Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Engineering Rydberg-pair interactions
in divalent atoms with hyperfine-split ionization thresholds

Frederic Hummel fhummel@atom-computing.com Atom Computing, Inc., Berkeley, California    Sebastian Weber weber@itp3.uni-stuttgart.de Institute for Theoretical Physics III and Center for Integrated Quantum Science and Technology, Universität Stuttgart, Germany    Johannes Mögerle Institute for Theoretical Physics III and Center for Integrated Quantum Science and Technology, Universität Stuttgart, Germany    Henri Menke Department of Physics, Friedrich-Alexander-Universität Erlangen-Nürnberg, Germany    Jonathan King Atom Computing, Inc., Berkeley, California    Benjamin Bloom Atom Computing, Inc., Berkeley, California    Sebastian Hofferberth hofferberth@iap.uni-bonn.de Institute of Applied Physics, University of Bonn, Germany    Ming Li mli@atom-computing.com Atom Computing, Inc., Berkeley, California
(July 31, 2024)
Abstract

Quantum information processing with neutral atoms relies on Rydberg excitation for entanglement generation. While the use of heavy divalent or open-shell elements, such as strontium or ytterbium, has benefits due to their optically active core and a variety of possible qubit encodings, their Rydberg structure is generally complex. For some isotopes in particular, hyperfine interactions are relevant even for highly excited electronic states. We employ multi-channel quantum defect theory to infer the Rydberg structure of isotopes with non-zero nuclear spin and perform non-perturbative Rydberg-pair interaction calculations. We find that due to the high level density and sensitivities to external fields, experimental parameters must be precisely controlled. Specifically in 87Sr, we study an intrinsic Förster resonance, unique to divalent atoms with hyperfine-split thresholds, which simultaneously provides line stability with respect to external field fluctuations and enhanced long-range interactions. Additionally, we provide parameters for pair states that can be effectively described by single-channel Rydberg series. The explored pair states provide exciting opportunities for applications in the blockade regime as well as for more exotic long-range interactions such as largely flat, distance-independent potentials.

Rydberg atoms, hyperfine interaction, pair-interaction potentials, Förster resonance

I Introduction

The long-range interactions between Rydberg atoms cause exotic processes in cold quantum gases and enable applications in non-linear quantum optics [1, 2, 3]. Coherent Rydberg excitations are essential to quantum information processing using neutral atoms in order to introduce and mediate entanglement [4, 5, 6]. Ensembles of individually trapped atoms in optical tweezer arrays have enabled incredible progress for quantum simulation and computation with defect-free assembly [7, 8], reconfigurable lattice geometries [9, 10, 11, 12], mid-circuit erasure conversion and readout [13, 14], continuous reloading [15, 16], and unseen connectivities by aid of rearrangement [17]. Recently, a variety of atomic species has been explored for quantum gas experiments and quantum technology applications exploiting different aspects of the atomic structure [18, 19, 20, 21, 22]. While alkali metals benefit from a long history of scientific research, the presence of a second valence electron in alkaline-earth and some lanthanide elements provides compelling control opportunities [23], e.g. the core electron can be utilized to match polarizabilities of ground- and Rydberg states for magic wavelength trapping [24] and isolated core excitations can be exploited to introduce coherent light shifts and probe Rydberg states [25, 26]. Some isotopes allow for the storage of quantum information in their nuclear spin, achieving coherence times on the order of tens of seconds [27].

Most schemes to entangle neutral atoms rely on strong dipolar interactions between Rydberg states, which are either coupled by near-resonant laser fields [28, 29, 30, 31] or by dressing ground states with far off-resonant lasers [32, 33, 34, 35]. Yet, for the Rydberg structure, typically only one or at most a few energy levels are considered explicitly. Non-perturbative Rydberg-pair interaction calculations are required to confirm the absence of other pair states and leakage channels in the Rydberg spectrum [18, 36, 37].

In this article, we present calculations of non-perturbative Rydberg-pair interactions for divalent atoms with hyperfine-split ionization thresholds in order to study interaction tunabilty and the robustness of perturbative predictions. The complex spectra of these atoms are described within multi-channel quantum defect theory [38], which has been adapted to include nuclear spin using frame transformation [39]. Recently, 87Sr was predicted to feature unusual properties not encountered in atoms with a simpler structure, namely, for some Rydberg states, simultaneous insensitivity to electric fields and resonant dipolar interaction in equal-state pairs [40]. We find this to be a general feature caused by the hyperfine interaction between different Rydberg series when the energy splitting of adjacent Rydberg manifolds becomes comparable to the threshold splitting in the ionic core. Specifically, the coupling of Rydberg series belonging to different ionic-core hyperfine configurations gives rise to varying electron spin character of some Rydberg states as a function of the principal quantum number [41, 42] and to a corresponding energy dependence of their quantum defects. The resulting properties are appealing for quantum technology applications and Rydberg gas experiments: High degrees of level stability help preparing Rydberg states that minimize amplitude and phase errors. Intrinsic Förster resonances lead to strong dipole-dipole interactions, allowing for the engineering of exotic many-body Hamiltonians and for the implementation of two- and multi-qubit gates [43]. These Förster resonances do not require careful tuning of external fields, which is often the case for atoms with a single valence electron [44, 45, 46, 47, 48], but tuning of the principal quantum number is in general sufficient as we will demonstrate. Moreover, the Förster resonances occur for two atoms being in the same state, allowing for global addressing of the atoms via a single laser beam.

This article is structured as follows: In Sec. II, we review the theory and methods, particularly the Hamiltonian for a pair of Rydberg atoms in Sec. II.1. After this general discussion of Rydberg interactions, we provide the methods specific to divalent Rydberg atoms in Sec. II.2, where we outline multi-channel quantum defect theory and frame transformation. Sec. II.3 details 87Sr, with comments on exact diagonalization in Sec. II.4 and on the developed software in Sec. II.5. Sec. III contains the results, discussing the single-atom spectral structure in Sec. III.1 and introducing field-free pair-interaction potentials for the Förster resonant Rydberg series in Sec. III.2. Sec. III.3 covers the response of single atoms and Rydberg pairs to magnetic fields, similarly for electric fields in Sec. III.4. A comparison to a different, single-channel Rydberg series is provided in Sec. III.5. Finally, Sec. IV contains the conclusion and an outlook.

II Theory and Methods

II.1 Rydberg-pair Hamiltonian

Refer to caption
Figure 1: The molecular system consisting of two Rydberg atoms with internuclear coordinate 𝑹𝑹\bm{R}bold_italic_R and electronic coordinates 𝒓isubscript𝒓𝑖\bm{r}_{i}bold_italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT for atom i𝑖iitalic_i, respectively. For each individual atom, the total hyperfine spin f𝑓fitalic_f is a good quantum number. In the fragmentation frame, 𝒇𝒇\bm{f}bold_italic_f is obtained by coupling the ionic core’s hyperfine angular momentum 𝒇csubscript𝒇𝑐\bm{f}_{c}bold_italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and the Rydberg electron’s total angular momentum 𝒋rsubscript𝒋𝑟\bm{j}_{r}bold_italic_j start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT. The coupling scheme for short electron separations 𝒓isubscript𝒓𝑖\bm{r}_{i}bold_italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT depends on the particular eigenstate.

We consider a pair of atoms as illustrated in Fig. 1 and modelled by the Hamiltonian

H2=H111+11H1+Hint(𝑹),subscript𝐻2tensor-productsubscript𝐻111tensor-product11subscript𝐻1subscript𝐻int𝑹H_{2}=H_{1}\otimes 1\!\!1+1\!\!1\otimes H_{1}+H_{\text{int}}(\bm{R}),italic_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_H start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⊗ 1 1 + 1 1 ⊗ italic_H start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT int end_POSTSUBSCRIPT ( bold_italic_R ) , (1)

where H1subscript𝐻1H_{1}italic_H start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is the single-atom Hamiltonian detailed further below and (using atomic units throughout)

Hint(𝑹)=κ1,κ2=1Vκ1κ2|𝑹|1κ1κ2subscript𝐻int𝑹superscriptsubscriptsubscript𝜅1subscript𝜅21subscript𝑉subscript𝜅1subscript𝜅2superscript𝑹1subscript𝜅1subscript𝜅2H_{\text{int}}(\bm{R})=\sum_{\kappa_{1},\kappa_{2}=1}^{\infty}V_{\kappa_{1}% \kappa_{2}}|\bm{R}|^{-1-\kappa_{1}-\kappa_{2}}italic_H start_POSTSUBSCRIPT int end_POSTSUBSCRIPT ( bold_italic_R ) = ∑ start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_κ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_V start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT | bold_italic_R | start_POSTSUPERSCRIPT - 1 - italic_κ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_κ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT (2)

is the multipole expansion of the electrostatic interaction between the two atoms suitable for internuclear separations exceeding the Le Roy radius and neglecting retardation effects. Here, Vκ1κ2subscript𝑉subscript𝜅1subscript𝜅2V_{\kappa_{1}\kappa_{2}}italic_V start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT is proportional to the spherical multipole operators on the Rydberg electron in atom 1 and 2, respectively. The exact form of Vκ1κ2subscript𝑉subscript𝜅1subscript𝜅2V_{\kappa_{1}\kappa_{2}}italic_V start_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT depends on the choice of the coordinate system. We choose the internuclear axis 𝑹𝑹\bm{R}bold_italic_R to align with the z𝑧zitalic_z axis and restrict Eq. 2 to dipole-dipole interactions:

V11=q=11(21+q)r1q(1)r1q(2),subscript𝑉11superscriptsubscript𝑞11binomial21𝑞superscriptsubscript𝑟1𝑞1superscriptsubscript𝑟1𝑞2V_{11}=-\sum_{q=-1}^{1}\binom{2}{1+q}r_{1q}^{(1)}r_{1-q}^{(2)},italic_V start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT = - ∑ start_POSTSUBSCRIPT italic_q = - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( FRACOP start_ARG 2 end_ARG start_ARG 1 + italic_q end_ARG ) italic_r start_POSTSUBSCRIPT 1 italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT 1 - italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT , (3)

where rκq(i)superscriptsubscript𝑟𝜅𝑞𝑖r_{\kappa q}^{(i)}italic_r start_POSTSUBSCRIPT italic_κ italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT is the spherical multipole operator of order κ𝜅\kappaitalic_κ for atom i𝑖iitalic_i.

