An optimized ion trap geometry to measure quadrupole shifts of 171Yb+ clocks
Batra N1, 2, Sahoo B K3, De S1, 2, †,
Academy of Scientific and Innovative Research, CSIR-NPL Campus, New Delhi, India
CSIR-National Physical Laboratory, Dr. K. S. Krishnan Marg, New Delhi 110012, India
Theoretical Physics Division, Physical Research Laboratory, Navrangpura, Ahmedabad 380009, India

 

† Corresponding author. E-mail: subhadeep@mail.nplindia.org

Abstract
Abstract

We propose a new ion-trap geometry to carry out accurate measurements of the quadrupole shifts in the 171Yb ion. This trap will minimize the quadrupole shift due to the harmonic component of the confining potential by an order of magnitude. This will be useful to reduce the uncertainties in the clock frequency measurements of the 6s 2S1/2 → 4f136s2 2F7/2 and 6s 2S1/2 → 5d 2D3/2 transitions, from which we can deduce the precise values of the quadrupole moments (Θs) of the 4f136s2 2F7/2 and 5d 2D3/2 states. Moreover, it may be able to affirm the validity of the measured Θ value of the 4f136s2 2F7/2 state, for which three independent theoretical studies defer almost by one order of magnitude from the measurement. We also calculate Θs using the relativistic coupled-cluster (RCC) method. We use these Θ values to estimate the quadrupole shift that can be measured in our proposed ion trap experiment.

1. Introduction

Advances in trapping and laser control of a single ion[13] began a new era for the frequency standards in the optical range, which are aimed to achieve a fractional accuracy of 10−16–10−18. A number of ions, such as 199Hg+,[4] 171Yb+,[59] 115In+,[10] 88Sr+,[11,12] 40Ca+,[13,14] 27Al+,[15] etc. have been undertaken in the experiments to attain such promising optical frequency standards. Among them, 171Yb+ is unique in the sense that it has three potential optical transitions that can be used for the clocks.[16] Out of these, there are two narrow 6s 2S1/2(F = 0, mF = 0) → 5d 2D3/2(F = 2, mF = 0),[17,18] 6s 2S1/2(F = 0, mF = 0) → 5d 2D5/2(F = 2, mF = 0) [19] quadrupole (E2) transitions and an ultra-narrow 6s 2S1/2(F = 0, mF = 0) → 4f136s2 2F7/2(F = 3, mF = 0) octupole (E3) transition[9,20,21] with their respective wavelengths at 435.5 nm, 411 nm, and 467 nm. The transitions at the wavelengths 435.5 nm and 467 nm with low systematic shifts are the most suitable ones for precision frequency standards owing to their extremely small natural line-widths 3.02 Hz and 1 nHz, respectively. Their precisely measured transition frequencies νo have already been reported as 688358979309307.82(36) Hz[18] and 688358979309308.42(42) Hz[6] for the E2 transition and 642121496772645.36(39) Hz[9] and 642121496772644.91(37) Hz[6] for the E3 transition. These two transitions are endorsed by the international committee for weight and measures (CIPM) for the secondary representation of the standard international (SI) second owing to their being least sensitive to the external electromagnetic fields. Other than being a potential candidate for the frequency standards, Yb+ is also being considered for studying the parity non-conservation effect,[22,23] violation of the Lorentz symmetry,[24] searching for possible temporary variation of the fine structure constant,[25,26] etc.

