Yb+, which do not involve excitation from the 4f4f shell are in good agreement with previous calculations (R=1.12R=1.12 and R=1.42R=1.42, see Table 8). This means that present single-electron estimations work well and that accurate many-body calculations are possible for simple systems.

To study the LPI violation one needs to compare two clocks with different values of the relativistic factors RR (see formula (14)) over at least half of a year. Table 8 shows that there is a wide range of choices for such clock pairs. In particular, the Au clock is practically as good as the Hg+ clock which was used before for this purpose in combination with the Al+ clock Hg+clock0 ; EEP . The Cu and Ag clocks, which have RR values close to unity, can be used in combination with clocks with large relativistic effects, e.g. Yb or Yb+ clocks.

IV.3 LLI violation

The LLI violation term is a tensor operator

H^LLI=16C0(2)T0(2),\hat{H}_{\rm LLI}=-\frac{1}{6}C_{0}^{(2)}T_{0}^{(2)}, (16)

where C0(2)C_{0}^{(2)} is unknown constant and the relativistic form of the T0(2)T_{0}^{(2)} operator is given by T0(2)=cγ0(γjpj3γ3p3)T_{0}^{(2)}=c\gamma_{0}(\gamma^{j}p_{j}-3\gamma^{3}p_{3}).

To study the effect of the LLI violating term (16) one needs long-lived atomic states with large value of the total electron angular momentum JJ, J>1/2J>1/2. All clock states of Cu, Ag and Au satisfy this requirement. This term should cause a dependence of the atomic frequencies on the apparatus orientation in space (e.g., due to Earth rotation). Interpretation of the measurements requires knowing the values of the reduced matrix elements of the operator T0(2)T_{0}^{(2)} for the clock states. We calculate these matrix elements using the CIPT method to obtain wave functions and the RPA method to obtain effective operator for valence electrons.

The results are presented in Table 8. The results of earlier calculations for Yb II Yb+LLI are also presented for comparison. In contrast to the search of the α\alpha-variation and LPI violation, one clock state is sufficient for the search of the LLI violation. The comparison of frequencies is done for states with different projections of the total angular momentum JJ Lia ; Yb+LLI . The large value of the matrix element is important but it is not the most important parameter, e.g. the lifetime of the metastable state is even more important (see, e.g. Yb+LLI for more discussion). The calculations for Cu and Au show that these systems are also suitable for the search of the LLI violation. Ag appears less well suited because its upper clock state has a significantly shorter lifetime.

V Conclusion

We studied electric quadrupole transitions between ground and excited metastable states of Cu, Ag and Au and demonstrated that the transitions have all features of optical clock transitions. Important systematic effects such as black-body radiation shift, Stark and Zeeman shifts, etc. are similar to or smaller than in current top-performing optical clocks. On the other hand, some of the transitions are more sensitive to new physics beyond the standard model than the currently used neutral-atom optical clocks, or of complementary value. The studied effects included time variation of the fine structure constant, Local Lorentz Invariance violation and Local Position Invariance violation.


Acknowledgements.
This work was supported by the National Natural Science Foundation of China (Grant No. 11874090) and the Australian Research Council. V.A.D. would like to express special thanks to the Institute of Applied Physics and Computational Mathematics in Beijing for its hospitality and support. This research includes computations using the computational cluster Katana supported by Research Technology Services at UNSW Sydney.

Appendix A Stark and electric quadrupole shifts.

Stark shift of the frequency of the transition between atomic states aa and bb due to interaction with residual static electric field ε\varepsilon is

δωab=Δαab(0)(ε2)2,\delta\omega_{ab}=-\Delta\alpha_{ab}(0)\left(\frac{\varepsilon}{2}\right)^{2}, (17)

where Δαab(0)\Delta\alpha_{ab}(0) is the difference between static scalar polarizabilities of states aa and bb. The shift is quadratic in electric field and usually small. It is further suppressed for considered clock transitions due to small difference in the polarizabilities (see Table 5).