We find a matrix representation of H2subscript𝐻2H_{2}italic_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT by expanding in a set of appropriately symmetrized wave functions of pairs of atoms. In the absence of external fields, H2subscript𝐻2H_{2}italic_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT preserves reflection symmetry through a plane containing the internuclear axis. Since we consider homonuclear pairs, H2subscript𝐻2H_{2}italic_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT also preserves inversion symmetry. In the case of external field alignment with the internuclear axis, the total magnetic quantum number M=m1+m2𝑀subscript𝑚1subscript𝑚2M=m_{1}+m_{2}italic_M = italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is a conserved quantity, where we introduce an index, indicating misubscript𝑚𝑖m_{i}italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT corresponds to atom i𝑖iitalic_i. Finally, since only dipole-dipole interaction are considered in the multipole expansion (Eq. 2), H2subscript𝐻2H_{2}italic_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT preserves permutation symmetry. We use these symmetries to decompose H2subscript𝐻2H_{2}italic_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT into blocks that can be treated independently of each other [49].

The Hamiltonian for a single Rydberg atom is given by

H1=H0+Hs+Hz+Hd,subscript𝐻1subscript𝐻0subscript𝐻ssubscript𝐻zsubscript𝐻dH_{1}=H_{0}+H_{\text{s}}+H_{\text{z}}+H_{\text{d}},italic_H start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT s end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT z end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT d end_POSTSUBSCRIPT , (4)

where H0subscript𝐻0H_{0}italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the field-free Hamiltonian with the good quantum numbers f𝑓fitalic_f and m𝑚mitalic_m corresponding to the total hyperfine spin, as well as parity. The eigenvalues of H0subscript𝐻0H_{0}italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT,

Eνf=EIRμν2,subscript𝐸𝜈𝑓subscript𝐸𝐼subscript𝑅𝜇superscript𝜈2E_{\nu f}=E_{I}-\frac{R_{\mu}}{\nu^{2}},italic_E start_POSTSUBSCRIPT italic_ν italic_f end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT - divide start_ARG italic_R start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG start_ARG italic_ν start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (5)

are conveniently expressed in terms of an effective principal quantum number ν𝜈\nuitalic_ν, defined with respect to the ionization threshold EIsubscript𝐸𝐼E_{I}italic_E start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT. Here, Rμsubscript𝑅𝜇R_{\mu}italic_R start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT is the mass scaled Rydberg constant. If an eigenstate is a superposition of multi-channel configurations as described in the following section, it is still useful to choose a specific EIsubscript𝐸𝐼E_{I}italic_E start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT in order to label eigenstates by a global ν𝜈\nuitalic_ν and, hence, indicate their ordering. For hyperfine split-ionization thresholds, we choose EIsubscript𝐸𝐼E_{I}italic_E start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT to be the energetically higher one, which some single-channel Rydberg series converge to.

The effect of static and homogeneous external fields is considered by the following three terms: For the electric field 𝑭𝑭\bm{F}bold_italic_F, the Stark Hamiltonian is given by

Hs=𝒓𝑭,subscript𝐻s𝒓𝑭H_{\text{s}}=-\bm{r}\cdot\bm{F},italic_H start_POSTSUBSCRIPT s end_POSTSUBSCRIPT = - bold_italic_r ⋅ bold_italic_F , (6)

where 𝒓𝒓\bm{r}bold_italic_r is the electronic coordinate (and, since the electron charge e=1𝑒1e=1italic_e = 1 in atomic units, corresponds to the electric dipole moment). For the magnetic field 𝑩𝑩\bm{B}bold_italic_B, the Zeeman Hamiltonian is given by

Hz=𝝁𝑩,subscript𝐻z𝝁𝑩H_{\text{z}}=-\bm{\mu}\cdot\bm{B},italic_H start_POSTSUBSCRIPT z end_POSTSUBSCRIPT = - bold_italic_μ ⋅ bold_italic_B , (7)

where 𝝁𝝁\bm{\mu}bold_italic_μ is the magnetic dipole moment taking into account the electronic and nuclear spins

𝝁=μB[c+r+gs(𝒔c+𝒔r)]+μI𝑰,𝝁subscript𝜇𝐵delimited-[]subscriptbold-ℓ𝑐subscriptbold-ℓ𝑟subscript𝑔𝑠subscript𝒔𝑐subscript𝒔𝑟subscript𝜇𝐼𝑰\bm{\mu}=-\mu_{B}\left[\bm{\ell}_{c}+\bm{\ell}_{r}+g_{s}\left(\bm{s}_{c}+\bm{s% }_{r}\right)\right]+\mu_{I}\bm{I},bold_italic_μ = - italic_μ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT [ bold_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT + bold_ℓ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_g start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( bold_italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT + bold_italic_s start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) ] + italic_μ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT bold_italic_I , (8)

where μBsubscript𝜇𝐵\mu_{B}italic_μ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT and μIsubscript𝜇𝐼\mu_{I}italic_μ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT are the Bohr magneton and the nuclear magnetic moment and gssubscript𝑔𝑠g_{s}italic_g start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is the electron spin g-factor. Here, the subscripts, c𝑐citalic_c and r𝑟ritalic_r label core and Rydberg electron quantum numbers, while bold-ℓ\bm{\ell}bold_ℓ and 𝒔𝒔\bm{s}bold_italic_s are orbital and electronic spin angular momentum, respectively, and 𝑰𝑰\bm{I}bold_italic_I is the nuclear spin (cf. Fig. 1). The diamagnetic Hamiltonian is given by

Hd=18(𝒓2𝑩2|𝒓𝑩|2).subscript𝐻d18superscript𝒓2superscript𝑩2superscript𝒓𝑩2H_{\text{d}}=\frac{1}{8}\left(\bm{r}^{2}\bm{B}^{2}-\left|\bm{r}\cdot\bm{B}% \right|^{2}\right).italic_H start_POSTSUBSCRIPT d end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 8 end_ARG ( bold_italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - | bold_italic_r ⋅ bold_italic_B | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) . (9)

This term is typically considered for ground-state atoms only when the magnetic field is on the order of 1 atomic unit (104similar-toabsentsuperscript104\sim 10^{4}∼ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT T). However, as it scales with the surface area of the atom, Hdn4similar-tosubscript𝐻dsuperscript𝑛4H_{\text{d}}\sim n^{4}italic_H start_POSTSUBSCRIPT d end_POSTSUBSCRIPT ∼ italic_n start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT, for the Rydberg states considered here, diamagnetic interaction is comparable to Zeeman interaction for fields as low as 100similar-toabsent100\sim 100∼ 100 G.

In practise, before dealing with the interactions between two atom, the total single-atom Hamiltonian H1subscript𝐻1H_{1}italic_H start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is constructed within a restricted basis (see Sec. II.4) and diagonalized for a specific external field configuration; a process we call prediagonalization [49]. An appropriate two-atom basis is then constructed by pairwise, symmetrized combinations of the resulting eigenstates with the option to provide an energy cutoff. This is particularly relevant in order to only include well converged single-atom states in the two-atom basis. Finally, Rydberg pair-interaction potentials are obtained by constructing H2subscript𝐻2H_{2}italic_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT in the two-atom basis and subsequent diagonalization at the internuclear distances of interest.

II.2 Multi-channel quantum defect theory

In order to find the eigensystem of H0subscript𝐻0H_{0}italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, we follow the formalism presented by Robicheaux et al. combining multi-channel quantum defect theory (MQDT) and frame transformations to capture complex spin interactions in heavy atoms [39]. Generally, the perspective of quantum defect theory is to consider an atom as a collision of an electron and an ionic core [50]. In MQDT, the collision is described in terms of the real, symmetric K𝐾Kitalic_K matrix that parameterizes the scattering between different spin configurations, so-called channels [38]. The frame transformation is the process of unitarily transforming K𝐾Kitalic_K matrices according to different coupling schemes that depend on the relevant length scale of the Rydberg electronic coordinate [51]. Each frame has a different advantage: In the fragmentation frame, it is straight-forward to incorporate multiple ionization thresholds and account for atomic hyperfine structure (or in the case of molecules, for vibrational or rotational degrees of freedom). The close-coupling frame, on the other hand, provides a diagonal representation of the K matrix.

From a scattering perspective, the total hyperfine spin of the Rydberg atom f𝑓fitalic_f is formed by coupling the asymptotically separated spins of the Rydberg electron jrsubscript𝑗𝑟j_{r}italic_j start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and the ionic core’s hyperfine spin fcsubscript𝑓𝑐f_{c}italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. However, at small separations, jrsubscript𝑗𝑟j_{r}italic_j start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and fcsubscript𝑓𝑐f_{c}italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT are not meaningful and the fragmentation-frame coupling scheme has to be replaced by a more suitable close-coupling scheme, e.g. spin-orbit coupling. This provides the additional advantage that the the K matrix in the close-coupling frame, obtained from spectroscopic data, are applicable (to a good approximation) independently of the isotope and the corresponding nuclear spin configuration. An intuitive explanation for this is provided by comparing the energy scales of different spin couplings: Close to the ground-state, the hyperfine structure is only a small correction to the eigenstates dominated by spin-orbit coupling (or jcjrsubscript𝑗𝑐subscript𝑗𝑟j_{c}j_{r}italic_j start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_j start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT-coupling for some excited states, jcsubscript𝑗𝑐j_{c}italic_j start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT being the total angular momentum of the core electrons) and therefore has very limited impact on the coupling parameters.