In an atomic clock, the measured frequency of the interrogated transition is always different than its absolute value due to the systematics. The net frequency shift depends upon many environmental factors and experimental conditions, which may or may not be canceled out at the end. Thus, they need to be accounted to establish accurate frequency standards. Careful design of the ion trap is also very important for minimizing the systematics caused by the environmental factors.[27] Electric quadrupole shift is one of the major systematics when the states associated with the clock transition have finite quadrupole moments (Θs). In the 171Yb+ ion, the experimentally measured Θ value of the 5d 2D3/2 state[28] is 50 times larger than the 4f136s2 2F7/2 state. A recent theoretical study suggested a more precise Θ value of the 5d 2D3/2 state.[16] Three independent theoretical investigations[16,29,30] showed very large disagreements with the measured Θ value of the 4f136s2 2F7/2 state.[8] Therefore, it is imperative to carry out further investigations to attain more precise and reliable values of Θs and probe the reason for the anomalies between the calculated and the experimental results. Here we also perform another calculation using the relativistic coupled-cluster (RCC) method to evaluate the Θ values of the 5d 2D3/2 and 4f136s2 2F7/2 states to use them in the present analysis. We also analyze in detail the suitable confining potentials, electric fields, and field gradients that can create a nearly ideal quadrupole trap condition for carrying out precise measurements of the Θ values. Using these inputs, we estimate typical values of the quadrupole shifts of the 6s 2S1/2(F = 0, mF = 0) → 5d 2D3/2(F = 2, mF = 0) and 6s 2S1/2(F = 0, mF = 0) → 4f136s2 2F7/2(F = 3, mF = 0) clock transitions considering a number of ion trap geometries and discuss their possible pros and cons to make an appropriate choice. This analysis identifies a suitable geometry of the end cap ion trap to measure Θs of the 5d 2D3/2 and 4f136s2 2F7/2 states of Yb+, which is being developed at the National Physical Laboratory (NPL), India.[31,32]

2. Electric quadrupole shift

Electric quadrupole shift ΔνQ to an atomic state with angular momentum F arises due to the interaction of the quadrupole moment Θ(γ,F) with an applied external electric field gradient ∇E, where γ represents the other quantum numbers of the state. A non-zero atomic angular momentum results in a non-spherical charge distribution, thus the atom acquires higher order moments. Following this, it is advantageous to choose states with J < 1 or F < 1 in a clock transition for which Θ = 0. However, the excited states of the 6s 2S1/2(F = 0, mF = 0) → 5d 2D3/2(F = 2, mF = 0) and 6s 2S1/2(F = 0, mF = 0) → 4f136s2 2F7/2(F = 3, mF = 0) clock transitions have J = 3/2, F = 2 and J = 7/2, F = 3, resulting in nonzero quadrupole shifts. These shifts can be estimated by calculating the expectation value of the Hamiltonian[33]

where ranks of the E and Θ tensors are two and their components are indicated by subscript q. The expectation value of HQ in reduced form can be expressed as[34]

where mF is the magnetic quantum number, D0q are the rotation matrix elements of the projecting components of ∇E in the principal axis frame that are used to convert from the trap axes to the lab frame,[35] Θ(γ,J) is the quadrupole moment of the atomic state with angular momentum J, and

Here the quantities within ( ) and {} represent the 3j and 6j coefficients, respectively. Both the excited states of the above mentioned clock transitions acquire Due to the axial symmetry of the trap, the frequency shift contributions from D0±1 cancel with each other, thus finite contributions come only from the D00 = (3 cos2θ −1)/2 and components for the Euler angles θ and ϕ. In this work, we use our calculated Θ values for the 5d 2D3/2(F = 2) and 4f136s2 2F7/2(F = 3) states to find out the optimum electrode geometries that can produce nearly-ideal quadrupole confining potentials after interacting with the resultant electric field gradients of the non-ideal multipole potentials with the order of multipole k of the effective trapping potentials.

3. Methods for calculation

To calculate the atomic state wave functions for the determination of the Θ(γ,J) values, we adopt the Bloch approach.[36] Following this approach, we express the wave function |Ψv〉 of the 5d 2D3/2 state with the valence orbital v in the 5d3/2 orbital as

and the wave function |Ψa〉 of the 4f136s2 2F7/2 state with the valence orbital a in the 4 f7/2 orbital as

of the Yb+ ion, where Ωv and Ωa are the wave operators for the corresponding reference states |Φv〉 and |Φa〉, respectively. We use two different ways to construct these reference states. For the computational simplification, we choose the working reference states as the Dirac–Hartree–Fock (DHF) wave functions of the closed-shell configurations (denoted by ) in place of the above mentioned respective actual reference states |Φv〉 and |Φa〉 having open valence orbitals. In our calculations, we have obtained for the [4f14] configuration, while is calculated with the [4f14]6s2 configuration. Then, the actual reference states are obtained by appending the valence orbital v = 5d3/2 and removing the spin partner of the valence orbital a = 4 f7/2 of the respective references. In the second quantization formalism, it is given as

We employ the Dirac–Coulomb Hamiltonian for the calculations, which in the atomic unit (a.u.) is given by

where α and β are the usual Dirac matrices and Vn(r) represents the nuclear potential.