The energy shift due to a gradient of a residual static electric field ε\varepsilon is described by a corresponding term in the Hamiltonian

H^Q=12Q^εzz,\hat{H}_{Q}=-\frac{1}{2}\hat{Q}\frac{\partial\varepsilon_{z}}{\partial z}, (18)

where Q^\hat{Q} is the atomic quadrupole moment operator (Q^=|e|r2Y2m\hat{Q}=|e|r^{2}Y_{2m}, the same as for E2 transitions). The energy shift of a state with total angular momentum JJ is proportional of the atomic quadrupole moment of this state. It is defined as twice the expectation value of the Q^\hat{Q} operator in the stretched state

QJ=2J,Jz=J|Q^|J,Jz=J.Q_{J}=2\langle J,J_{z}=J|\hat{Q}|J,J_{z}=J\rangle. (19)

Calculations using the CIPT method for wave functions and the RPA method for the operator give the values QJ=0.431Q_{J}=0.431 a.u. for the 2D5/2 clock state of Cu, QJ=0.296Q_{J}=0.296 a.u. for the 2D3/2 clock state of Cu, QJ=0.966Q_{J}=0.966 a.u. for the clock state of Ag, and QJ=1.47Q_{J}=1.47 a.u. for the clock state of Au. Quadrupole moments of the ground states of these atoms are zero due to small value of the total electron angular momentum (JJ=1/2).

Consider transitions between hyperfine structure (hfs) components of the ground and clock states with definite values of the total angular momentum FF. The quadrupole shift is proportional to 3Fz2F(F+1)3F_{z}^{2}-F(F+1), where FzF_{z} is the projection of 𝐅\mathbf{F}. For F=3F=3 and Fz=±2F_{z}=\pm 2 this factor is zero and the quadrupole shift vanishes. Note that clock states with F=3F=3 exist for all stable isotopes of all three considered atoms (see Table 3). Using these states would lead to a linear Zeeman shift. It cancels out by averaging over the transition frequencies to the states with Fz=2F_{z}=-2 and Fz=2F_{z}=2.