The formalism presented in [39] and outlined below solves for the bound states of H0subscript𝐻0H_{0}italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT after transforming from the closed-coupling frame into the fragmentation frame. This is equivalent to, but leads to different equations than solving directly in the close-coupling frame, presented e.g. in [52]. Let us start in the close-coupling frame, in which we assume the K𝐾Kitalic_K matrix to be diagonal, denoted as 𝑲¯¯𝑲\bar{\bm{K}}over¯ start_ARG bold_italic_K end_ARG, with elements

K¯α=tanπμα,subscript¯𝐾𝛼𝜋subscript𝜇𝛼\bar{K}_{\alpha}=\tan\pi\mu_{\alpha},over¯ start_ARG italic_K end_ARG start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = roman_tan italic_π italic_μ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT , (10)

where μαsubscript𝜇𝛼\mu_{\alpha}italic_μ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT are (possibly energy dependent) quantum defects and α𝛼\alphaitalic_α labels the close-coupling channels. We choose the transformation into the fragmentation frame to be expressed as a series of unitary matrices

𝑼^ccfj=𝑼^jjfj𝑼^LSjj𝑼^ccLS,subscript^𝑼𝑐𝑐𝑓𝑗subscript^𝑼𝑗𝑗𝑓𝑗subscript^𝑼𝐿𝑆𝑗𝑗subscript^𝑼𝑐𝑐𝐿𝑆\hat{\bm{U}}_{cc-fj}=\hat{\bm{U}}_{jj-fj}\hat{\bm{U}}_{LS-jj}\hat{\bm{U}}_{cc-% LS},over^ start_ARG bold_italic_U end_ARG start_POSTSUBSCRIPT italic_c italic_c - italic_f italic_j end_POSTSUBSCRIPT = over^ start_ARG bold_italic_U end_ARG start_POSTSUBSCRIPT italic_j italic_j - italic_f italic_j end_POSTSUBSCRIPT over^ start_ARG bold_italic_U end_ARG start_POSTSUBSCRIPT italic_L italic_S - italic_j italic_j end_POSTSUBSCRIPT over^ start_ARG bold_italic_U end_ARG start_POSTSUBSCRIPT italic_c italic_c - italic_L italic_S end_POSTSUBSCRIPT , (11)

where the subscripts label coupling schemes (detailed below), the first two matrices are solely determined by the corresponding angular momentum recouplings. 𝑼^ccLSsubscript^𝑼𝑐𝑐𝐿𝑆\hat{\bm{U}}_{cc-LS}over^ start_ARG bold_italic_U end_ARG start_POSTSUBSCRIPT italic_c italic_c - italic_L italic_S end_POSTSUBSCRIPT accounts for corrections to pure spin-orbit (LS𝐿𝑆LSitalic_L italic_S) coupling that is typically expressed as a series of simple rotations between specific LS𝐿𝑆LSitalic_L italic_S channels, the (possibly energy dependent) angles of which serve as an optimization parameter in order to fit experimental data111 For light atoms, this is often not required and the K𝐾Kitalic_K matrix is diagonal in the LS𝐿𝑆LSitalic_L italic_S frame. As pointed out in [40], for 87Sr, a coupling between the 5snd5𝑠𝑛𝑑5snd5 italic_s italic_n italic_d series singlet and triplet states, D21superscriptsubscript𝐷21{}^{1}D_{2}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT italic_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and D23superscriptsubscript𝐷23{}^{3}D_{2}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT italic_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, can be accounted for by this additional rotation.. We list here the representations of the coupling schemes that make for the frame transformation matrices analogously to references [39, 40]. When the Rydberg electron is close to the core, we write the wave function as

|LS=|(((scsr)S(lclr)L)JI)fm,ket𝐿𝑆ketsubscript𝑠𝑐subscript𝑠𝑟𝑆subscript𝑙𝑐subscript𝑙𝑟𝐿𝐽𝐼𝑓𝑚|LS\rangle=|(((s_{c}s_{r})S(l_{c}l_{r})L)JI)fm\rangle,| italic_L italic_S ⟩ = | ( ( ( italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) italic_S ( italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_l start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) italic_L ) italic_J italic_I ) italic_f italic_m ⟩ , (12)

where parentheses indicate which spins are coupled consecutively. This configuration is recoupled in an intermediate step to

|jj=|(((sclc)jc(srlr)jr)JI)fm,ket𝑗𝑗ketsubscript𝑠𝑐subscript𝑙𝑐subscript𝑗𝑐subscript𝑠𝑟subscript𝑙𝑟subscript𝑗𝑟𝐽𝐼𝑓𝑚|jj\rangle=|(((s_{c}l_{c})j_{c}(s_{r}l_{r})j_{r})JI)fm\rangle,| italic_j italic_j ⟩ = | ( ( ( italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) italic_j start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_s start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_l start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) italic_j start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) italic_J italic_I ) italic_f italic_m ⟩ , (13)

before the final recoupling to the fragmentation frame

|fj=|(((sclc)jcI)fc(srlr)jr)fm.ket𝑓𝑗ketsubscript𝑠𝑐subscript𝑙𝑐subscript𝑗𝑐𝐼subscript𝑓𝑐subscript𝑠𝑟subscript𝑙𝑟subscript𝑗𝑟𝑓𝑚|fj\rangle=|(((s_{c}l_{c})j_{c}I)f_{c}(s_{r}l_{r})j_{r})fm\rangle.| italic_f italic_j ⟩ = | ( ( ( italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) italic_j start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_I ) italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_s start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_l start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) italic_j start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) italic_f italic_m ⟩ . (14)

This scheme can be summarized to be propagating the coupling of the valence electrons to the nuclear spin through different stages.

Finally, the K𝐾Kitalic_K matrix in the fragmentation frame is obtained as

𝑲^=𝑼^ccfj𝑲¯𝑼^ccfj.^𝑲subscript^𝑼𝑐𝑐𝑓𝑗¯𝑲superscriptsubscript^𝑼𝑐𝑐𝑓𝑗\hat{\bm{K}}=\hat{\bm{U}}_{cc-fj}\bar{\bm{K}}\hat{\bm{U}}_{cc-fj}^{\dagger}.over^ start_ARG bold_italic_K end_ARG = over^ start_ARG bold_italic_U end_ARG start_POSTSUBSCRIPT italic_c italic_c - italic_f italic_j end_POSTSUBSCRIPT over¯ start_ARG bold_italic_K end_ARG over^ start_ARG bold_italic_U end_ARG start_POSTSUBSCRIPT italic_c italic_c - italic_f italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT . (15)

Generally, the K𝐾Kitalic_K matrix is block diagonal in f𝑓fitalic_f and in parity. Thus, under the assumption c=0subscript𝑐0\ell_{c}=0roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 0, it further decomposes into blocks of equal orbital angular momentum rsubscript𝑟\ell_{r}roman_ℓ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT.

A bound state |ψνfketsubscript𝜓𝜈𝑓|\psi_{\nu f}\rangle| italic_ψ start_POSTSUBSCRIPT italic_ν italic_f end_POSTSUBSCRIPT ⟩ at energy Eνfsubscript𝐸𝜈𝑓E_{\nu f}italic_E start_POSTSUBSCRIPT italic_ν italic_f end_POSTSUBSCRIPT is obtained by solving

det𝑴^=0,^𝑴0\displaystyle\det\hat{\bm{M}}=0,roman_det over^ start_ARG bold_italic_M end_ARG = 0 , (16a)
𝑴^=tan𝜷𝟏𝟏^+𝑲^,^𝑴𝜷^11^𝑲\displaystyle\hat{\bm{M}}=\tan\bm{\beta}\ \hat{\bm{1\!\!1}}+\hat{\bm{K}},over^ start_ARG bold_italic_M end_ARG = roman_tan bold_italic_β over^ start_ARG bold_1 bold_1 end_ARG + over^ start_ARG bold_italic_K end_ARG , (16b)

where the elements of 𝜷𝜷\bm{\beta}bold_italic_β, βi=π(νir,i)subscript𝛽𝑖𝜋subscript𝜈𝑖subscript𝑟𝑖\beta_{i}=\pi(\nu_{i}-\ell_{r,i})italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_π ( italic_ν start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - roman_ℓ start_POSTSUBSCRIPT italic_r , italic_i end_POSTSUBSCRIPT ), are related to the relative principal quantum numbers νisubscript𝜈𝑖\nu_{i}italic_ν start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT defined with respect to the associated ionization threshold of the i𝑖iitalic_i-th channel, i.e.

E=EiRμνi2,i,𝐸subscript𝐸𝑖subscript𝑅𝜇superscriptsubscript𝜈𝑖2for-all𝑖E=E_{i}-\frac{R_{\mu}}{\nu_{i}^{2}},\ \forall i,italic_E = italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - divide start_ARG italic_R start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG start_ARG italic_ν start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , ∀ italic_i , (17)

for total energy E𝐸Eitalic_E and thresholds Eisubscript𝐸𝑖E_{i}italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. At a bound state energy Eνfsubscript𝐸𝜈𝑓E_{\nu f}italic_E start_POSTSUBSCRIPT italic_ν italic_f end_POSTSUBSCRIPT, the coefficients for each channel Aisubscript𝐴𝑖A_{i}italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are defined via the one-dimensional null space

𝑨=ker((𝝂3/2cos𝜷).𝑴^),\displaystyle\bm{A}=\ker\left((\bm{\nu}^{-3/2}\cos\bm{\beta})^{\top}.\hat{\bm{% M}}\right),bold_italic_A = roman_ker ( ( bold_italic_ν start_POSTSUPERSCRIPT - 3 / 2 end_POSTSUPERSCRIPT roman_cos bold_italic_β ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT . over^ start_ARG bold_italic_M end_ARG ) , (18a)
𝑨2=1,superscript𝑨21\displaystyle\bm{A}^{2}=1,bold_italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1 , (18b)

where the exponentiation of 𝝂𝝂\bm{\nu}bold_italic_ν, containing elements νisubscript𝜈𝑖\nu_{i}italic_ν start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, as well as the matrix multiplication (.) are assumed to be element-wise operations. Then, the total MQDT wave function is given as

|ψνf=iAi(νf)Pνir,i(r)|ϕi,ketsubscript𝜓𝜈𝑓subscript𝑖superscriptsubscript𝐴𝑖𝜈𝑓subscript𝑃subscript𝜈𝑖subscript𝑟𝑖𝑟ketsubscriptitalic-ϕ𝑖|\psi_{\nu f}\rangle=\sum_{i}A_{i}^{(\nu f)}P_{\nu_{i}\ell_{r,i}}(r)|\phi_{i}\rangle,| italic_ψ start_POSTSUBSCRIPT italic_ν italic_f end_POSTSUBSCRIPT ⟩ = ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_ν italic_f ) end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_ℓ start_POSTSUBSCRIPT italic_r , italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_r ) | italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ , (19)

with the channel function |ϕiketsubscriptitalic-ϕ𝑖|\phi_{i}\rangle| italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ representing the ionic core as well as the spin and angular configuration of the Rydberg electron. The radial representation of the Rydberg electron is given by

Pνir,i(r)=(ficosβi+gisinβi)νi3/2,subscript𝑃subscript𝜈𝑖subscript𝑟𝑖𝑟subscript𝑓𝑖subscript𝛽𝑖subscript𝑔𝑖subscript𝛽𝑖superscriptsubscript𝜈𝑖32P_{\nu_{i}\ell_{r,i}}(r)=\left(f_{i}\cos\beta_{i}+g_{i}\sin\beta_{i}\right)\nu% _{i}^{-3/2},italic_P start_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_ℓ start_POSTSUBSCRIPT italic_r , italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_r ) = ( italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_cos italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_g start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_sin italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_ν start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 3 / 2 end_POSTSUPERSCRIPT , (20)

with the regular and irregular Coulomb functions fisubscript𝑓𝑖f_{i}italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and gisubscript𝑔𝑖g_{i}italic_g start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, respectively. Eq. 16 guarantees the Rydberg wave function to go to zero asymptotically.

The total wave functions (Eq. 19) containing the radial representation only for the Rydberg electron (Eq. 20), but angular representation for both the Rydberg electron and the ionic core, are used for the calculation of all matrix elements, relevant for constructing the Hamiltonians discussed in the previous sections, particularly the electric multipole moments of order κ𝜅\kappaitalic_κ

ψνf|rκq|ψνf=i,jAi(νf)Aj(νf)Rijκϕi|Yκq|ϕj,quantum-operator-productsubscript𝜓𝜈𝑓subscript𝑟𝜅𝑞subscript𝜓superscript𝜈superscript𝑓subscript𝑖𝑗superscriptsubscript𝐴𝑖𝜈𝑓superscriptsubscript𝐴𝑗superscript𝜈superscript𝑓superscriptsubscript𝑅𝑖𝑗𝜅quantum-operator-productsubscriptitalic-ϕ𝑖subscript𝑌𝜅𝑞subscriptitalic-ϕ𝑗\left\langle\psi_{\nu f}\right|r_{\kappa q}\left|\psi_{\nu^{\prime}f^{\prime}}% \right\rangle=\sum_{i,j}A_{i}^{(\nu f)}A_{j}^{(\nu^{\prime}f^{\prime})}R_{ij}^% {\kappa}\left\langle\phi_{i}\right|Y_{\kappa q}\left|\phi_{j}\right\rangle,⟨ italic_ψ start_POSTSUBSCRIPT italic_ν italic_f end_POSTSUBSCRIPT | italic_r start_POSTSUBSCRIPT italic_κ italic_q end_POSTSUBSCRIPT | italic_ψ start_POSTSUBSCRIPT italic_ν start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ⟩ = ∑ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_ν italic_f ) end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_ν start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) end_POSTSUPERSCRIPT italic_R start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT ⟨ italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_Y start_POSTSUBSCRIPT italic_κ italic_q end_POSTSUBSCRIPT | italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⟩ , (21)

with the spherical harmonics Yκqsubscript𝑌𝜅𝑞Y_{\kappa q}italic_Y start_POSTSUBSCRIPT italic_κ italic_q end_POSTSUBSCRIPT and the radial integrals

Rijκ=drrκPνir,i(r)Pνjr,j(r),superscriptsubscript𝑅𝑖𝑗𝜅differential-d𝑟superscript𝑟𝜅subscript𝑃subscript𝜈𝑖subscript𝑟𝑖𝑟subscript𝑃subscriptsuperscript𝜈𝑗subscriptsuperscript𝑟𝑗𝑟R_{ij}^{\kappa}=\int\mathrm{d}r\ r^{\kappa}P_{\nu_{i}\ell_{r,i}}(r)P_{\nu^{% \prime}_{j}\ell^{\prime}_{r,j}}(r),italic_R start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT = ∫ roman_d italic_r italic_r start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_ℓ start_POSTSUBSCRIPT italic_r , italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_r ) italic_P start_POSTSUBSCRIPT italic_ν start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r , italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_r ) , (22)

as well as the magnetic dipole moment

ψνf|𝝁|ψνf=i,jAi(νf)Aj(νf)Rij0ϕi|𝝁|ϕj.quantum-operator-productsubscript𝜓𝜈𝑓𝝁subscript𝜓superscript𝜈superscript𝑓subscript𝑖𝑗superscriptsubscript𝐴𝑖𝜈𝑓superscriptsubscript𝐴𝑗superscript𝜈superscript𝑓superscriptsubscript𝑅𝑖𝑗0quantum-operator-productsubscriptitalic-ϕ𝑖𝝁subscriptitalic-ϕ𝑗\left\langle\psi_{\nu f}\right|\bm{\mu}\left|\psi_{\nu^{\prime}f^{\prime}}% \right\rangle=\sum_{i,j}A_{i}^{(\nu f)}A_{j}^{(\nu^{\prime}f^{\prime})}R_{ij}^% {0}\left\langle\phi_{i}\right|\bm{\mu}\left|\phi_{j}\right\rangle.⟨ italic_ψ start_POSTSUBSCRIPT italic_ν italic_f end_POSTSUBSCRIPT | bold_italic_μ | italic_ψ start_POSTSUBSCRIPT italic_ν start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ⟩ = ∑ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_ν italic_f ) end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_ν start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) end_POSTSUPERSCRIPT italic_R start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ⟨ italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | bold_italic_μ | italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⟩ . (23)

This assumes the approximation that contributions from the dipole operators acting on the core are negligible compared to contributions from the Rydberg electron because the corresponding wave function differ vastly in size. Here, the representation in the fragmentation channel is beneficial: Rydberg and core quantum numbers are well separated therein.

II.3 Strontium 87

For the Rydberg structure of 87Sr, we use the experimentally reported quantum defects for 88Sr in the LS𝐿𝑆LSitalic_L italic_S frame as listed in [40]. We are interested in the 5sn5𝑠𝑛5sn\ell5 italic_s italic_n roman_ℓ Rydberg series close to the ionization threshold at relatively large principal quantum numbers n30greater-than-or-approximately-equals𝑛30n\gtrapprox 30italic_n ⪆ 30. In Sr, some doubly excited states of the 5pnp5𝑝𝑛𝑝5pnp5 italic_p italic_n italic_p-, 4dnd4𝑑𝑛𝑑4dnd4 italic_d italic_n italic_d-, and 4dnp4𝑑𝑛𝑝4dnp4 italic_d italic_n italic_p-series lie energetically below the first ionization threshold and are particularly relevant in order to accurately describe the spectrum at low n15𝑛15n\lessapprox 15italic_n ⪅ 15 [53]. However, for our purposes, the influence of these series is sufficiently captured by considering the quantum defects of the associated 5sn5𝑠𝑛5sn\ell5 italic_s italic_n roman_ℓ series as slowly varying functions of energy [40]. In the case of 87Sr+ with its fermionic core and neglecting electron excitations, i.e. c=0subscript𝑐0\ell_{c}=0roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 0, the finite nuclear spin I=92𝐼92I=\frac{9}{2}italic_I = divide start_ARG 9 end_ARG start_ARG 2 end_ARG couples to the electronic angular momentum jc=sc=12subscript𝑗𝑐subscript𝑠𝑐12j_{c}=s_{c}=\frac{1}{2}italic_j start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG giving rise to two core hyperfine configurations fcsubscript𝑓𝑐f_{c}italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT with energies separated by roughly 5 GHz. The energies of the ionic ground states with different hyperfine configurations serve as distinct ionization thresholds Efcsubscript𝐸subscript𝑓𝑐E_{f_{c}}italic_E start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUBSCRIPT for the Rydberg electron.

To visualize which pair-interaction potentials can possibly be excited via an intermediate level —c¿, we use a color code highlighting the potentials according to the dipole coupling of Rydberg-pair states |ΨketΨ\left|\Psi\right\rangle| roman_Ψ ⟩ to the pair state |c,tket𝑐𝑡\left|c,t\right\rangle| italic_c , italic_t ⟩, where |tket𝑡\left|t\right\rangle| italic_t ⟩ is the Rydberg target state to be specified and we have exemplary chosen |cket𝑐\left|c\right\rangle| italic_c ⟩ to be the P03superscriptsubscript𝑃03{}^{3}P_{0}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT f=92𝑓92f=\frac{9}{2}italic_f = divide start_ARG 9 end_ARG start_ARG 2 end_ARG clock state. Thus, the coloring corresponds to the matrix element D=|Ψ|𝒓11|c,t|𝐷quantum-operator-productΨtensor-product𝒓11𝑐𝑡D=|\left\langle\Psi\right|\bm{r}\otimes 1\!\!1\left|c,t\right\rangle|italic_D = | ⟨ roman_Ψ | bold_italic_r ⊗ 1 1 | italic_c , italic_t ⟩ |. Note that the state |c,tket𝑐𝑡\left|c,t\right\rangle| italic_c , italic_t ⟩ is not symmetrized, therefore the coloring is not sensitive to being in the dark- or bright-state sector of the Hamiltonian. In pure Born-Oppenheimer approximation, these sectors are well separated, however, considering motional degrees of freedom, there is some finite coupling. We consider this choice for coloration to be more instructive than coloring according to the overlap with a specified target state, because the latter neglects the impact of other dipole coupled states with for example differing hyperfine configuration or different orbital angular momentum. For e.g. blockade experiments, it is crucial to find an energy window as well as a region of internuclear distances, which is “clean”, i.e. free of dipole-coupled potentials. Unless otherwise specified, we choose for all pair-interaction potential figures in this article, D>105ea0𝐷superscript105𝑒subscript𝑎0D>10^{-5}\ ea_{0}italic_D > 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT italic_e italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for colored points and D>109ea0𝐷superscript109𝑒subscript𝑎0D>10^{-9}\ ea_{0}italic_D > 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT italic_e italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for gray, high-opacity points to inform about the encountered level densities.

II.4 Remarks on basis sets

A major challenge for the exact diagonalization of Rydberg pair Hamiltonians is the choice of the basis set and the corresponding convergence of eigenstates. For van-der-Waals type interaction, the pair states relevant to converge a given pair potential can be energetically distant, defining a requirement for the considered energy window for diagonalization. In general external field configurations, when M𝑀Mitalic_M is not a conserved quantity, the pair state level density becomes large and the number of required basis states is typically the limiting factor. Adding to this precarity, for multi-channel atoms, the single-atom level density is generally larger than e.g. in alkali atoms due to the relevance of the hyperfine interaction.

To illustrate this, let us consider the number of bound states within a single Rydberg manifold restricting the maximal orbital angular momentum to max=3subscriptmax3\ell_{\text{max}}=3roman_ℓ start_POSTSUBSCRIPT max end_POSTSUBSCRIPT = 3. In 87Rb, where I=32𝐼32I=\frac{3}{2}italic_I = divide start_ARG 3 end_ARG start_ARG 2 end_ARG, there are technically 32 levels with average degeneracy of 4 each. However, since the hyperfine interaction scales as n3superscript𝑛3n^{-3}italic_n start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, the nuclear spin can be disregarded for sufficiently large principal quantum numbers shrinking the number of levels to 7, or 32 in the case of resolved degeneracy. Contrary for 87Sr, hyperfine interaction cannot be disregarded222 In alkali metals, the hyperfine interaction scales with the principal quantum number due to the overlap of the electron wave function with the nuclear core. While in alkaline-earths the Rydberg electron is similarly disconnected from the nucleus, the second valence electron is always close to the core probing its structure. The electron-spin coupling of the two valence electrons prevents decoupling of the hyperfine structure and the corresponding scaling. and the number of levels is 64 with average degeneracy 10 each, due to the large nuclear spin I=92𝐼92I=\frac{9}{2}italic_I = divide start_ARG 9 end_ARG start_ARG 2 end_ARG. A factor of 20 in the single atom level density has devastating consequences for the diagonalization of pair states, where the number of states scales quadratically and the number of matrix elements to construct the Hamiltonian scales with the power of 4.

In order to achieve convergence on the sub-MHz scale for the figures in this article, the typical basis sizes range from 10k to 25k states, after symmetrization. For external fields not aligned with the z𝑧zitalic_z axis (not discussed here), we were only able to converge potentials to the tens-of-MHz scale using around 45k basis states. Unless otherwise specified, we take into account 3 to 4 neighboring n𝑛nitalic_n manifolds, i.e. 7 to 9 manifolds in total, and orbital angular momenta \ellroman_ℓ from 0 to 3 in the single atom basis. States further separated from the target state typically do not have the coupling strength to significantly contribute, although even energetically distant states could, in principle, make a degenerate pair.

For calculations with finite electric field, we compare the results including higher \ellroman_ℓ up to 4 or 5, however this does not serve as a proper convergence check in the case of high electric fields, because contributions from degenerate high-\ellroman_ℓ states could, in principle, add up to modify the level structure, hence, the focus on small electric fields, where we expect those contributions to be small. Combinations of such single atom eigenstates are considered in the pair basis, if their total energy is within a specified energy window of the target state. We found 20 GHz to be sufficient for our convergence goal for the considered states with n60𝑛60n\approx 60italic_n ≈ 60, however, in arbitrary external field configurations, this may lead to intractable basis sizes and the convergence criterion may need to be loosened.

II.5 Pairinteraction software

In order to account for the increased requirements associated with calculating pair-interaction potentials of multi-channel atoms discussed in the previous sections, we extended the pairinteraction software. A corresponding open-source release is currently pending.333 We refer to the software release and the accompanying publication for a more complete description of the improvements and the corresponding capabilities. The software will be available at https://pairinteraction.github.io/. Notably, the software supports sophisticated diagonalization routines such as the FEAST algorithm [54]. It is planned to offer cloud access to the pairinteraction software making use of CUDA for distributed computation on GPUs [55]. Furthermore, we propose a database standard for atomic state data. Such a database contains the bound state energies and appropriate state labeling according to the relevant quantum numbers as well as, importantly, the reduced matrix elements required for constructing Hamiltonians with high-performance [56].

Such a database separates the calculation of pair-interaction potentials from the more fundamental calculation of atomic spectra and, particularly, removes any requirement for the software to handle MQDT wave functions. Instead, we provide a Julia [57] implementation of MQDT using frame transformations following Robicheaux et al. [39] and outlined above as a utility tool for pairinteraction that will be available as open-source software. An MQDT spectrum is generally sensitive on the underlying quantum defect data and the considered channel couplings. Therefore any experimental advances in the spectroscopy of multi-channel atoms and improved efforts to fit quantum defects will, in principle, lead to new databases that can be used with the pairinteraction software.

III Results

In the following we will show that the multi-channel couplings stemming from the hyperfine splitting of the ionic core and resulting in singlet-triplet mixed Rydberg series extends the complexity and versatility of Rydberg pair interactions. Specifically, we focus on an intrinsic equal-state Förster resonance in the 5sns5𝑠𝑛𝑠5sns5 italic_s italic_n italic_s f=92𝑓92f=\frac{9}{2}italic_f = divide start_ARG 9 end_ARG start_ARG 2 end_ARG series of 87Sr. We can control the corresponding Rydberg-pair interactions in four ways, namely by the choice of the principal quantum number, by the magnitude of homogeneous magnetic and electric fields, and by the internuclear distance between atoms. In contrast, the 5sns5𝑠𝑛𝑠5sns5 italic_s italic_n italic_s f=112𝑓112f=\frac{11}{2}italic_f = divide start_ARG 11 end_ARG start_ARG 2 end_ARG series of 87Sr is effectively described in a single channel approximation and provides the features known from Rydberg states in alkali metals with properties regularly scaling with the principal quantum number.

III.1 Overview of the Förster resonance

Refer to caption
Figure 2: Transition energies to the 5snp5𝑠𝑛𝑝5snp5 italic_s italic_n italic_p manifold and reduced dipole matrix elements for the mixed singlet-triplet series 5sns5𝑠𝑛𝑠5sns5 italic_s italic_n italic_s f=92𝑓92f=\frac{9}{2}italic_f = divide start_ARG 9 end_ARG start_ARG 2 end_ARG. Each Rydberg manifold provides a total of ten different available states (colored points, some on top of each other), three of which are particularly strongly coupled. Rydberg scaling of energies and dipole elements is taken into account and the s𝑠sitalic_s state’s energy is set to zero for reference. (a) Each p𝑝pitalic_p series has a characteristic energy dependence. (b-c) Zooms into (a), the three strongly coupled states are equidistant from the s𝑠sitalic_s state around ν=60𝜈60\nu=60italic_ν = 60. Each combination of these series gives rise to a resonant pair with slightly different Förster defect leading to complex Rydberg-pair interaction behavior.

Fig. 2 illustrates the level structure and coupling elements of 87Sr relevant for the composition of Rydberg pair-interaction potentials as functions of the effective principal quantum number ν𝜈\nuitalic_ν. (For this series nν+3.5𝑛𝜈3.5n\approx\nu+3.5italic_n ≈ italic_ν + 3.5, where n𝑛nitalic_n counts the electronic orbital.) Accounting for the relevant scaling associated with these properties, we show the level separation of the 5sns5𝑠𝑛𝑠5sns5 italic_s italic_n italic_s f=92𝑓92f=\frac{9}{2}italic_f = divide start_ARG 9 end_ARG start_ARG 2 end_ARG series to its dipole-coupled states, i.e. states of 5snp5𝑠superscript𝑛𝑝5sn^{\prime}p5 italic_s italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_p f={72,92,112}superscript𝑓7292112f^{\prime}=\{\frac{7}{2},\frac{9}{2},\frac{11}{2}\}italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = { divide start_ARG 7 end_ARG start_ARG 2 end_ARG , divide start_ARG 9 end_ARG start_ARG 2 end_ARG , divide start_ARG 11 end_ARG start_ARG 2 end_ARG } character. Those states determine both the polarizabilty and the character of the pair-interaction potentials. Each Rydberg manifold provides in total ten relevant states (not all of which are resolved in the figure), that can be grouped according to their total electronic character S𝑆Sitalic_S, which can be mixed or pure. A subset of three states has very similar mixed singlet-triplet character to the 5sns5𝑠𝑛𝑠5sns5 italic_s italic_n italic_s f=92𝑓92f=\frac{9}{2}italic_f = divide start_ARG 9 end_ARG start_ARG 2 end_ARG series and their dipole coupling is correspondingly strong.

Importantly, the position of the states of neighboring manifolds within the spectrum relative to the s𝑠sitalic_s series slowly changes as a function of ν𝜈\nuitalic_ν and around ν61𝜈61\nu\approx 61italic_ν ≈ 61 the states of the neighboring manifolds are equidistant, cf. Fig. 2(b-c). In other words, the Förster defects, i.e. the energy differences, of pairs (n1)p+np𝑛1𝑝𝑛𝑝(n-1)p+np( italic_n - 1 ) italic_p + italic_n italic_p and ns+ns𝑛𝑠𝑛𝑠ns+nsitalic_n italic_s + italic_n italic_s have zero crossings giving rise to a broad Förster resonance. Additionally, the contribution of these transitions to the Stark effect cancel giving rise to near zero polarizabilty. Both of these effects have been predicted by perturbative considerations [40]. Fig. 2 reveals the complex interplay of the contributing states. The more weakly coupled states as seen in Fig. 2(a), which have strong triplet character, contribute to the overall structure and illustrate the significant energy dependence of quantum defects, which originates in the hyperfine-split ionization threshold of 87Sr+.

In the following, we analyze the conditions for this Förster resonance, indicating that it is intrinsic to an equal-state pair in the s𝑠sitalic_s series strongly coupling to the adjacent p𝑝pitalic_p levels at the zero crossing of the associated Förster defect, which is unique to specific Rydberg series in divalent atoms. In Rydberg series with slowly varying energy dependence of the quantum defects, for example the 5sns5𝑠𝑛𝑠5sns5 italic_s italic_n italic_s f=112𝑓112f=\frac{11}{2}italic_f = divide start_ARG 11 end_ARG start_ARG 2 end_ARG series in 87Sr, the functional behavior of the Förster defect features the regular n3superscript𝑛3n^{-3}italic_n start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT scaling. This scaling and the consequential absence of a zero crossing of the Förster defect prohibits the existence of a Förster resonance. Particularly, this scaling is generally present in alkali atoms, where Förster resonances are not typically found in equal-state pairs and rarely without additional tuning of energy levels with external fields [44]. Instead, the existence of a Förster resonance requires the analysis of many spectral properties and the careful tuning of the quantum numbers to different values for the involved atoms in order to achieve degeneracy of a pair [47]. Examples are p𝑝pitalic_p series in which the fine structure sufficiently shifts the state to be resonant with adjacent s𝑠sitalic_s states or an nsnd𝑛𝑠superscript𝑛𝑑nsn^{\prime}ditalic_n italic_s italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_d pair or inter-species resonances using different principal quantum numbers for each species [58].

In hyperfine-split ionic-core species, however, the existence of a mixed single-triplet s𝑠sitalic_s series featuring a Förster resonance is guaranteed, because the hyperfine interaction leads to an exchange in electronic character as a function of n𝑛nitalic_n, when the intrinsic splitting of adjacent Rydberg states is comparable to the ionic hyperfine splitting. For example, we expect such a resonance for the 6sns6𝑠𝑛𝑠6sns6 italic_s italic_n italic_s f=12𝑓12f=\frac{1}{2}italic_f = divide start_ARG 1 end_ARG start_ARG 2 end_ARG series of 171Yb, however, the uncertainty of quantum defects of the 5snp5𝑠𝑛𝑝5snp5 italic_s italic_n italic_p series makes a quantitative prediction of its location in the spectrum currently dispensable444 On the other hand, the quantum defects for the s𝑠sitalic_s and d𝑑ditalic_d series in Yb are rather well known. It is, however, inherently difficult to probe the p𝑝pitalic_p series in heavy divalent atoms, in which Rydberg spectroscopy is typically performed using two-photon transitions, because the single photon transition is too deeply ultra violet. We hope to see new Yb spectroscopy data soon or that this article can serve as a motivation for additional experimental efforts. .

The unique combination of the features makes the Förster resonant state an interesting candidate for quantum technology applications: The bare Rydberg line is largely insusceptible to external fields and their fluctuations. Beyond the small polarizability, the mixed electron-spin character of the state also leads to a suppressed linear Zeeman splitting, because singlet states have a vanishing g𝑔gitalic_g factor, and the response to magnetic field is mostly quadratic driven by diamagnetic interaction requiring large fields. At the same time, the pair state is subject to dipole-dipole interactions that provide a larger blockade radius than typically encountered for pair states subject to induced-dipole (i.e. van-der-Waals) interactions. And finally, the equal-state pair can be easily addressed in a global beam setup requiring only a single wavelength as opposed to Förster states involving pairs of states differing in n𝑛nitalic_n or \ellroman_ℓ or even interspecies resonances as recently demonstrated [59].

III.2 Field-free pair-interaction potentials

Refer to caption
Figure 3: Pair-interaction potentials for the 5sns5𝑠𝑛𝑠5sns5 italic_s italic_n italic_s f=92𝑓92f=\frac{9}{2}italic_f = divide start_ARG 9 end_ARG start_ARG 2 end_ARG equal-state pairs of 87Sr for different n={59,62,64,66,69}𝑛5962646669n=\{59,62,64,66,69\}italic_n = { 59 , 62 , 64 , 66 , 69 } (a-e). Varying n𝑛nitalic_n, the potentials change character from repulsive to attractive, while on resonance (c), the potentials are symmetric and asymptotically degenerate. However, many other pair states contribute to the potential energy landscape introducing avoided crossings and possible channels for population transfer. The color indicates the dipole transition strength from the P03superscriptsubscript𝑃03{}^{3}P_{0}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT clock state. Only the M=9𝑀9M=-9italic_M = - 9 symmetry sector is shown.

In Fig. 3 we show the Rydberg-pair interaction potentials for equal pairs in the 5sns5𝑠𝑛𝑠5sns5 italic_s italic_n italic_s f=92𝑓92f=\frac{9}{2}italic_f = divide start_ARG 9 end_ARG start_ARG 2 end_ARG series in the vicinity of the Förster resonance at n=64𝑛64n=64italic_n = 64. Here, the asymptotic energy of the 5sns+5sns5𝑠𝑛𝑠5𝑠𝑛𝑠5sns+5sns5 italic_s italic_n italic_s + 5 italic_s italic_n italic_s pair is set to zero and the coloring indicates dipolar coupling to the P03superscriptsubscript𝑃03{}^{3}P_{0}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT f=9/2𝑓92f=9/2italic_f = 9 / 2 clock state. Over a relatively short range of principal quantum numbers, n={59,62,64,66,69}𝑛5962646669n=\{59,62,64,66,69\}italic_n = { 59 , 62 , 64 , 66 , 69 } shown in subfigures (a) through (e), the potentials drastically change character. At low n𝑛nitalic_n far from resonance, the potential is repulsive and scales as R6superscript𝑅6R^{-6}italic_R start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT with the internuclear distance, indicating induced-dipolar interactions, and the pair interaction is well described using perturbation theory. Closer to resonance, the scaling changes to R3superscript𝑅3R^{-3}italic_R start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, indicating permanent-dipole interactions due to the mixing of pair states with s𝑠sitalic_s and p𝑝pitalic_p character. At n=64𝑛64n=64italic_n = 64, the s𝑠sitalic_s and p𝑝pitalic_p pairs are asymptotically degenerate and the molecular eigenstates are symmetric superpositions, which leads to symmetrically attractive and repulsive potentials as is characteristic for a Förster resonance. However, the fact that several pairs of p𝑝pitalic_p character contribute, leads to some non-generic level crossings and state mixings at large internuclear distances and at small detunings. Furthermore, the significant p𝑝pitalic_p state character of the potentials gives rise to yet more additional couplings to other pair states of either s𝑠sitalic_s or d𝑑ditalic_d character at larger energies (e.g. around E/h=150𝐸150E/h=150italic_E / italic_h = 150 MHz and E/h=250𝐸250E/h=250italic_E / italic_h = 250 MHz) that lead to avoided crossings. At larger n𝑛nitalic_n, the dominant potential remains attractive and the scaling transitions to dominantly R6superscript𝑅6R^{-6}italic_R start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT again. At short internuclear distances R<4𝑅4R<4italic_R < 4 µm, the potential is modified due to significant dipolar interactions stemming from other pair states at smaller energies. The larger n𝑛nitalic_n, the more likely such modifications become, due to the increasing level density and the increasing dipolar interaction strength.

III.3 Magnetic field response

Refer to caption
Figure 4: Magnetic field maps for Rydberg pairs in the vicinity of the Förster resonant pair. (a) Asymptotically, in the non-interacting limit, pair states are subject to vastly differing field responses all subject to diamagnetic interaction. Black indicates pairs of s𝑠sitalic_s states, blue pairs of p𝑝pitalic_p states, and gray additional pairs involving s𝑠sitalic_s, p𝑝pitalic_p, and d𝑑ditalic_d states. (b) At 4 µm internuclear distance and strong dipolar coupling, the varying Förster defects cause avoided crossings of the potentials. The colorcode in (b) indicates coupling strength to the clock state.

The multi-channel character of the different Rydberg series in 87Sr gives rise to diverse external field responses depending on the principal and orbital quantum numbers. High level densities and the requirement for magnetic field strengths sufficient to lift the degeneracy in the ground (S01superscriptsubscript𝑆01{}^{1}S_{0}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT italic_S start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) or clock (P03superscriptsubscript𝑃03{}^{3}P_{0}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) manifold, which is often the case for quantum technology applications, can lead to non-generic level shifts and avoided crossings in the Rydberg spectra of single atoms and pairs.

Fig. 4 shows magnetic field maps for the Förster resonant pair and surrounding states (a) asymptotically in the non-interacting limit and (b) at 4 µm internuclear spacing. Here, the magnetic field is aligned with the internuclear axis such that the total projection of the hyperfine state is still a conserved quantity and only the M=9𝑀9M=-9italic_M = - 9 symmetry sector is shown. For the 5s64s5𝑠64𝑠5s64s5 italic_s 64 italic_s equal-state pair, highlighted in black in Fig. 4(a), the linear magnetic field response is weak, due to a strong contribution from the singlet channel in that state, and the quadratic response due to diamagnetism dominates even at relatively small field strength. For such states that are field-seeking, i.e. they lower their energy with respect to the magnetic field strength due to Zeeman interaction, there exists a local minimum in the magnetic field map, where the pair state is insensitive to magnetic field fluctuations due to a balance between attractive Zeeman and repulsive diamagnetic interaction. For the state of interest this occurs at approximately B=50𝐵50B=50italic_B = 50 G.

Multiple pair states with p𝑝pitalic_p character, which are close in energy at zero field, are shifted to smaller energies due to their strong linear response, i.e. Zeeman effect. Note that most of the states shown in Fig. 4 are field seeking because of M𝑀Mitalic_M being negative. Only in the case of f<I𝑓𝐼f<Iitalic_f < italic_I, such states are shifted to larger energies. However, some pairs of p𝑝pitalic_p states, while initially being shifted out of resonance for small fields, become resonant with the s𝑠sitalic_s pair again at B>150𝐵150B>150italic_B > 150 G. This is striking and highlights the importance of diamagnetism: If only linear shifts were considered, all p𝑝pitalic_p pairs would be shifted out of resonance, which would drastically change the character of the pair interaction potentials.

At short internuclear distances in Fig. 4(b), dipole-dipole interactions are additionally relevant, and the diverse magnetic field responses of the pair states leads to non-trivial functional behavior of interaction potentials, with a multitude of avoided crossings, and of their dipole addressability. In particular, at B=200𝐵200B=200italic_B = 200 G, one of the p𝑝pitalic_p pairs initially resonant at zero field, becomes degenerate with a different pair state. This pair, which is detuned by approximately 200200-200- 200 MHz at zero field (cf. subfigure (a)), is from the same 5sns5𝑠𝑛𝑠5sns5 italic_s italic_n italic_s f=92𝑓92f=\frac{9}{2}italic_f = divide start_ARG 9 end_ARG start_ARG 2 end_ARG Rydberg series, however, for different n𝑛nitalic_n, namely n=n±2superscript𝑛plus-or-minus𝑛2n^{\prime}=n\pm 2italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_n ± 2, and one of the two states has slightly different singlet-triplet configuration due to its multi-channel nature. In a magnetic field, this pair remains approximately equidistant to the Förster resonant state. However, at B=200𝐵200B=200italic_B = 200 G, the two s𝑠sitalic_s pairs become strongly coupled: Essentially, two Förster resonances occur simultaneously at approximately 200 MHz energy splitting and the involved p𝑝pitalic_p states mediate coupling between the s𝑠sitalic_s pairs.

Refer to caption
Figure 5: Competition of Förster resonances between pairs 64s+64s64𝑠64𝑠64s+64s64 italic_s + 64 italic_s, 63p+64p63𝑝64𝑝63p+64p63 italic_p + 64 italic_p, and 62s+66s62𝑠66𝑠62s+66s62 italic_s + 66 italic_s. Their Förster defects depend on the magnetic field strength. (a) At B=50𝐵50B=50italic_B = 50 G, the 64s+64s64𝑠64𝑠64s+64s64 italic_s + 64 italic_s and several 63p+64p63𝑝64𝑝63p+64p63 italic_p + 64 italic_p pairs are close to resonance, while 62s+66s62𝑠66𝑠62s+66s62 italic_s + 66 italic_s is detuned by 200similar-toabsent200\sim 200∼ 200 MHz. (b) At B=200𝐵200B=200italic_B = 200 G, both s𝑠sitalic_s pairs are resonant with different p𝑝pitalic_p pairs while their respective splitting is approximately maintained. However, the p𝑝pitalic_p pairs mitigate a strong coupling between the s𝑠sitalic_s pairs, hence the large avoided crossing and dipole addressability.

The corresponding pair potentials are shown in Fig. 5 for B=50𝐵50B=50italic_B = 50 G (a) and B=200𝐵200B=200italic_B = 200 G (b). Dipolar interactions exceed the energy scale of the asymptotic splitting and the pair interaction potentials feature a broad avoided crossing (cf. Fig. 5(b)) giving rise to an essentially flat dipole-coupled potential over the full range of R𝑅Ritalic_R. While for blockade experiments, this situation needs to be avoided, such a potential is highly unusual and may lead to facilitation, where Rydberg pairs are addressed resonantly.

This example illustrates that the magnetic field can effectively act as a knob for competing Förster resonances. The potentials in Fig. 5 can be qualitatively reproduced by a toy model that only includes the four properly symmetrized pair states that contribute dominantly:

{|nsns,|(n1)pnp,|nsn′′s,|(n1)pnp},ket𝑛𝑠𝑛𝑠ket𝑛1𝑝𝑛𝑝ketsuperscript𝑛𝑠superscript𝑛′′𝑠ket𝑛1superscript𝑝𝑛superscript𝑝\{\left|nsns\right\rangle,\left|(n-1)pnp\right\rangle,\left|n^{\prime}sn^{% \prime\prime}s\right\rangle,\left|(n-1)p^{\prime}np^{\prime}\right\rangle\},{ | italic_n italic_s italic_n italic_s ⟩ , | ( italic_n - 1 ) italic_p italic_n italic_p ⟩ , | italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_s italic_n start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT italic_s ⟩ , | ( italic_n - 1 ) italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_n italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ } , (24)

where only s𝑠sitalic_s and p𝑝pitalic_p pairs are coupled, respectively, by dipolar interaction (of similar strength) and two different Förster defects control the asymptotic detunings as functions of the magnetic field.

III.4 Non-perturbative polarizability calculations

Refer to caption
Figure 6: Electric field response of 5s64s5𝑠64𝑠5s64s5 italic_s 64 italic_s equal-state pairs in different Rydberg series at internuclear spacing R=6𝑅6R=6italic_R = 6 µm. (a) The single channel f=112𝑓112f=\frac{11}{2}italic_f = divide start_ARG 11 end_ARG start_ARG 2 end_ARG state follows the expectation from perturbation theory with linear behavior of the electric field derivative. (b) For the multi-channel f=92𝑓92f=\frac{9}{2}italic_f = divide start_ARG 9 end_ARG start_ARG 2 end_ARG state, hyperpolarizability is relevant even at small electric fields, giving rise to configurations with zero polarizability (here: zero local slope). At resonance, the polarizability of the f=92𝑓92f=\frac{9}{2}italic_f = divide start_ARG 9 end_ARG start_ARG 2 end_ARG pairs is larger than for f=112𝑓112f=\frac{11}{2}italic_f = divide start_ARG 11 end_ARG start_ARG 2 end_ARG pairs because of significant p𝑝pitalic_p state admixture.

After considering the effects of external magnetic fields in the previous section, we now turn our attention towards electric fields. Generally, electric fields pose a computational challenge because contributions from high-orbital-angular-momentum states become relevant even for moderate field strengths. Additionally, the influence of electric fields steeply scales with the principal quantum number (n5similar-toabsentsuperscript𝑛5\sim n^{5}∼ italic_n start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT). Therefore, here we restrict our analysis to electric field strengths of few hundreds of mV/cm. We find that even these small electric fields can cause responses not expected from perturbation theory.

This is illustrated in Fig. 6 for R=6𝑅6R=6italic_R = 6 µm internuclear spacing and electric fields aligned with the internuclear axis. While pairs of the pure triplet state 5s64s5𝑠64𝑠5s64s5 italic_s 64 italic_s f=112𝑓112f=\frac{11}{2}italic_f = divide start_ARG 11 end_ARG start_ARG 2 end_ARG show the expected linear response of the derivative of its eigenenergy with respect to the electric field strength dE𝐸Eitalic_E/dF𝐹Fitalic_F as a function of the electric field strength F𝐹Fitalic_F, the behavior of the Förster resonant pairs 5s64s5𝑠64𝑠5s64s5 italic_s 64 italic_s f=92𝑓92f=\frac{9}{2}italic_f = divide start_ARG 9 end_ARG start_ARG 2 end_ARG indicates the presence of a hyperpolarizability that becomes relevant for relatively small field strengths F150similar-to𝐹150F\sim 150italic_F ∼ 150 mV/cm. In fact, this can be exploited to achieve extremely high degrees of line stability, when the electric field is tuned to a local extremum of dE𝐸Eitalic_E/dF𝐹Fitalic_F. Overall, the f=92𝑓92f=\frac{9}{2}italic_f = divide start_ARG 9 end_ARG start_ARG 2 end_ARG pairs show a much larger polarizability than the f=112𝑓112f=\frac{11}{2}italic_f = divide start_ARG 11 end_ARG start_ARG 2 end_ARG pairs. This can be attributed to the strong p𝑝pitalic_p-state admixture at resonance. Asymptotically, the pure 5s64s5𝑠64𝑠5s64s5 italic_s 64 italic_s f=92𝑓92f=\frac{9}{2}italic_f = divide start_ARG 9 end_ARG start_ARG 2 end_ARG has near zero polarizability, approximately 2 orders of magnitude smaller than the 5s64s5𝑠64𝑠5s64s5 italic_s 64 italic_s f=112𝑓112f=\frac{11}{2}italic_f = divide start_ARG 11 end_ARG start_ARG 2 end_ARG state.

III.5 Single channel Rydberg series

Refer to caption
Figure 7: Pair potentials for the 5s64s5𝑠64𝑠5s64s5 italic_s 64 italic_s f=112𝑓112f=\frac{11}{2}italic_f = divide start_ARG 11 end_ARG start_ARG 2 end_ARG equal-state pairs for finite magnetic fields B=0𝐵0B=0italic_B = 0 G (a) and B=200𝐵200B=200italic_B = 200 G (b). These potentials are spherically symmetric and the characterizing C6 coefficient is only about 5% smaller in (b) compared to (a).

As mentioned previously, the 5sns5𝑠𝑛𝑠5sns5 italic_s italic_n italic_s f=112𝑓112f=\frac{11}{2}italic_f = divide start_ARG 11 end_ARG start_ARG 2 end_ARG Rydberg series in 87Sr provides, in a lot of aspects, less complexity than the f=92𝑓92f=\frac{9}{2}italic_f = divide start_ARG 9 end_ARG start_ARG 2 end_ARG series. It is a pure triplet series, well described in a single channel approximation and, hence, shows a lot of the regular scaling behaviors with respect to the principal quantum number expected from alkali metals or other single-channel Rydberg series. In this series, an equal-state Förster resonance is not expected and the polarizability scales as n7superscript𝑛7n^{7}italic_n start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT, while induced-dipolar interaction scales as n11superscript𝑛11n^{11}italic_n start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT.

Fig. 7 shows the pair interaction potentials (a) at zero magnetic field and (b) at B=200𝐵200B=200italic_B = 200 G aligned with the internuclear axis. Here, the M=11𝑀11M=-11italic_M = - 11 symmetry sector is shown and the coupling to the clock state is achieved by σ+subscript𝜎\sigma_{+}italic_σ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT driving. Essentially, the magnetic field does not have a strong impact on the potential. At B=200𝐵200B=200italic_B = 200 G the C6 coefficient is reduced by approximately 5% and the spherically symmetric structure of the potential is maintained (not shown in the figure). However, to achieve, e.g., Rydberg blockade for pairs within that series, still care must be taken for the choice of the principal quantum number in order to avoid accidental resonances.

IV Conclusion and Outlook

We have analysed the structure of Rydberg pair-interaction potentials of 87Sr, a multi-channel, divalent atomic species. The ionic core’s hyperfine structure gives rise to intriguing features within the Rydberg series, particularly states with vanishing polarizability and resonant dipolar interaction. While these properties make those states very promising candidates for application in quantum technology, the arising complexity requires careful consideration of external field configurations. We have shown that magnetic fields can undermine Rydberg blockade because of competing Förster resonances that give rise to avoided crossings in the potential energy landscape and to unanticipated potential shapes, such as an almost completely flat potential with respect to the internuclear distance. On the other hand, configurations can be found that offer insensitivity to electric fields with pair states that have zero polarizability.

The results from pair-interaction calculations are readily available as input for error models and fidelity estimations in the context of entangling gates in quantum computation [60]. Specifically, the results provide quantitative information about leakage into doubly excited states. Depending on the available channels, this undesired behavior can potentially be mitigated by appropriate pulse shaping.

We have integrated the tools to analyze the spectra of multi-channel divalent atoms based on quantum defect measurements into the pairinteraction software. Due to the increased level density of fermionic species, the calculation of pair potentials in arbitrary external field configurations remains challenging, particularly for 87Sr due to its large nuclear spin.

Recently, 171Yb has emerged as an interesting candidate for application in quantum computation. With a nuclear spin of I=12𝐼12I=\frac{1}{2}italic_I = divide start_ARG 1 end_ARG start_ARG 2 end_ARG it is a natural choice for nuclear qubit encoding. Due to a similar multi-channel structure, we expect the Rydberg spectrum of 171Yb to share a lot of features with 87Sr. Quantitative predictions have been enabled by related work including spectroscopy of the P11superscriptsubscript𝑃11{}^{1}P_{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and PJ3superscriptsubscript𝑃𝐽3{}^{3}P_{J}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT italic_P start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT series in 171Yb  [61], which we became aware of during the compilation of this manuscript.

Acknowledgements.
S.W. and J.M. acknowledge funding by the Federal Ministry of Education and Research (BMBF) under the Grants QRydDemo, MUNIQC-Atoms, and by the Horizon Europe program HORIZON-CL4-2021-DIGITALEMERGING-01-30 via the project 101070144 (EuRyQa).

References

  • Pritchard et al. [2010] J. D. Pritchard, D. Maxwell, A. Gauguet, K. J. Weatherill, M. P. A. Jones, and C. S. Adams, Cooperative atom-light interaction in a blockaded Rydberg ensemble, Phys. Rev. Lett. 105, 193603 (2010).
  • Gorshkov et al. [2011] A. V. Gorshkov, J. Otterbach, M. Fleischhauer, T. Pohl, and M. D. Lukin, Photon-photon interactions via Rydberg blockade, Phys. Rev. Lett. 107, 133602 (2011).
  • Firstenberg et al. [2016] O. Firstenberg, C. S. Adams, and S. Hofferberth, Nonlinear quantum optics mediated by Rydberg interactions, J. Phys. B: At. Mol. Opt. Phys. 49, 152003 (2016).
  • Saffman et al. [2010] M. Saffman, T. G. Walker, and K. Mølmer, Quantum information with Rydberg atoms, Rev. Mod. Phys. 82, 2313 (2010).
  • Adams et al. [2019] C. S. Adams, J. D. Pritchard, and J. P. Shaffer, Rydberg atom quantum technologies, J. Phys. B: At. Mol. Opt. Phys. 53, 012002 (2019).
  • Morgado and Whitlock [2021] M. Morgado and S. Whitlock, Quantum simulation and computing with Rydberg-interacting qubits, AVS Quantum Science 3, 023501 (2021).
  • Endres et al. [2016] M. Endres, H. Bernien, A. Keesling, H. Levine, E. R. Anschuetz, A. Krajenbrink, C. Senko, V. Vuletic, M. Greiner, and M. D. Lukin, Atom-by-atom assembly of defect-free one-dimensional cold atom arrays, Science 354, 1024 (2016).
  • Manetsch et al. [2024] H. J. Manetsch, G. Nomura, E. Bataille, K. H. Leung, X. Lv, and M. Endres, A tweezer array with 6100 highly coherent atomic qubits (2024), arXiv:2403.12021 [quant-ph] .
  • Barredo et al. [2016] D. Barredo, S. de Léséleuc, V. Lienhard, T. Lahaye, and A. Browaeys, An atom-by-atom assembler of defect-free arbitrary two-dimensional atomic arrays, Science 354, 1021 (2016).
  • Kim et al. [2016] H. Kim, W. Lee, H.-g. Lee, H. Jo, Y. Song, and J. Ahn, In situ single-atom array synthesis using dynamic holographic optical tweezers, Nature Commun. 7, 13317 (2016).
  • Bernien et al. [2017] H. Bernien, S. Schwartz, A. Keesling, H. Levine, A. Omran, H. Pichler, S. Choi, A. S. Zibrov, M. Endres, M. Greiner, et al., Probing many-body dynamics on a 51-atom quantum simulator, Nature 551, 579 (2017).
  • Ebadi et al. [2021] S. Ebadi, T. T. Wang, H. Levine, A. Keesling, G. Semeghini, A. Omran, D. Bluvstein, R. Samajdar, H. Pichler, W. W. Ho, et al., Quantum phases of matter on a 256-atom programmable quantum simulator, Nature 595, 227 (2021).
  • Ma et al. [2023] S. Ma, G. Liu, P. Peng, B. Zhang, S. Jandura, J. Claes, A. P. Burgers, G. Pupillo, S. Puri, and J. D. Thompson, High-fidelity gates and mid-circuit erasure conversion in an atomic qubit, Nature 622, 279 (2023).
  • Norcia et al. [2023] M. A. Norcia, W. B. Cairncross, K. Barnes, P. Battaglino, A. Brown, M. O. Brown, K. Cassella, C.-A. Chen, R. Coxe, D. Crow, et al., Midcircuit qubit measurement and rearrangement in a Yb171superscriptYb171{}^{171}\mathrm{Yb}start_FLOATSUPERSCRIPT 171 end_FLOATSUPERSCRIPT roman_Yb atomic array, Phys. Rev. X 13, 041034 (2023).
  • Tao et al. [2024] R. Tao, M. Ammenwerth, F. Gyger, I. Bloch, and J. Zeiher, High-fidelity detection of large-scale atom arrays in an optical lattice, Phys. Rev. Lett. 133, 013401 (2024).
  • Norcia et al. [2024] M. A. Norcia, H. Kim, W. B. Cairncross, M. Stone, A. Ryou, M. Jaffe, M. O. Brown, K. Barnes, P. Battaglino, T. C. Bohdanowicz, et al., Iterative assembly of 171YbYb\mathrm{Yb}roman_Yb atom arrays with cavity-enhanced optical lattices, PRX Quantum 5, 030316 (2024).
  • Bluvstein et al. [2022] D. Bluvstein, H. Levine, G. Semeghini, T. T. Wang, S. Ebadi, M. Kalinowski, A. Keesling, N. Maskara, H. Pichler, M. Greiner, et al., A quantum processor based on coherent transport of entangled atom arrays, Nature 604, 451 (2022).
  • Browaeys et al. [2016] A. Browaeys, D. Barredo, and T. Lahaye, Experimental investigations of dipole–dipole interactions between a few Rydberg atoms, J. Phys. B: At. Mol. Opt. Phys. 49, 152001 (2016).
  • Jenkins et al. [2022] A. Jenkins, J. W. Lis, A. Senoo, W. F. McGrew, and A. M. Kaufman, Ytterbium nuclear-spin qubits in an optical tweezer array, Phys. Rev. X 12, 021027 (2022).
  • Chen et al. [2022] N. Chen, L. Li, W. Huie, M. Zhao, I. Vetter, C. H. Greene, and J. P. Covey, Analyzing the Rydberg-based optical-metastable-ground architecture for Yb171superscriptYb171{}^{171}\mathrm{Yb}start_FLOATSUPERSCRIPT 171 end_FLOATSUPERSCRIPT roman_Yb nuclear spins, Phys. Rev. A 105, 052438 (2022).
  • Unnikrishnan et al. [2024] G. Unnikrishnan, P. Ilzhöfer, A. Scholz, C. Hölzl, A. Götzelmann, R. K. Gupta, J. Zhao, J. Krauter, S. Weber, N. Makki, H. P. Büchler, T. Pfau, and F. Meinert, Coherent control of the fine-structure qubit in a single alkaline-earth atom, Phys. Rev. Lett. 132, 150606 (2024).
  • Cao et al. [2024] A. Cao, W. J. Eckner, T. L. Yelin, A. W. Young, S. Jandura, L. Yan, K. Kim, G. Pupillo, J. Ye, N. D. Oppong, and A. M. Kaufman, Multi-qubit gates and ’Schrödinger cat’ states in an optical clock (2024), arXiv:2402.16289 [quant-ph] .
  • Dunning et al. [2016] F. B. Dunning, T. C. Killian, S. Yoshida, and J. Burgdörfer, Recent advances in Rydberg physics using alkaline-earth atoms, J. Phys. B: At. Mol. Opt. Phys. 49, 112003 (2016).
  • Wilson et al. [2022] J. T. Wilson, S. Saskin, Y. Meng, S. Ma, R. Dilip, A. P. Burgers, and J. D. Thompson, Trapping alkaline earth Rydberg atoms optical tweezer arrays, Phys. Rev. Lett. 128, 033201 (2022).
  • Burgers et al. [2022] A. P. Burgers, S. Ma, S. Saskin, J. Wilson, M. A. Alarcón, C. H. Greene, and J. D. Thompson, Controlling Rydberg excitations using ion-core transitions in alkaline-earth atom-tweezer arrays, PRX Quantum 3, 020326 (2022).
  • Pham et al. [2022] K.-L. Pham, T. F. Gallagher, P. Pillet, S. Lepoutre, and P. Cheinet, Coherent light shift on alkaline-earth Rydberg atoms from Isolated Core Excitation without autoionization, PRX Quantum 3, 020327 (2022).
  • Barnes et al. [2022] K. Barnes, P. Battaglino, B. J. Bloom, K. Cassella, R. Coxe, N. Crisosto, J. P. King, S. S. Kondov, K. Kotru, S. C. Larsen, et al., Assembly and coherent control of a register of nuclear spin qubits, Nature Commun. 13, 2779 (2022).
  • Jaksch et al. [2000] D. Jaksch, J. I. Cirac, P. Zoller, S. L. Rolston, R. Côté, and M. D. Lukin, Fast quantum gates for neutral atoms, Phys. Rev. Lett. 85, 2208 (2000).
  • Lukin et al. [2001] M. D. Lukin, M. Fleischhauer, R. Cote, L. M. Duan, D. Jaksch, J. I. Cirac, and P. Zoller, Dipole blockade and quantum information processing in mesoscopic atomic ensembles, Phys. Rev. Lett. 87, 037901 (2001).
  • Levine et al. [2019] H. Levine, A. Keesling, G. Semeghini, A. Omran, T. T. Wang, S. Ebadi, H. Bernien, M. Greiner, V. Vuletić, H. Pichler, and M. D. Lukin, Parallel implementation of high-fidelity multiqubit gates with neutral atoms, Phys. Rev. Lett. 123, 170503 (2019).
  • Jandura and Pupillo [2022] S. Jandura and G. Pupillo, Time-Optimal Two- and Three-Qubit Gates for Rydberg Atoms, Quantum 6, 712 (2022).
  • Bouchoule and Mølmer [2002] I. Bouchoule and K. Mølmer, Spin squeezing of atoms by the dipole interaction in virtually excited Rydberg states, Phys. Rev. A 65, 041803 (2002).
  • Johnson and Rolston [2010] J. E. Johnson and S. L. Rolston, Interactions between rydberg-dressed atoms, Phys. Rev. A 82, 033412 (2010).
  • Pupillo et al. [2010] G. Pupillo, A. Micheli, M. Boninsegni, I. Lesanovsky, and P. Zoller, Strongly correlated gases of rydberg-dressed atoms: Quantum and classical dynamics, Phys. Rev. Lett. 104, 223002 (2010).
  • Mitra et al. [2020] A. Mitra, M. J. Martin, G. W. Biedermann, A. M. Marino, P. M. Poggi, and I. H. Deutsch, Robust Mølmer-Sørensen gate for neutral atoms using rapid adiabatic Rydberg dressing, Phys. Rev. A 101, 030301 (2020).
  • de Léséleuc et al. [2018] S. de Léséleuc, S. Weber, V. Lienhard, D. Barredo, H. P. Büchler, T. Lahaye, and A. Browaeys, Accurate mapping of multilevel Rydberg atoms on interacting spin-1/2121/21 / 2 particles for the quantum simulation of Ising models, Phys. Rev. Lett. 120, 113602 (2018).
  • Weber et al. [2018] S. Weber, S. de Léséleuc, V. Lienhard, D. Barredo, T. Lahaye, A. Browaeys, and H. P. Büchler, Topologically protected edge states in small Rydberg systems, Quantum Sci. Technol. 3, 044001 (2018).
  • Aymar et al. [1996] M. Aymar, C. H. