Measuring the Using Circular Rydberg in an Intensity-Modulated Optical Lattice

Andira Ramos1,∗ Kaitlin Moore2, and Georg Raithel1,2 1Department of Physics, University of Michigan, Ann Arbor, Michigan 48105, USA and 2Applied Physics Program, University of Michigan, Ann Arbor, Michigan 48105, USA (Dated: May 9, 2017)

A method for performing a precision measurement of the Rydberg constant, R∞, using cold circular Rydberg atoms is proposed. These states have long lifetimes, as well as negligible quantum- electrodynamics (QED) and no nuclear-overlap corrections. Due to these advantages, the measure- ment can help solve the “ radius puzzle” [Bernauer, Pohl, Sci. Am. 310, 32 (2014)]. The atoms are trapped using a Rydberg- optical lattice, and transitions are driven using a recently- demonstrated lattice-modulation technique to perform Doppler-free spectroscopy. The circular-state transition frequency yields R∞. Laser and beam geometries are selected such that the lattice-induced transition shift is minimized. The selected transitions have no first-order Zeeman and Stark corrections, leaving only manageable second-order Zeeman and Stark shifts. For Rb, the projected relative uncertainty of R∞ in a measurement under the presence of the Earth’s gravity is 10−11, with the main contribution coming from the residual lattice shift. This could be reduced in a future micro-gravity implementation. The next-important systematic arises from the Rb+ (relative-uncertainty contribution of ≈ 3 × 10−12).

PACS numbers: 06.20.Jr, 32.80.Rm, 42.62.Fi

I. INTRODUCTION

1 0 0 0 | n = 5 3 , n = 1 , n = 1 , m = 5 0 Knowing the value of the Rydberg constant (R∞) accu- 1 2 l rately has been of interest for decades due to its relation 5 0 0 - ) 0

to other fundamental constants and its role in calcula- a ( 0 + tions of atomic energy levels. More recently, the large z R b + + discrepancies in the proton [1, 2] and deuteron [3] radii - 5 0 0 that were found using muonic and deuterium, - ∆ν= 9 3 . 6 5 7 G H z respectively, have reinforced the need to confirm the ac-

curacy of R∞. Previous precision experiments with this | n = 5 1 , n 1 = 0 , n 2 = 0 , m l = 5 0 goal have involved low-lying states of hydrogen, limited 5 0 0 ) 0

typically by statistical uncertainties, AC Stark shifts and a ( 0 + second-order Doppler shifts [4]. These have led to the z R b + current CODATA relative uncertainty for the Rydberg 0 constant value of 5.9 × 10−12 [5]. There has also been a - 5 0 0

study involving circular Rydberg states of hydrogen [6] - 1 0 0 0 −11 (relative uncertainty of 2.1 × 10 ) and a proposal in- 0 1 0 0 0 2 0 0 0 3 0 0 0 4 0 0 0 ρ volving circular states of [7] (expected relative ( a 0 ) uncertainty of about 10−10). The approaches involving low-lying states and circular states deal with significantly different frequency regimes: optical versus microwave. FIG. 1. Normalized probability density, |ψ|2, for the circular Therefore, measurements involving low-lying states of hy- state (bottom) and near-circular state (top) of the transition drogen can have a better relative uncertainty δν/ν than of interest in the proposed experiment. The kets are labeled in principal, parabolic and magnetic quantum numbers, n, n1, results for circular states (under the assumption of sim- + 2 ilar absolute uncertainty, δν). However, circular states n2, m [9]. Near the Rb core, |ψ| = 0 for both states. The arXiv:1705.02682v1 [physics.atom-ph] 7 May 2017 signs refer to the polarity of ψ. It is seen that the transition are insensitive to several systematics that are limiting requires a quadrupolar interaction. in spectroscopy of low-lying states, as discussed in this paper. It has also been proposed to measure the Ryd- trapped circular Rydberg atoms. Transitions are driven berg constant using high- states of using a recently-developed spectroscopic method, which hydrogen-like [8]. is based on lattice modulation at microwave frequencies We propose an experiment to obtain an indepen- [10]. The critical advantages of circular states, in com- dent measurement of the Rydberg constant using cold, parison with low-lying states of hydrogen, are their long radiative lifetimes (on the order of ms) [11], their small QED corrections [12], and the absence of any overlap ∗ [email protected] with the nucleus, hence eliminating nuclear charge dis- 2 tribution effects (see Fig. 1). a) nF c) F, B In contrast with previous efforts to measure R [4, 13], 7/2 ∞ 5D it is proposed to trap cold Rydberg atoms using a pon- 5/2 5P deromotive potential optical lattice (POL) [14] instead of 3/2 F, B using cold atomic beams. Trapping the atoms allows for 5S1/2 increased interaction times. The light shift introduced by the lattice trap can be addressed by the use of “magic” b) conditions under which the trap-induced shifts of upper |53, 1, 1, 50 and lower states cancel [15]. Residual imperfections in ∆ν =93.657 GHz α z the magic-lattice trap can be addressed by performing 3 |51, 0, 0, 50 the experiment under micro-gravity conditions, which al- low for an overall reduction in the magic-lattice shift. A key method is ponderomotive spectroscopy [10], FIG. 2. (color online). a) Excitation scheme used to pre- which allows for circular-to-near-circular transitions to pare atoms before circularization. b) After circularization, be driven by amplitude-modulating the ponderomotive the transition between an initial, circular state and a final, optical lattice. The method permits us to drive the near-circular state (expressed using parabolic quantum num- bers |n, n , n , m i) is driven by ponderomotive spectroscopy, quadrupolar transition needed in our work as a first-order 1 2 l indicated by the curly arrow. c) Parallel stabilization fields, process and to eliminate Doppler broadening. By driving F and B, define the axis (z-axis) of the atoms. transitions between states of the same magnetic quantum The angles between different pairs of counter-propagating lat- number, we eliminate the first-order Zeeman effect. To tice beams, αi, can be varied to change the periodicities of the obtain a zero first-order Stark shift, we select states with lattice in the ith direction. parabolic quantum numbers n1 = n2. In these ways, the proposed experiment addresses im- In order to take advantage of the circular atoms’ long portant issues that have been encountered in preceding lifetimes (tens of ms at 4 K), it is necessary to trap the Rydberg constant measurements. In addition, due to Rydberg atoms. To achieve this, a three-dimensional the elimination of nuclear charge overlap, the measure- standing-wave optical lattice is adiabatically turned on, ment of R∞ is also independent of the radius of the pro- and the atoms are trapped via the ponderomotive po- −11 ton. Overall, a relative uncertainty of 10 is expected, tential [22] (see Fig. 2c). This ponderomotive poten- which would already be enough to shed light onto the tial emerges from the last term in the minimal-coupling proton radius puzzle. An improvement beyond the cur- Hamiltonian (in SI units), rent CODATA uncertainty, δν/ν = 5.9 × 10−12 [5], is possible by an implementation of the experiment under ˆ 1 2 micro-gravity conditions. H = (2|e|A(ˆr) · ˆp + e A(ˆr) · A(ˆr)), (1) 2me