References

  • (1) A. D. Ludlow, M. M. Boyd, J. Ye, E. Peik, and P. O. Schmidt, Rev. Mod. Phys. 87, 637 (2015).
  • (2) C. W. Chou, D. B. Hume, J. C. J. Koelemeij, D. J. Wineland, and T. Rosenband, Phys. Rev. Lett. 104, 070802 (2010).
  • (3) 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).
  • (4) K. Beloy, N. Hinkley, N. B. Phillips, J. A. Sherman, M. Schioppo, J. Lehman, A. Feldman, L. M. Hanssen, C. W. Oates, and A. D. Ludlow, Phys. Rev. Lett. 113, 260801 (2014).
  • (5) I. Ushijima, M. Takamoto, M. Das, T. Ohkubo, and H. Katori, Nature Photon. 9, 185 (2015).
  • (6) T. L. Nicholson, S. L. Campbell, R. B. Hutson, G. E. Marti, B. J. Bloom, R. L. McNally, W. Zhang, M. D. Barrett, M. S. Safronova, G. F. Strouse, W.L. Tew and J. Ye, Nature Commun. 6, 6896 (2015).
  • (7) N. Nemitz, T. Ohkubo, M. Takamoto, I. Ushijima, M. Das, N. Ohmae, and H. Katori, Nature Photonics 10, 258 (2016).
  • (8) V. V. Flambaum and V. A. Dzuba, Can. J. Phys. 87, 25 (2009).
  • (9) V. A. Dzuba and V. V. Flambaum, Phys. Rev. D 95, 015019 (2017).
  • (10) R. Shaniv, R. Ozeri, M. S. Safronova, S. G. Porsev, V. A. Dzuba, V. V. Flambaum, and H. Häffner, Phys. Rev. Lett. 120, 103202 (2018).
  • (11) V. A. Dzuba, V. V. Flambaum, M. S. Safronova, S. G. Porsev, T. Pruttivarasin, M. A. Hohensee, H. Häffner, Nature Physics 12, 465 (2016).
  • (12) S. Schiller, Phys. Rev. Lett. 98, 180801 (2007)
  • (13) J. C. Berengut, V. A. Dzuba, V. V. Flambaum, Phys. Rev. Lett. 105, 120801 (2010).
  • (14) J. C. Berengut, V. A. Dzuba, V. V. Flambaum, A. Ong, Phys. Rev. A 86, 022517 (2012).
  • (15) V. A. Dzuba and V. V. Flambaum, Hyperfine Interactions, 236, 79 (2015).
  • (16) C. J. Campbell, A. G. Radnaev, A. Kuzmich, V. A. Dzuba, V. V. Flambaum, and A. Derevianko, Phys. Rev. Lett. 108, 120802 (2012).
  • (17) T. Rosenband, D. B. Hume, P. O. Schmidt, C. W. Chou, A. Brusch, L. Lorini, W. H. Oskay, R. E. Drullinger, T. M. Fortier, J. E. Stalnaker, S. A. Diddams, W. C. Swann, N. R. Newbury, W. M. Itano, D. J. Wineland, and J. C. Bergquist, Science 319, 1808 (2008).
  • (18) Zhang Xiang, Lu Ben-Quan, Li Ji-Guang, and Zou Hong-Xin, Acta Physica Sinica 68, 043101 (2019).
  • (19) V. A. Dzuba, V. V. Flambaum, S. Schiller, Phys. Rev. A 98, 022501 (2018).
  • (20) C. Sanner, N. Huntemann, R. Lange, C. Tamm, E. Peik, M. S. Safronova, and S. G. Porsev, Nature 567, 204 (2019).
  • (21) V. A. Dzuba, V. V. Flambaum, J. K. Webb, Phys. Rev. A, 59, 230 (1999).
  • (22) T. Badr, M. D. Plimmer, P. Juncar, M. E. Himbert, Y. Louyer, and D. J. E. Knight, Phys. Rev. A 74, 062509 (2006).
  • (23) V. A. Dzuba, J. C. Berengut, C. Harabati, and V. V. Flambaum, Phys. Rev. A 95, 012503 (2017).
  • (24) W. R. Johnson, and J. Sapirstein, Phys. Rev. Lett. 57, 1126 (1986).
  • (25) V. A. Dzuba, V. V. Flambaum, P. G. Silvestrov, O. P. Sushkov, J. Phys. B: At. Mol. Phys. 20, 1399 (1987).
  • (26) P. Hannaford, P. L. Larkins and R. M. Lowe, J. Phys. B: At. Mol. Phys. 14, 2321 (1981).
  • (27) G. Uhlenberg, J. Dirscherl, H. Walther, Phys. Rev. A 62, 063404 (2000).
  • (28) L. D. Landau and E. M. Lifshitz, Quantum Mechanics. Non-relativistic Theory. Pergamon Press, Oxford, second edition (1965).
  • (29) J. Tenenbaum, I. Smilanski, S. Gabay, L. A. Levin, G. Erez, and S. Lavi, Opt. Comm. 32, 473 (1980).
  • (30) B. K. Ankush and M. N. Deo, J. Quant. Spect. Rad. Transfer 134, 21 (2014).
  • (31) W. J. Childs and L. S. Goodman, Phys. Rev. 141, 176 (1966).
  • (32) A. G. Blachman, D. A. Landman, A. Lurio, Phys. Rev. 161, 60(1967)
  • (33) W. M. Itano, J. Res. Natl. Inst. Stand. Technol. 105, 829 (2000).
  • (34) J. Bieroń, C. Froese Fischer, P. Indelicato, P. Jönsson, and P. Pyykkö, Phys. Rev. A 79, 052502 (2009).
  • (35) N. Brahms, B. Newman, C. Johnson, T. Greytak, D. Kleppner, and J. Doyle, Phys. Rev. Lett. 101, 103002 (2008).
  • (36) V. A. Dzuba, A. Kozlov, and V. V. Flambaum, Phys. Rev. A 89, 042507 (2014).
  • (37) V. A. Dzuba, V. V. Flambaum, and M. G. Kozlov, Phys. Rev. A, 54, 3948 (1996).
  • (38) S. G. Porsev and A. Derevianko, Phys. Rev. A 74, 020502(R) (2006); Erratum Phys. Rev. A 86, 029904(E) (2012).
  • (39) P. Schwerdtfeger and J. K. Nagle, Molecular Physics 117, 1200 (2019).
  • (40) A. Kramida, Yu. Ralchenko, J. Reader, and NIST ASD Team (2018). NIST Atomic Spectra Database (ver. 5.5.2), [Online]. Available: https://physics.nist.gov/asd [2018, January 10]. National Institute of Standards and Technology, Gaithersburg, MD.
  • (41) V. A. Dzuba, V. V. Flambaum, and M. V. Marchenko, Phys. Rev. A 68, 022506 (2003).
  • (42) H. Dahmen and S. Penselin, Zeitschrift f. Physik 200, 456(1967)
  • (43) A. Arvanitaki, J. Huang, and K. Van Tilburg, Phys. Rev. D 91, 015015 (2015).
  • (44) K. Van Tilburg, N. Leefer, L. Bougas, D. Budker, Phys. Rev. Lett. 115, 011802 (2015).
  • (45) Y. V. Stadnik and V. V. Flambaum, Phys. Rev. Lett. 115, 201301 (2015).
  • (46) V. V. Flambaum and V. A. Dzuba, Canadian Journal of Physics, 87, 25 (2009).
  • (47) M. S. Safronova, S. G. Porsev, C. Sanner, and J. Ye, Phys. Rev. Lett. 120, 173001 (2018).
  • (48) V. A. Dzuba and V. V. Flambaum, Phys. Rev. A 77, 012514 (2008).
  • (49) V. A. Dzuba and V. V. Flambaum, Phys. Rev. A 77, 012515 (2008).