In a perturbative procedure, Ωv/a can be expressed as

where and χv/a are responsible for carrying out excitations from due to the residual interaction Vr = HH0 for the DHF Hamiltonian H0. In a series expansion, they are given as

where superscript k refers to the number of times Vr being considered in the many-body perturbation theory (MBPT(k) method). The kth order amplitudes for the and χv/a operators are obtained by solving the following equations:[36]

where the projection operators are defined as Pv/a = |Φv/a〉〈Φv/a|, and Qv/a = 1 – Pv/a. The exact energies for the states having the closed-shell and open-shell configurations are evaluated using the effective Hamiltonians

In the RCC theory framework, the wave functions of the considered states are expressed as (e.g., see Refs. [16] and [37])

where Tv and Ta excite the core electrons from the new reference states and respectively, to account for the electron correlation effects. The Sv and (eTv − 1)Sv operators excite electrons from the valence and the valence with core orbitals from |Φv〉. Similarly, the Ra and (eTa − 1)Ra operators excite electrons from the valence and the valence with core orbitals from |Φa〉. In this work, we consider only the singles and doubles excitations in the RCC theory (CCSD method), which are identified by the RCC operators with the subscripts 1 and 2, respectively, as

When only the linear terms are retained in Eqs. (14) and (15) with the singles and doubles excitations approximation in the RCC theory, we refer to it as the LCCSD method. The amplitudes of the above operators are evaluated by the equations

where and are the excited state configurations with respect to the DHF states and |Φv/a〉, respectively, and with subscript c representing the connected terms only. Here and are the attachment energy of the electron in the valence orbital v and the ionization potential of the electron in the orbital a, respectively. With Eqs. (12) and (13), ΔEv/a are evaluated as

To improve the quality of the wave functions, we use the experimental values of ΔEv/a instead of the calculated values in the CCSD method and refer to the approach as the CCSDex method. This is obviously better than the approach which improves the Ev/a values by incorporating contributions from the important triple excitations in a perturbative approach in the CCSD method (CCSD(T) method) that was employed in Refs. [16] and [37].

After obtaining amplitudes of the MBPT and RCC operators using the equations described above, the Θ values of the considered states are evaluated using the expression

This gives rise to a finite number of terms for the MBPT(2) and LCCSD methods, but it involves two non-terminating series in the numerator and the denominator, which are eTv/a Θ eTv/a and eTv/a eTv/a, respectively, in the CCSD method. We account contributions from these non-truncative series by adopting iterative procedures as described in our previous works.[38,41] We also give results considering only the linear terms of Eq. (22) that appear exactly in the LCCSD method, but using amplitudes of the RCC operators from the CCSD method (referred to as the CCSD(2) method). Therefore from the differences between the results of the LCCSD and CCSD methods, one can infer the importance of the non-linear terms in the calculations of the wave functions; while from the differences between the results of the CCSD(2) and CCSD methods, one can understand the roles of the non-linear effects appearing in Eq. (22) for the estimations of the Θ values. We also give results considering a maximum of four T operators in the above non-truncative series as the CCSD(4) method.

We present in Table 1 the Θ values of the 4f145d 5D3/2 and 4f136s2 2F7/2 states of Yb+ obtained from various methods described above. We also compare our results with the other calculation results and available experimental values. We consider the results from the CCSDex method as our recommended values as this method accounts for the more physical effects. We have also estimated uncertainties to the CCSDex results by estimating the neglected contributions due to truncation in the basis functions and from the omitted correlation effects that could mainly arise through the triply excited configurations. We had also presented these values using the CCSD(4) method in our previous work.[16] Differences in the results from the CCSD(4) and the CCSD methods indicate that the higher non-linear terms are insignificant for the 4f145d 2D3/2 state, but they contribute slightly to the 4f136s2 2F7/2 state. As said before, these series are solved iteratively to include infinity numbers of terms in this work. Again, we have removed uncertainties due to the calculated energies that enter into the amplitude solving Eq. (18) of the CCSD method by using the experimental energies.

Table 1.

Demonstration of trends in the Θ values (in a.u.) from lower to higher order methods. Our results are compared with the values available from other calculations and the experimental results. The estimated final values are shown in bold fonts. Uncertainties to these values are also quoted separately.

.
4. Ion trap induced shift

We plan to employ a modified Paul trap[42] of end cap geometry as shown in Fig. 1(a). In reality, such traps are not capable of producing pure quadrupole potential Φ(2) due to the geometric modifications of the hyperbolic electrode, machining inaccuracies, and misalignments. On the other hand, for precision measurements with ions stored in a non-ideal trap, the anharmonic components of the potential Φ(k>2)(x,y,z) are non-negligible due to the fact that they change the ion dynamics and also affect the systematics. For minimizing such effects several groups, such as MPQ Germany,[43] NRC Canada,[11] PTB Germany,[44] and European Metrology Research Programme on ion clocks,[45] have come up with different end cap trap designs for establishing single ion frequency standards. Here we aim to identify a new end cap trap geometry in which the trap induced quadrupole shift can be minimized. This trap can also add minimum anharmonicity to the confining potential and small micromotions.[46] In a cylindrically symmetric trap as shown in Fig. 1(a), only the even order multipoles contribute. Here, in order to estimate the quality of the trap potentials, we consider k up to 10 since the amplitudes of Φ(k) fall drastically at higher k. The tensor components of ∇E for each multipole potential Φ(k) are opted from their electric field components Ex,y,z. The corresponding fractional quadrupole shift ΔνQ/ν0 at each k is estimated from Eq. (2). The variation of ΔνQ/ν0 for all multipole potentials up to k = 10 at two different distances from the trap center are estimated for the E2 and E3 clock transitions, which are shown in Fig. 2. The reported experimental values of the quadrupole shifts for these two transitions are also depicted in the same figure for comparison.

Fig. 1. (a) Our proposed electrode assembly of the end-cap trap with 2z0 = 0.7 mm, 2z2 ≈ 1.0 mm, 2r1 = 1 mm, 2r2 = 1.4 mm, 2r3 = 2 mm, θi = 10°, and θo = 45°. (b) Variation of the trap depth as a function of dc voltage U for a fixed radio-frequency ωrf and an ac voltage V.
Fig. 2. Fractional electric quadrupole shifts due to the multipoles up to k = 10 for the (a) E2 and (b) E3 transitions. The ion is off centered by a distance of 0.1 μm (green) and 10 μm (black). The harmonic potential gives a spatially independent |ΔνQ/νo|. The reported fractional accuracies lie within the grey bands and so far the best accuracies are 1.1 × 10−16[8] and 5 × 10−17[18] for the E2 and E3 transitions, respectively.

The dominating perturbation of Φ(2) arises from the octupole term Φ(4), which may largely affect the quadrupole shift of the trapped ion frequency standards due to the wrong choice of electrode geometry. As an example, we simplify the analysis considering Φ(x,y,z) ≃ Φ(2) + Φ(4) since other higher orders are less significant for the quadrupole shift (Fig. 2). In the absence of any asymmetries, the trap potential can be written as

where and VT(t) = U + V cos(ωrft), which depends on the dc and rf components of the trapping voltages with amplitudes U and V, respectively. The dimensionless coefficients c2 and c4 depend on the electrode geometry. Here we consider |x| = |y| = r, since the trap is axially symmetric. The harmonic part of the potential can produce the restoring force on the ion and the resultant axial trap depth yields [47] where and Q and m are the charge and the mass of the ion, respectively. As an example, in Fig. 1(b), we show the variation of Dz with U for fixed values V = 500 V and ωrf = 2π × 12 MHz. Since the first order quadrupole shift from the rf averages to zero and its second order is also zero for 171Yb+,[28] we estimate the shift considering U = 10 V. The quadrupole shift as given by ∑qEqD0q results in due to the harmonic part of the potential and it is constant within the trapping volume. The spatial dependency comes from the higher orders, for example, Φ(4) results in a quadrupole shift of

The strengths of the multipole potentials depend on ck as given in Eq. (23). The magnitudes of ck depend on the geometric parameters of the trap electrodes, such as radius ro, angle θi of the electrode carrying rf (that is the inner electrode), inside and outside radii r1 and r2, angle θo of the dc carrying electrode (that is the outer electrode) which is coaxial to the inner one and also their mutual tip-to-tip separations 2z0 and 2z1, respectively, as shown in Fig. 1(a). We have obtained the trap potentials for various choices of these geometric factors by carrying out numerical simulations using the boundary element method.[48] Then, we have characterized the multipole components in it by fitting ∑k Φ(k) for k up to 10. The potentials are obtained for various combinations of θi, θo, 2r0, and 2z0 but at the fixed values of 2r1 = 1.4 mm, 2r2 = 2 mm, and 2z1 = 1.16 mm. We fix these parameters keeping in mind that the laser beams from the three orthogonal directions can impinge on the ion without any blockage as described in Ref. [31]. These three laser beams will be used for detecting the micromotions in all the three directions independently.[49,50] After studying a series of trap geometries, we find that the diameters of the outer electrode have a weak influence on Φ(k) which is below the accuracy that is expected from the machining tolerances. In Fig. 3, the variation of the quadrupole shift at the center of the trap with the diameter and the tip-to-tip separation of the inner electrode is shown, keeping the θi and θo values fixed. When placing the inner electrodes further away from each other, it is necessary to operate the trap at a larger voltage to obtain the required trap depth. On the other hand, placing them close to each other or having larger diameters can introduce optical blockage for the three orthogonal laser beams. Also micromotions of the ions can increase at large 2ro and 2z0.[46] We obtain the optimized values as 2r0 = 1 mm and 2z0 = 0.7 mm for which ΔνQ is reduced but not minimized. Further attempt to minimize the quadrupole shift causes increase in anharmonicity and micromotions. The dependences of the quadrupole shift on θi and θo are shown in Figs. 3(e)3(f). These clearly show that the quadrupole shift increases at large θi but it has a relatively weak influence on θo. A pair of inner electrodes with flat surfaces will introduce minimum shift. However θi = 0 gives optical blockage at our optimized 2z0 = 0.7 mm for impinging three orthogonal Gaussian laser beams of waist ∼ 30 μm on the ion and overlapping them with the other laser beams. To avoid the optical blockage, we have optimized the values of θi and θo at 10° and 45°, respectively. These would produce an insignificant number of scattered photons from the tails of the Gaussian laser beams, which will propagate along the three mutually orthogonal directions in our described design reported in Ref. [31]. For our trap geometry, the coefficients c2/2R2 and c4/2R4 are estimated to be 0.93 × 106 m−2 and 0.11 × 1012 m−4, respectively.

Fig. 3. Fractional electric quadrupole shifts for the E2 and E3 clock transitions corresponding to the different geometrical parameters: (a), (b) diameter 2ro of the inner electrode; (c), (d) tip-to-tip separation 2z0 of the inner electrodes; and (e), (f) angles θo at which the outer electrodes are machined. Each figure consists of a set of plots for angles of the inner electrode θi, which are 0° (black), 10° (red), 20° (gray), 30° (green), 40° (cyan), 50° (blue), and 60° (purple). The dashed line connects data points for the fixed values of θi and 2ro = 1 mm and 2z0 = 0.7 mm for all of them.

Due to the residual thermal motion of the ion after the laser cooling, it is unlikely to probe the clock transition while the ion is sitting at the center of the trap. Also there is a possibility of shift in the mean position of the ion from the trap center due to imperfect stray field compensation. This results in a spatially dependent ΔνQ in a non-ideal ion trap. At any position, the shift due to potentials for k > 2 increases following a power law of order k – 2 to the separation of the ion from the trap center, as shown in Fig. 2. Figure 4 shows the spatial variation of ΔνQ/ν0 due to Φ(4) and compares that with the shift resulting from Φ(2). These two figures clearly show that the quadrupole shift due to the dominating anharmonic potential is insignificant for a frequency standard of accuracy ∼ 10−18 when the ion is positioned in the vicinity of the trap center within sub-micron precision and at a tight confinement. The quadrupole shift due to Φ(2) can be eliminated by measuring the clock frequency along the three mutually orthogonal orientations in the lab frame and averaging them while quantizing the ion using the magnetic fields of equal amplitudes.[34,49] Such angular averaging also eliminates the quadrupole shift due to Φ(4) provided that the trap is perfectly axially symmetric, which is shown in Fig. 4. In practice, the trap can deviate from such an ideal situation, which could lead to inaccuracy in eliminating the quadrupole shift by angular averaging. In such a non ideal trap, the inaccuracies of eliminating the quadrupole shift are generally induced by Φ(4) but it is below the accuracy of the frequency standard that we are aiming for. The present analysis helps us in building a suitable trap electrode where the effect of Φ(2) in the quadrupole shift is reduced.

Fig. 4. Spatial dependence of the fractional electric quadrupole shifts ΔνQ/νo for the (a) E2 and (b) E3 clock transitions of 171Yb+. The quadrupole trapping potential produces a constant shift (gray). The spatial dependency in the shift along the radial (red) and axial (blue) directions arise from the anharmonic components k > 2 of Φ(k).

The formula for evaluating the quadrupole shift is given by

Substituting values of Θs from Table 1, we estimate these shifts for a constant c2U/R2 = 932 V/cm2 which is expected in our trap geometry at U = 10 V. All the resultant quadrupole shifts for the E2 and E3 clock transitions are shown in Figs. 5(a) and 5(b), respectively. This shows that in our trap, we will be able to measure the quadrupole moment of the 5d 2D3/2(F = 2) state at an accuracy of 1 part in 103. This uncertainty will be one order of magnitude better than the previous measurement.[28] Similarly, the quadrupole moment of the 4f136s2 2F7/2(F = 3) state was previously measured with 12% accuracy. We also expect to improve the accuracy of this quantity using our proposed ion trap. This will help to verify the reported discrepancies among the experimental and theoretical results.

Fig. 5. The electric quadrupole shifts obtained in the previously reported works and estimated values from this work are shown in panel (a) for the 6s 2S1/2 → 5d 2D3/2 and panel (b) for the 6s 2S1/2 → 4f136s2 2F7/2 clock transitions.
5. Conclusion

We have proposed a suitable ion trap geometry for carrying out accurate measurements of the quadrupole shifts of the 6s 2S1/2 → 5d 2D3/2 and 6s 2S1/2 → 4f136s2 2F7/2 clock transitions in 171Yb+. We have also calculated the Θ values of the 5d 2D3/2 and 4f136s2 2F7/2 states of 171Yb+ using the RCC method that is used in our analysis. We have identified an end cap ion trap geometry which can produce nearly an ideal harmonic confinement to minimize the electric quadrupole shift. We also showed that the anharmonic component of the potential adds an insignificant quadrupole shift for a frequency standard with an accuracy ∼ 10−18, when the ion is positioned at the trap center within sub-micron precision. To obtain a figure of merit of our design, we have estimated the quadrupole shifts along with their uncertainties in our proposed setup and compared the result with the previously available values. In our optimized ion trap, we expect to measure the quadrupole moment of the 5d 2D3/2 state with an accuracy of one part in 103. We are also aiming to measure the quadrupole moment of the 4f136s2 2F7/2 state reliably that could possibly explain the discrepancy between the experimental and theoretical results.

Reference
1BauchATelleH R 2002 Rep. Prog. Phys. 65 789
2PoliNOatesC WGillPTinoG M 2013 Riv. Nuovo Cimento 36 555
3TakamotoMUshijimaIDasMNemitzNOhkuboTYamanakaKOhmaeNTakanoTAkatsukaTYamaguchiAKatoriH 2015 C. R. Phys. 16 489
4RosenbandT 2008 Science 319 1808
5GillPBarwoodG PKleinH AHuangGWebsterS ABlytheP JHosakaKLeaS NMargolisH S 2003 IEEE Meas. Sci. Technol. 14 1174
6GodunR M 2014 Phys. Rev. Lett. 113 210801
7TammCWeyersSLipphardtBPeikE 2009 Phys. Rev. 80 043403
8HuntemannNOkhapkinMLipphardtBWeyersSTammCPeikE 2012 Phys. Rev. Lett. 108 090801
9HuntemannNLipphardtBTammCGerginovVWeyersSPeikE 2014 Phys. Rev. Lett. 113 210802
10WangY HDumkeRLiuTStejskalAZhaoY NZhangJLuZ HWangL JBeckerTWaltherH O 2007 Opt. Comm. 273 526
11DubéPMadejA AZhouZBernardJ E 2013 Phys. Rev. 87 023806
12BarwoodG PHuangGKleinH AJohnsonL A MKingS AMargolisH SSzymaniecKGillP 2014 Phys. Rev. 89 050501
13HuangYLiuPBianWGuanHGaoK 2014 Appl. Phys. 114 189
14KajitaMLiYMatsubaraKHayasakaKHosokawaM 2005 Phys. Rev. 72 043404
15ChouC WHumeD BKoelemeijJ C JWinelandD JRosenbandT 2010 Phys. Rev. Lett. 104 070802
16NandyD KSahooB K 2014 Phys. Rev. 90 050503
17WebsterS AGodunR MKingS AHuangGWaltonBTsatourianVMargolisH SLeaSGillP 2010 IEEE Trans. on Ultrasonics, Ferroelectrics, and Frequency Control 57 3
18TammCHuntemannNLipphardtBGerginovVNemitzNKazdaMWeyersSPeikE 2014 Phys. Rev. 89 023820
19RobertsMTaylorPGateva-KostovaS VClarkeR B MRowleyW R CGillP 1999 Phys. Rev. 60 2867
20KingS AGodunR MWebsterS AMargolisH SJohnsonL A MSzymaniecKBairdP E GGillP 2012 New J. Phys. 14 013045
21HuntemannNSannerCLipphardtBTammCPeikE 2016 Phys. Rev. Lett. 116 063001
22DzubaV AFlambaumV V 2011 Phys. Rev. 83 052513
23SahooB KDasB P 2011 Phys. Rev. 84 010502
24DzubaV AFlambaumV VSafronovaM SPorsevS GPruttivarasinTHohenseeM AHäffnerH 2016 Nat. Phys. 12 465
25DzubaV AFlambaumV VMarchenkoM V 2003 Phys. Rev. 68 022506
26DzubaV AFlambaumV V 2008 Phys. Rev. 77 012515
27Nisbet-JonesP B RKingS AJonesJ MGodunR MBaynhamC F ABongsKDolez̆alMBallingPGillP 2016 Appl. Phys. 122 57
28SchneiderTPeikETammC 2005 Phys. Rev. Lett. 94 230801
29BlytheP JWebsterS AHosakaKGillP 2003 J. Phys. 36 981
30PorsevS GSafronovaM SKozlovM G 2012 Phys. Rev. 86 022504
31RastogiABatraNRoyAThangjamJKalsiV P SPanjaSDeS 2015 MAPAN J. Metrology Society of India 30 169
32DeSBatraNChakraborySPanjaSSenguptaA2014Current Science1061348
33RamseyN F1956Molecular BeamsLondonOxford Univ. Press10.1093/acprof:oso/9780198520214.001.0001
34ItanoW M 2000 J. Res. Nat. Inst. Stand. Tech., USA 105 829
35EdmondsA R1974Angular Momentum in Quantum MechanicsNew JerseyPrinceton Univ. Press
36LindgrenIMorrisonJ1986Atomic Many-Body Theory2nd Edn.BerlinSpringer-Verlag10.1007/978-3-642-61640-2
37NandyD KSahooB K 2013 Phys. Rev. 88 052512
38SahooB KNandyD KDasB PSakemiYPhys. Rev. A91042507
39ItanoW M 2006 Phys. Rev. 73 022510
40LathaK V P 2007 Phys. Rev. 76 062508
41SinghYSahooB K 2015 Phys. Rev. 91 030501
42PaulW 1990 Rev. Mod. Phys. 62 531
43SchramaC APeikESmithW WWaltherH 1993 Opt. Commun. 101 32
44SteinB2010Ph. D. thesisHannover: University of Hannover
45Dolez̆alMBallingPNisbet-JonesP B RKingS AJonesJ MKleinH AGillPLindvallTWallinA EMerimaaMTammCSannerCHuntemannNScharnhorstNLerouxI DSchmidtP OBurgermeisterTMehlstäublerT EPeikE2015arXiv:1510.05556
46BatraNDeSPanjaS2016The ion trap geometry dependent dynamics have been studied[Manuscript in preparation]
47MajorF GGheorgheV NWerthG2010Charged Particle TrapsNew YorkSpringer10.1007/b137836
483D Charged Particle Optics Program (CPO-3D) CPO Ltd., USA
49BerkelandD JMillerJ DBergquistJ CItanoW MWinelandD J 1998 J. Appl. Phys. 83 10
50KellerJPartnerH LBurgermeisterTMehlstäublerT E 2015 J. Appl. Phys. 118 104501