which is proportional to laser intensity and arises when II. PROPOSED EXPERIMENTAL OUTLINE a quasi-free Rydberg is placed in a rapidly oscil- lating field. Here, A(ˆr) is the vector potential of the op- A. Atom Preparation and Spectroscopy tical field, ˆp is the momentum operator, ˆr is the position operator, me is the electron mass and e is the electron We use cold atoms to reduce interaction-time broad- charge. For non-degenerate states in a monochromatic ening and to limit the interaction volume, thereby reduc- lattice, the adiabatic trapping potential is given by ing field inhomogeneity effects. We propose to use 85Rb 2 Z atoms, pre-cool them in a MOT to ∼ 100 µK, and fur- e 2 2 3 Vad(R) = |ψ(r)| |A(r + R)| d r, (2) ther cool them to ∼ 10 µK and ∼ 1 µK in bright and 2me gray optical molasses, respectively [16, 17]. Since the atom densities desired for this experiment are moderate where R is the atomic center-of-mass position. < 8 −3 (∼ 10 cm ), to avoid Rydberg-Rydberg interactions, The of the lattice is chosen to match a the use of optical molasses is ideal. “magic” condition for the desired transition (equal po- In order to circularize atoms, we use a modified adia- tentials for lower and upper states). To illustrate this, batic rapid passage (ARP) method [18, 19], which in Rb in Fig. 3a we show adiabatic lattice-potential depths ob- requires that the Rydberg atoms are initially prepared tained from Eq. 2 for the two states of interest, n=51 in ml=3. We employ a three-level excitation scheme, and n=53, as a function of wavelength. In the figure it 5S1/2 → 5P3/2 (wavelength of 780 nm), 5P3/2 → 5D5/2 is seen that for the two states considered there are two (776 nm), and 5D5/2 → nF7/2 (1260 nm) (see Fig. 2a). options to achieve magic conditions, 290 nm and 532 nm. The modified ARP circularization method is optimal for For the proposed experiment we choose a 532-nm lattice. < states n ∼ 50 (our case). If much higher principal quan- The same ponderomotive term that traps atoms is tum numbers were desired, the “E×B” method [20, 21] also used to drive transitions between circular and near- could be used. circular states [10]. In ponderomotive spectroscopy, the 3 optical-field intensity varies substantially within the vol- units). This stabilization method is suitable for high- ume of the atom, and the lattice amplitude modulation precision spectroscopy [6, 22]. frequency is resonant with an atomic transition or a sub- In order to probe the atoms, we trap them in a pon- harmonic of it [23]. In the proposed experiment, the spa- deromotive potential. The ponderomotive shift must be tial variation of the lattice is about the same as the size smaller than the shift caused by the stabilizing electric of the atom, and the lattice-amplitude modulation fre- and magnetic fields [22]. The potential must also be deep quency (atomic transition frequency) is about 100 GHz. enough to trap atoms at their thermal . Finally, the population transfer between initial and fi- Hence, the laser-cooling temperature sets minimum val- nal states is measured using state-selective field ioniza- ues for the fields that we use in both trapping and sta- tion [11]. bilization of our states. The hierarchy of shifts is shown The proposed experiment occurs inside a cryogenic en- in Table I for MOT temperatures (∼ 100µK), tempera- closure that provides a radiation temperature near 4 K. tures in gray optical molasses (∼ 1µK) [16, 17] and Bose- This is done to decrease blackbody radiation effects, Einstein condensate (BEC) temperatures (∼ 10 nK). Ta- which reduce the Rydberg-state lifetimes [11] but cause ble II shows the corresponding typical field magnitudes only minor shifts of the transition frequencies of interest and provides guidance in designing the circular-state sta- (see section III.G). bilization scheme.

TABLE I. Hierarchy of level shifts in three atomic tempera- B. Stabilization Fields ture regimes.a T(µK) Thermal Energy b POL Magnetic Electric The <2n2-fold degeneracy of the circular and hy- ∼ 100 1000 3100 9400 28000 drogenic states must be lifted using stabilization fields. To stabilize the circular Rydberg states against state- 1 10 31 94 280 mixing, we employ an electric field, F , and a weak, par- 0.01 0.1 0.31 0.94 2.8 allel magnetic field, B (B/2 < (3/2)nF < n−3, in atomic a All energies are expressed in kHz. b Thermal Energy = kBT/2. ) p a ) V | n = 5 1 , n 1 = 0 , n 2 = 0 , m = 5 0

f 0 . 4

o | n = 5 3 , n = 1 , n = 1 , m = 5 0

1 2

s TABLE II. Magnetic and electric fields suitable for three t i

n ΙΙΙ ΙΙ Ι temperature regimes. The fields satisfy optical-trap depth u (

≈ kBT < B/2 < (3/2)nF , for n=51.

h 0 . 2 t

p T(µK) (mT) (mV/cm) e D

100 0.67 290 l a

i −3 t 0 . 0 1 6.7 × 10 2.9 n

e b ) −5 t 0.01 6.7 × 10 0.029 o P

c Ι i - 0 . 2 t

a Since we are always in the Paschen-Back regime of the

b ΙΙ a

i fine structure (see sections III B and E), we use the basis

d ΙΙΙ

A - 0 . 4 set {|n, n1, n2, ml, msi} [9] throughout this paper, where 2 0 0 4 0 0 6 0 0 8 0 0 1 0 0 0 n1 and n2 are parabolic quantum numbers. W a v e l e n g t h ( n m ) Parabolic and spherical bases are related by [11] X FIG. 3. a) Ponderomotive adiabatic potential depth (in units |n, n , n , m i = Cn1,n2 |n, l, m i, (3) 1 2 l l,ml l of the free electron potential, Vp, the energy of a free electron l in the lattice laser field) as a function of wavelength for the two states of interest in a one-dimensional lattice formed by where the Clebsch-Gordan coefficients are related counter-propagating beams (α = 0). The points at which the to the Wigner 3J symbols by [11] Cn1n2 = lml two plots cross are the “magic” wavelengths for this particu- hn, n , n , m |n, l, m i, lar pair of states. The magic wavelength we choose for this 1 2 l l experiment is shown with a white dashed line, and it occurs √ Cn1n2 = (−1)(1−n+ml+n1−n2)/2+l 2l + 1 at about 532 nm. b) Schematics of the projection of the wave- lml function density onto the lattice, as the wavelength is varied ! n−1 n−1 l (4) and the atom size remains fixed. Labels I, II and III corre- × 2 2 . ml+n1−n2 ml−n1+n2 spond to those in part a). The oscillatory behavior and flip in 2 2 −ml signs in a) are related to how many lattice periods fall within the volume of the atom [14]. Note that for the states in Figs. 1, 2 and 3, the sum over l has at most two non-zero terms. 4

III. ENERGY SHIFTS

In this section we discuss the various energy-level shifts; the results are summarized in the discussion (sec- tion IV).

A. Lattice-Induced Shift

In its interaction with the optical-lattice field, the Rydberg electron behaves as a quasi-free particle. For a plane-wave linearly-polarized field of the form xˆFL(r) sin(ωt)(xˆ is a unit vector), Eq. 1 leads to the free-electron ponderomotive potential

2 2 e |FL(r)| Vp = 2 , (5) 4meω FIG. 4. Effects of wavefunction projections on the depth where ω is the angular frequency of the laser electric field of the BO adiabatic potential for a 532-nm lattice extending 2 9 2 and |FL(r)| is proportional to the spatially-varying field along x and z, with single-beam intensities 4 × 10 W/m intensity [24]. The ponderomotive potential is the aver- (αi = 0). a) Alignment of the optical-lattice standing-waves age kinetic energy of the free electron in the lattice laser and the circular-state probability distribution. The ampli- field, and it is polarization- and phase-independent. tude of the z-direction lattice is modulated in time. b) Pro- 2 The position-dependent ponderomotive potential, jections of |ψ| along x and z. The overlap of the projections with the optical-lattice standing waves determines the BO Vp(ˆr + R), is added as a perturbation to the Rydberg electron’s Hamiltonian. Diagonalization of the Rydberg adiabatic trapping potentials along the respective coordinate directions (see Eq. 6). c) Trapping potentials (as a function Hamiltonian yields the Born-Oppenheimer (BO) adia- of the center-of-mass position of the atom) calculated from batic potential surfaces, Vp(R), for the atom’s center- Eq. 6; the zero position corresponds to a lattice field node. of-mass motion, as well as the associated adiabatic The different depths and phases are a result of the quite dis- Rydberg-electron wavefunction, ψ(r; R) [22]. tinct wavefunction projections onto x and z. Generally, ψ(r; R) is unknown and must be simulta- neously solved for along with the BO potentials [25]. In our regime, where the shifts due to the parallel electric The ratio of laser intensities of the lattice axes (up and magnetic stabilization fields dominate the optical to three) and the aspect ratio between the atom’s size shifts, the adiabatic states are given by the parabolic ba- (defined by its known state) and the optical lattice peri- sis states, |n, n1, n2, ml, msi. This greatly simplifies the odicities (defined by λi and αi) (Fig. 3 and Fig. 4) can calculation of the BO adiabatic potential because ψ(r; R) be experimentally controlled. This allows us to vary the is no longer dependent on R. Since in our case, the op- depth and the minimum potential value of Vad(R) (see tical lattice is formed by three sets of lattice beams (see Fig. 3) and to realize a “magic”-lattice condition (where Fig. 2), the three-dimensional BO adiabatic potential fol- the two states in the transition experience the same en- lows from ergy shift in the BO potential Vad(R)). Experimental considerations suggest to use α = 0 and Z 2 2 i X e |FLi cos(∆ki · (R + r))| V (R) = d3r to choose common laser wavelengths λi, leaving the in- ad m ω2 i e i (6) tensity ratios and the atomic quantum numbers to at- tain a magic condition. Here we consider the n = 51 → × |ψ (r)|2. n,n1,n2 n = 53 transition, which, as seen in Fig. 3, has a magic wavelength at about 532 nm (the second harmonic of In the integral in Eq. 6, |ψ (r)|2 acts as a weight- n,n1,n2 a ND:YAG laser). Using a magic lattice, most of the ing factor. There, i is the summing index over optical- lattice-induced shift in the proposed experiment can be lattice directions (for a 3D lattice, i = 1, 2, 3), which need eliminated (see section IV.A). not be orthogonal to each other; FLi is the field ampli- tude of a single beam; ψn,n1,n2 (r) is the R-independent Rydberg electron wavefunction; r is the B. First-Order Zeeman and Stark Shifts (relative) position, ωi is the angular frequency of the lat- tice beam; and |∆ki| = |ki1 − ki2| = 2kicos(αi/2), where ki1 and ki2 are the wavevectors corresponding to the pair To avoid undesired state-mixing due to minute stray th of lattice beams along the i direction, and αi is the electric and magnetic fields, stabilization fields must be angle between a pair of counter-propagating beams (see applied to the atoms. The most suitable stabilization Fig. 2c). scheme is one where (3/2)nF > B/2, in atomic units 5

(see section II.B). The Stark Hamiltonian for an electric wherex ˆ andy ˆ are the x and y-direction position opera- field, F, pointing along z is tors. This Hamiltonian can be rewritten in the spherical basis as HˆS = F ez,ˆ (7) wherez ˆ is the z-component of the position operator. In 2 2 ˆ e B 2 2 ˆ the basis of parabolic states, |n, n , n , m i, with quan- HD = (ˆr sin θ), (13) 1 2 l 8me tization axis along z, the first-order eigenvalues of the Stark Hamiltonian are where the operator θˆ is the angle with respect to the z- 3 axis. Using Eq. 13, we obtain a diamagnetic energy shift E = F ea n(n − n ). (8) of S 2 0 1 2 X e2B2 D E The states and transitions in this work are of the type E = |Cn1n2 |2 nlm |rˆ2 sin2 θˆ|nlm , (14) SZ lml l l 8me |n, 0, 0, n − 1i ↔ |n + 2, 1, 1, n − 1i; in this case the linear l Stark shifts for both levels, as well as for the transition, where the angular matrix elements are given in [9] and are identical zero. the radial matrix elements can be computed numerically. We apply a weak magnetic field, B, in the z-direction As shown in Table III, these second-order shifts lead to that removes the remaining degeneracies between states uncertainty contributions below the current uncertainty of interest and other Stark levels. This results in a Zee- goal. man Hamiltonian Be HˆZ = (gLLˆz + geSˆz), (9) 2me D. Quantum Defects where g is the electron spin g-factor, g is the electron’s e L In alkali atoms, polarization and penetration quantum orbital g-factor (gL = 1) and Lˆz and Sˆz are the orbital an- gular momentum and spin angular momentum operators, defects are introduced as corrections to the hydrogenic respectively. In the Paschen-Back regime, the parabolic eigenvalue [11, 26] states with spin, |n, n1, n2, ml, msi, are eigenstates of the 1 E = −hcR , (15) Zeeman Hamiltonian. This shift is given by Rb 2 (n − δl) B¯he where c is the speed of light, n is the principal quantum EZ = (ml + gems), (10) M + 2me number, RRb = R∞ (M is the mass of Rb ) and me+M where ¯h is the reduced Planck’s constant. Since m and δl is the . This δl can be expressed as the l sum of the polarization and penetration quantum defects, ms are equal for both states involved in the transition considered, the linear Zeeman shift of the transition is δl = δpol + δpen, which is commonly expanded by using zero. the Rydberg-Ritz formula [11]. It is critical that the angle between the electric and This δl decreases significantly with increasing l. In magnetic field be close to zero for this to hold, since any the proposed experiment, where transitions from circular departure from zero would introduce x- or y-components to near-circular states are driven, δpen = 0 because the of the fields, leading to the appearance of additional probability density of circular Rydberg states is zero in second-order shifts. These can be estimated using Eq. the ionic core region (Fig. 1). Core polarization, how- 2.15 in [6], which yields an upper limit for the allowed ever, must still be considered, with the shift due to the effective dipole polarizability, α0 , being the leading term, angular misalignment between the fields of about one de- d followed by an almost negligible shift due to the effective gree. 0 quadrupole polarizability, αq. The effective polarizabili- ties consist of the DC polarizability and a non-adiabatic C. Second-Order Stark and Diamagnetic Shifts correction. The polarization potential is given by [27] −e2 1 α0 1 α0  Vˆ = d + q + ... , (16) Second-order perturbation theory for the Stark effect pol 16π22 2 rˆ4 2 rˆ6 Hamiltonian yields an energy shift of [6, 11] 0 0 0 where the values of αd and αq are obtained from [28] and −4π a3F 2n4 E = 0 0 can be converted to SI units [29]. For l  n, δpol cor- SS 16 (11) responding to this potential is approximately (in atomic  2 2 2  × 17n − 3(n1 − n2) − 9ml + 19 , units) which is small (see Table III). The diamagnetic Hamilto- 3 α0   l2  δ ≈ d 1 − O nian is pol 4 l5 d n2 (17) 2 2 35 α0   l2  ˆ e B 2 2 q HD = (ˆx +y ˆ ), (12) + 9 1 − Oq 2 , 8me 16 l n 6 where the O(l2/n2) terms are corrections that can be ex- where α is the fine-structure constant. In the Paschen- actly resolved by using the analytically known expres- Back regime (our case) the fine-structure-induced correc- sions for r−4 and r−6 (see appendix of reference tion to the energy levels is [26]). Since in the proposed experiment high-angular- momentum states are employed, the exact analytic ex- 4 2 α mec X n1n2 2 mlms −4 −6 ESO = |C | . (20) pressions for r and r need to be used. 2 lml n3l(l + 1)(l + 1 ) The quadrupole term in Eq. 16 becomes negligible at l 2 large r values such as the ones found in circular Rydberg states. The dipole polarizability term needs to be care- The relativistic contribution to the fine-structure shift fully considered as it leads to corrections of the order of results from expanding the expression for the relativis- several kHz. tic kinetic energy of a particle. This yields a correction Hamiltonian of [34]

4 1. ˆ pˆ Hrel = − 3 2 . (21) 8me c The polarizabilities in the polarization quantum defect are not well known. The most recent experimental limits Following the same procedure presented in [34], in first- 0 0 are αd = 9.12 and αq = 14 (in atomic units) [28], which order perturbation theory the relativistic shift is are not consistent with previous theory work [30, 31] for 4 2 the dipole and quadrupole polarizabilities, respectively. α mec X n1n2 2 Erel = − |C | The current uncertainties in the experimental values of 2 lml polarizabilities are of order 10−3, which lead to a relative l (22) −12  1 3  uncertainty of the order of 10 in the proposed Ryd- × − . 3 1 4 berg constant measurement, making this one of the main n (l + 2 ) 4n sources of uncertainty. Putting both terms together we obtain the fine- structure energy shift 2. Non-adiabatic Effects 4 2 α mec X n1n2 2 EFS = − |C | The quantum defect theory discussed so far assumes 2n3 lml that the Rb+ response to the Rydberg electron’s field is l (23)    adiabatic. However, this is not necessarily the case. The −mlms 1 3 × 1 + 1 − . non-adiabaticity of the electron’s motion makes it neces- l(l + 1)(l + 2 ) (l + 2 ) 4n sary to redefine Vˆpol [32] and hence the polarizabilities as For our states of interest, the relativistic correction is q around 6 kHz while the spin- correction is around −e2 1 α yd 1 α y + α yd  ˆ d 0 q 0 d 1 200 Hz. Vpol = 2 2 4 + 6 + ... , (18) 16π 0 2 rˆ 2 rˆ

d d q where y0 , y1 , and y0 vary slowly with n and l. Compar- ing this expression to Eq. 16, we see that the corrected F. Quantum Electrodynamic Corrections and the adiabatic polarizabilities are related as follows: 0 d 0 q d αd = y0 αd, αq = y0αq + y1 αd [26]. Using Ref. [33] the Quantum electrodynamics (QED) introduces the self- corrected dipolar and quadrupolar polarizabilities can be energy and the vacuum polarization QED corrections, 85 calculated for Rb. The experimental values of the po- which together form the Lamb shift. For circular Ryd- larizability used for this work already include the non- berg states, a first-order account of QED corrections is adiabatic correction. sufficient; the result is

4 3 8Z α X n1n2 2 E. Fine Structure Correction ELamb = hcR∞ |C | 3πn3 lml l (24) For Rydberg atoms with large l, the form of the fine-  3 c  × L(n, l) + lj , structure shift is the same as for the . 8 2l + 1 It has two contributions, the relativistic mass correction and the spin-orbit coupling [34]. The spin-orbit Hamil- where tonian is ( −1 α¯h 1 (l + 1) for j = l + 1/2 Hˆ = Lˆ · Sˆ, (19) clj = (25) SO 2 3 −l−1 for j = l − 1/2 2me c rˆ 7

The Bethe logarithm is L(n, l) [35], which can be extrap- being considered since ∆ω is defined by the transition in olated for n ≥ 4, l ≥ 3 as question. Approximations for the limiting cases of Eq. 27 are " 3/2 3/2# 0.1623834 1  1  given in [11]. In our case, the transition frequency of L(n, l) = − 2l + 1 l n interest is about 100 GHz, which is on the order of the " !# (26) peak of the radiation spectrum at 4 K. As a result, in 1 1 l + 13/2 × 1 ± − . order to calculate the blackbody shift, Eq. 27 has to be 2 4 n explicitly evaluated. For treating parabolic states, Eq. 3 can be used. The first term in Eq. 24 corresponds to vacuum polar- ization and the second term to the self energy. The latter gives rise to the anomalous magnetic moment of the elec- H. Hyperfine Structure Correction tron. The electron’s g-factor is 2+α/π (to lowest order). This changes the electron’s magnetic moment, leading to The interaction of the nuclear magnetic moment with the second term in squared brackets in Eq. 24, which is the magnetic field caused by the valence electron gives equivalent to accounting for the lowest-order correction rise to the hyperfine-structure Hamiltonian [37] of the electron’s g-factor in Eqs. 19 and 20.

In Eq. 24, the self-energy term is typically two orders µ g e2 Lˆ · ˆI g of magnitude higher than the vacuum polarization term, Hˆ = 0 I − e Sˆ · ˆI HFS 4πm m 2r3 4r3 with the values of the circular state of interest being e p (29) 0.59 Hz and 1.1 mHz, respectively, and values for the g g δ(r)  + e 3(Sˆ · ˆr)(ˆI · ˆr) + e Sˆ · ˆI , near-circular state being 0.31 Hz and 0.89 mHz. These 4r3 3 r2 corrections are small and lead to a small transition energy shift due to the Lamb shift as shown in Table III. which acts on the space {|n, n1, n2, ml, ms, mii}, where mi is the nuclear magnetic quantum number. Above, µ0 is the permeability of free space, gI is the g-factor of the G. Blackbody Shift nucleus, mp is the mass of the proton, ˆI is the nuclear spin operator and δ(r) is the Dirac delta function. The Blackbody radiation has two effects on Rydberg atoms: last term in the Hamiltonian is a contact term, where the on-resonant portion affects their lifetime and the off- the energy depends on the wavefunction density at the resonant portion can cause an energy shift [36]. Since position of the nucleus, which is zero for the states we Rydberg-Rydberg transition frequencies are in the range consider in this experiment (Fig. 1). of thermal blackbody radiation, thermal transitions lead Noting that the hyperfine structure is in the Paschen- to a lifetime reduction. For instance, the lifetime of the Back regime and using first-order perturbation theory n = 50 circular state is reduced from 30 ms at 0 K to and the analytic expression given in [9] for the r−3 matrix 10 ms at 4 K. Since both states in the transition of in- elements, we obtain a hyperfine energy shift of terest (Fig. 1) are affected by this lifetime reduction, we n1n2 2 expect a linewidth in the range of 30 Hz, which is suffi- X |Clm | E = m l ciently narrow for the current purpose. HFS i 3 3 1 a0n (l + 1)(l + )l The blackbody shift caused by the off-resonant portion l 2 2 2 is potentially of greater concern. The blackbody energy µ0gI¯h e  gems (30) × ml − shift is given by 8πmemp 2  2 2  2 Z ∞ n0l0 2 2 2l + 2l − 2m − 1 e X |R | |Fb| ∆ω l Enl = nl dω , (27) × 1 − 3 , BBR ¯h 2(∆ω2 − ω2) b (2l + 3)(2l − 1) n0,l0 0 b which leads to a negligible energy shift (see Table III). where ωb is the angular frequency of the blackbody ra- 2 diation, |Fb| is the blackbody field-amplitude squared n0l0 per frequency unit, Rnl is the radial matrix element, I. Doppler Effect and ∆ω is the transition angular frequency difference be- tween the final and initial states being considered. The To drive transitions, one of the components of the field amplitude, |F |2, can be obtained using the spectral b optical lattice is amplitude-modulated (see Fig. 4; opti- energy density form of the Planck radiation law cal carrier angular frequency ω and modulation angular 3 frequency Ω). Ponderomotive spectroscopy involves the 2 2¯hωb |Fb| = , (28) inelastic scattering of two counter-propagating optical-  π2c3(ehω¯ b/kBT − 1) 0 lattice of angular-frequency difference Ω [38], where kB is Boltzmann’s constant and T is the temper- which is at the atomic transition frequency. In tra- ature. Eq. 27 has an implicit dependency on the states ditional Raman spectroscopy with counter-propagating 8

area under the central peaks in Fig. 5. For a substan- 0 . 5 tial number of atoms to be captured, molasses tempera- tures are required, as shown in Fig. 5. At 100 µK, only about 6% of the population is trapped, whereas 1 µK 0 . 4 yields 51% trapped population. When temperatures are lowered, the full width of the Doppler-broadened back-

e 0 . 3 ground signal is ≈ 4vω/πc, where the thermal velocity P p 2 0 k H z v = kBT/Matom (see Eq. 31). The gaps between the 0 . 0 0 3 central peak and the onset of the Doppler background reflect the fact that atoms within a range of velocities 0 . 0 0 2 3 0 0 k H z are trapped. The trapped atoms experience no Doppler 0 . 0 0 1 shift and generate the Fourier-limited feature at the cen-

0 . 0 0 0 ter of the spectrum. They essentially undergo recoil-free - 1 0 0 - 5 0 0 5 0 1 0 0 absorption within the lattice wells. A more complete dis- D e t u n i n g ( k H z ) cussion of this topic will be given in a future paper.

FIG. 5. Simulation of the excited-state population, Pe, as a function of detuning for two temperature regimes: 100 µK IV. DISCUSSION (red dashed line) and 1 µK (blue solid line) for a potential depth of 35 kHz (motivated by Table I) and an interaction time of 5 ms. The inset shows that the widths of the narrow In Table III, we summarize the sources of frequency features at the center are Fourier-limited. The arrows indi- shifts and their respective relative uncertainties for the cate the approximate half widths of the Doppler-broadened sample atomic transition |51, 0, 0, 50i → |53, 1, 1, 50i con- background signals. sidered in this paper. These lead to an expected relative uncertainty in the proposed measurement of the Rydberg −11 beams, the Doppler shift between the drive frequency constant in the low 10 range. In the following, we dis- experienced by the atoms, Ω0, and Ω would follow the cuss the leading sources of uncertainty and how these can expression be improved in order to attain a state-of-the-art uncer- tainty. In contrast with measurements performed with v Ω v2 low-lying states of hydrogen (from which the best cur- Ω0 − Ω ≈ −2ωkˆ · + , (31) c 2 c2 rent uncertainty of 5.9 × 10−12 is obtained), our mea- surement is independent of nuclear effects and therefore where v is the center-of-mass velocity of an atom and the could contribute to solving the proton radius puzzle [40]. unit vector kˆ marks the direction of propagation of the beams. While the second-order Doppler effect is entirely negligible at MOT temperatures, the first-order Doppler effect would lead to a shift of about 20 kHz (for temper- A. Main sources of uncertainty ature of 1 µK). In ponderomotive spectroscopy Eq. 31 does not apply. While it approximately accounts for the The main source of uncertainty on the proposed overall widths of the spectra (indicated by the arrows in ground-based experiment is the residual lattice-induced Fig. 5), it fails to describe the central, narrow peak ob- shift. The resulting uncertainty is 3.2 × 10−11; it is served in ponderomotive spectroscopy. There, the cool- mainly due to laser-intensity fluctuations. This uncer- ing and trapping of the atoms in a magic lattice [23], tainty value is obtained assuming a 1% lattice-intensity combined with the fact that the phase of the Rabi fre- uncertainty (better stabilities are likely possible). The quency is constant within a potential well, allow us to lattice shift results presented in Table III are obtained achieve Doppler-free, Fourier-limited line-widths of the through simulations based on Eq. 6 where we specify the central peak (see Fig. 5). two atomic states of interest, arbitrary laser-beam geom- In order to model the spectra, we employ a simula- etry, wavelengths and intensities of the beams that form tion program that treats the center-of-mass dynamics the lattice. For these calculations, it is assumed that of the atoms (due to lattice-induced forces) classically the alignment between the normal vectors of the lattice and the internal, modulation-driven dynamics quantum- planes and the quantization axis is perfect. As suggested mechanically [10, 23, 39]. The effects of temperature on by Fig. 4, conditions can be chosen such that one pair the population fraction that becomes excited into the of lattice beams causes a transition shift with a different upper state are shown in Fig. 5 for a potential depth polarity than that due to another pair of lattice beams, of 35 kHz in a one-dimensional lattice. The central, such that the induced shifts can cancel each other out. Fourier-limited features shown correspond to trapped To achieve this type of magic lattice, the intensity of one atoms (no Doppler effect). As the temperature is low- of the beams is adjusted until the calculated transition ered, the fraction of atoms trapped in the optical-lattice lattice shift reaches zero. In Table III, the corresponding wells increases, leading to a corresponding increase of the uncertainties are generated by the assumption that the 9

TABLE III. Transition frequency shifts, relative transition shifts and relative uncertainties for ground-based experiment under conditions suitable for a kinetic temperature of 1 µK. Improved micro-gravity-conditions shifts and uncertainties are shown in square brackets for a temperature of 10 nK. The second-order Stark and diamagnetic shifts are lowered under these conditions since the field values are determined based on the kinetic temperature of the atoms (see section II.B). Here, we use ms = 1/2. Source ∆ν ∆ν/ν δ∆ν/ν (×10−12) Residual Lattice Shift 0 (3) 0 32 [0 (0.1)] [1.0] Dipolar Polarization Quantum Defect 120.1(3) Hz 1.283 × 10−9 2.8 2nd order Stark -6.8 (1) Hz −7.3 × 10−11 1.5 [-0.73 (1) mHz] [1.6 × 10−4] Diamagnetic 0.94 (4) Hz 1.0 × 10−11 0.4 [94 (4) µHz] [4.0 × 10−5] Mass Correction -605.08747 (3) kHz 6.4606271 × 10−6 0.3 Lamb Shift -84.1 (5) mHz −8.98 × 10−13 5.0 × 10−3 Blackbody a 0.64 (6) mHz 6.8 × 10−15 6.2 × 10−4 Quadrupolar Polarization Quantum Defect 26 (5) µHz 2.8 × 10−10 6.0 × 10−5 2nd order Doppler 0.05 (7)nHz 5 × 10−22 7.3 × 10−10 Fine Structure 488.0332466612(5)Hz 5.210818188587 × 10−9 1.6 × 10−9 Hyperfine Structure 32.89402(7) µHz 3.512153 × 10−16 7.2 × 10−12 1st order Stark 0 0 0 and Zeeman 0 0 0 1st order Doppler 0 0 0 a at 4 K. intensities can be controlled with a 1% relative uncer- be avoided in the first place by using hydrogen instead tainty. of rubidium. However, experimental obstacles due to the Smaller uncertainties are possible through the use of large recoil energy of hydrogen and laser-cooling on the shallower lattices and fewer laser beams. To achieve this, Lyman-α line are prohibitive at this time, leaving rubid- lower atomic temperatures are needed, for which other ium as an attractive option. well-known cooling methods can be employed [41]. More- The second-order Stark shift leads to an uncertainty over, for a similar experiment in micro-gravity conditions, due the electric field not being precisely known. For the the depth of the lattice could be further decreased or values displayed in Table III, we assume an electric-field even only used for driving transitions. This reduction uncertainty of 1% of the field magnitudes provided in of the lattice-induced uncertainty may be the only im- Table II. The resulting uncertainty is the third-largest in provement necessary to achieve a competitive Rydberg- Table III. It is seen in Table III that by performing the constant measurement. In Table III, lattice-shift esti- experiment under micro-gravity conditions the second- mates are presented for two cases of the lattice depth, order Stark shift uncertainty can be decreased by four one suitable for ground-based experiments and the other orders of magnitude (because the electric field can be for micro-gravity experiments. Table III also shows that dropped by two orders of magnitude) and hence the shift the lattice-induced shift represents the by far dominant goes from border-line significant to insignificant. source of systematic uncertainty and needs to be ad- dressed first in any incremental improvement of the ex- periment. B. Other sources of uncertainty The next-significant systematic shift arises from the dipole polarizability of the ionic core (see Table III), The following discussion shows that the remaining which currently stands at 9.12(2) [28], with experiments shifts listed in Table III present negligible uncertain- underway to improve the uncertainty in this value [38]. ties at the current level of precision (5.9 × 10−12), but The uncertainty due to the quadrupolar shift these are discussed here for completeness. is negligible in the overall uncertainty budget, because of The uncertainty displayed for the second-order Zeeman 6 the 1/r dependence of that shift [42, 43]. shift in Table III assumes the magnetic field is known to The quantum defect corrections are due to deviations 2% of its value that is dictated by the kinetic temperature from the hydrogenic 1/r potential. This applies to both of the atoms (see Table II). This precision can be achieved the penetration [11] (zero for the case of circular states) by monitoring the atomic Larmor precession of cold-atom and the polarization quantum defects. This issue could samples using the Faraday rotation technique. 10

The finite-mass correction, which accounts for the non- ture inside the spectroscopy enclosure. When calculating infinite mass of the nucleus, consists of a dominant first- the blackbody shift uncertainty, it is assumed that the order term and several higher-order terms. The first temperature is known to ±0.5 K. order can be considered by multiplying the Rydberg The main disturbance caused by the blackbody radia- constant by a factor of µ/me (µ = meM/(me + M)) tion is the broadening it induces on the spectral line due [9], where M is the Rb+ mass. The correction is to thermally-induced decays and excitations. A precise −605.08747 kHz, as shown in Table III. The mass correc- determination of the location of our measured frequency tion introduces an insignificant uncertainty to our mea- is essential in obtaining a successful measurement of R∞. surement since the mass of Rb+ (84.911 245 324 a.u.) Assuming a good signal-to-noise ratio, we expect to de- and that of the electron are well known (relative uncer- termine the line center to within 1/100 of the line-width. tainties of 4.4 × 10−8 [44] and 1.2 × 10−8 [5], respec- At a radiation temperature of 4 K, and for the states con- tively). The higher-order terms show up as factors of the sidered in this paper, the corresponding statistical uncer- η form (µ/me) in the fine structure (η = 1), second-order tainty is 0.3 Hz, corresponding to a relative uncertainty Stark effect (η = 3), diamagnetic shift (η = 1), Lamb of 3 × 10−12. shift (η = 2) and hyperfine structure (η = 1) corrections. Despite producing a relatively large shift, the fine- When the mass correction factor is considered for these, structure correction can be calculated very accurately the shifts decrease by 3 mHz, 0.1 mHz, 6 µHz, 1 µHz, and since the fine-structure constant is well known (relative 212 pHz, respectively. Since these differences are negligi- uncertainty of 2.3 × 10−10 [5]). As a result, the relative ble, we do not carry out these corrections in the results uncertainty introduced by the fine-structure shift is only shown in Table III. 1.6 × 10−21. In contrast to measurements on low-lying states of hy- Even though the hyperfine-structure shift in itself is drogen [4], through the use of circular states, we obtain negligible, its uncertainty is nevertheless estimated. The low QED corrections, since the valence electron has zero main sources of this uncertainty are the electron mass, probability of being in the vicinity of the nucleus, and proton mass and Planck’s constant (all of them with rel- the size of the Rydberg function becomes very large. ative uncertainties of 1.2 × 10−8) and the electron charge The main source of uncertainty in the Lamb shift is the (relative uncertainty of 6.1 × 10−9) [5]. The g-factors for Bethe logarithm (relative uncertainty of 3.3 × 10−8 and the nucleus ( 0.000 293 640 0) and the electron (2.002 319 1.0×10−7 or less for the circular and near-circular states, 304 361 53) are also well known (relative uncertainties of respectively [35]), leading to a negligible uncertainty for 6.0 × 10−10 [37] and 2.6 × 10−13 [5], respectively). our measurement. Effective trapping of atoms in the lattice leads to zero The blackbody shift is lowered three orders of mag- first-order and negligible second-order Doppler effects nitude by placing our system in thermal contact with (Fig. 5). As discussed in section III.B, the first-order liquid helium, which has a temperature of 4 K. Even so, Stark and Zeeman shifts are zero for our transition. This at 300 K the blackbody radiation shift for the transition is possible by choosing the lower and upper state so that of interest is just −21 mHz, making this shift negligible either the shift on each individual state is zero (first- for a wide range of typical experimental temperatures. order Stark shift) or both states experience the same shift The results of numerical calculations shown in Table III (first-order Zeeman shift). follow a similar procedure to that presented in [36]. How- ever, we consider only bound states up to about n=300. This truncation of the basis set does not affect the re- V. CONCLUSION sults significantly. By leaving out the last 250 states in the calculation, at worst, the calculated shift for the in- We have discussed an experimental method to help dividual states (about 2.4 kHz at 300 K and 0.42 Hz at solve the proton radius puzzle using a cold-atom-based 4 K) changes only by about 0.1 mHz at a temperature measurement of the Rydberg constant (R∞), which uti- of 300 K and by nano-Hertz at a temperature of 4 K. lizes (near-)circular Rydberg states and is free of QED This leads us to the conclusion that it is not necessary to shifts and sensitivity to nuclear charge overlap. Previ- include the continuum states in our calculations, which ous efforts to measure R∞ with Rydberg atoms have en- is also reaffirmed in [36]. We treat the near-circular state countered several experimental challenges which are ad- as a sum of spherical states multiplied by their respec- dressed in this proposed measurement. The first-order tive 3J symbols squared. The radiation field is taken to Zeeman and Stark shifts are both zero, owing to appro- be isotropic inside our spectroscopy enclosure, since at priate selection of parabolic atomic states involved in the the frequencies considered, the cavity density of states transition. By applying cooling and trapping techniques, approaches that of free space. With this treatment, we the interaction times are increased, leading to a reduc- obtain results comparable to those obtained in [36] for tion of the Fourier width. By using a new method of temperatures of 300 K. The radial matrix elements used spectroscopy in modulated optical lattices, the Doppler in the calculations are correct to four significant figures. broadening is eliminated. An implementation of the pro- The main source of uncertainty for the shift presented in posed experiment at atomic temperatures of 1 µK is pro- Table III is dictated by how well we know the tempera- jected to yield a Rydberg constant value with an uncer- 11 tainty in the low 10−11 range, limited almost exclusively VI. ACKNOWLEDGMENTS by lattice-trap-induced shifts. Since the proposed ex- periment differs from spectroscopy on low-lying atomic states, in that it is entirely insensitive to the proton ra- dius, this level of precision would be sufficient to make A.R. acknowledges support from the National Science a statement about the proton radius puzzle. The trap- Foundation Graduate Research Fellowship under Grant induced shifts could be very well addressed by perform- No. DGE 1256260. K.M. acknowledges support from ing the experiment under micro-gravity conditions, which the University of Michigan Rackham Pre-doctoral Fel- could lead to an uncertainty in the ≈ 3 × 10−12 range, lowship. This work was supported by NSF Grant No. almost a factor of two improvement over the current un- PHY1506093, and NASA Grant No.NNH13ZTT002N certainty. NRA.

[1] R. Pohl, A. Antognini, F. Nez, F. D. Amaro, F. Biraben, [21] D. Delande and J. Gay, EPL (Europhysics Letters) 5, J. M. Cardoso, D. S. Covita, A. Dax, S. Dhawan, L. M. 303 (1988). Fernandes, et al., Nature 466, 213 (2010). [22] S. K. Dutta, J. R. Guest, D. Feldbaum, A. Walz- [2] R. J. Hill, arXiv preprint arXiv:1702.01189 (2017). Flannigan, and G. Raithel, Phys. Rev. Lett. 85, 5551 [3] R. Pohl, F. Nez, L. M. Fernandes, F. D. Amaro, (2000). F. Biraben, J. M. Cardoso, D. S. Covita, A. Dax, [23] K. R. Moore and G. Raithel, Phys. Rev. Lett. 115, S. Dhawan, M. Diepold, et al., Science 353, 669 (2016). 163003 (2015). [4] P. J. Mohr and B. N. Taylor, Rev. Mod. Phys. 77, 1 [24] H. Friedrich, Theoretical atomic physics, Vol. 3 (Springer, (2005). 2006). [5] P. Mohr, B. Taylor, and D. Newell, Web Version 7 [25] K. C. Younge, S. E. Anderson, and G. Raithel, New (2014). Journal of Physics 12, 023031 (2010). [6] J. De Vries, “A precision millimeter-wave measurement [26] R. R. Freeman and D. Kleppner, Phys. Rev. A 14, 1614 of the rydberg frequency phd thesis,” (2001). (1976). [7] J. Hare, A. Nussenzweig, C. Gabbanini, M. Weidemuller, [27] L. J. Curtis, Atomic structure and lifetimes: a conceptual P. Goy, M. Gross, and S. Haroche, Instrumentation and approach (Cambridge University Press, 2003). Measurement, IEEE Transactions on 42, 331 (1993). [28] J. Lee, J. Nunkaew, and T. Gallagher, Physical Review [8] U. D. Jentschura, P. J. Mohr, and J. N. Tan, Journal A 94, 022505 (2016). of Physics B: Atomic, Molecular and Optical Physics 43, [29] J. a. Mitroy, M. Safronova, and C. W. Clark, Journal 074002 (2010). of Physics B: Atomic, Molecular and Optical Physics 43, [9] H. A. Bethe and E. E. Salpeter, of 202001 (2010). one-and two-electron atoms (Springer Science & Business [30] J. Heinrichs, The Journal of Chemical Physics 52 (1970). Media, 2012). [31] R. M. Sternheimer, Phys. Rev. A 1, 321 (1970). [10] K. R. Moore, S. E. Anderson, and G. Raithel, Nature [32] H. Eissa and U. Opik,¨ Proceedings of the Physical Society communications 6 (2015). 92, 556 (1967). [11] T. F. Gallagher, Rydberg atoms, Vol. 3 (Cambridge Uni- [33] S. Krishnagopal, S. Narasimhan, and S. Patil, The Jour- versity Press, 2005). nal of chemical physics 83, 5772 (1985). [12] U. D. Jentschura, P. J. Mohr, J. N. Tan, and B. J. [34] J. J. Sakurai and J. Napolitano, Modern quantum me- Wundt, Phys. Rev. Lett. 100, 160404 (2008). chanics (Addison-Wesley, 2011). [13] P. J. Mohr, B. N. Taylor, and D. B. Newell, Rev. Mod. [35] G. W. Erickson, Journal of physical and chemical refer- Phys. 84, 1527 (2012). ence data 6, 831 (1977). [14] S. E. Anderson, K. C. Younge, and G. Raithel, Phys. [36] J. W. Farley and W. H. Wing, Phys. Rev. A 23, 2397 Rev. Lett. 107, 263001 (2011). (1981). [15] M. Takamoto, F.-L. Hong, R. Higashi, and H. Katori, [37] E. Arimondo, M. Inguscio, and P. Violino, Rev. Mod. Nature 435, 321 (2005). Phys. 49, 31 (1977). [16] C. Salomon, J. Dalibard, W. D. Phillips, A. Clairon, and [38] K. R. Moore and G. Raithel, “Measurement of rb g-series S. Guellati, AIP Conference Proceedings 233 (1991). quantum defect using two- ng → (n + 2)g mi- [17] D. Boiron, A. Michaud, P. Lemonde, Y. Castin, C. Sa- crowave spectroscopy,”. lomon, S. Weyers, K. Szymaniec, L. Cognet, and A. Cla- [39] K. C. Younge, S. E. Anderson, and G. Raithel, New iron, Phys. Rev. A 53, R3734 (1996). Journal of Physics 12, 113036 (2010). [18] P. Nussenzveig, F. Bernardot, M. Brune, J. Hare, J. M. [40] J. C. Bernauer and R. Pohl, Scientific American 310, 32 Raimond, S. Haroche, and W. Gawlik, Phys. Rev. A 48, (2014). 3991 (1993). [41] H. J. Metcalf and P. Straten, Laser cooling and trapping [19] R. G. Hulet and D. Kleppner, of neutral atoms (Wiley Online Library, 2007). 51, 1430 (1983). [42] W. Li, I. Mourachko, M. W. Noel, and T. F. Gallagher, [20] J. Hare, M. Gross, and P. Goy, Phys. Rev. Lett. 61, 1938 Phys. Rev. A 67, 052502 (2003). (1988). [43] J. Han, Y. Jamil, D. V. L. Norum, P. J. Tanner, and T. F. Gallagher, Phys. Rev. A 74, 054502 (2006). 12

[44] B. J. Mount, M. Redshaw, and E. G. Myers, Physical Review. A 82, 042513 (2010).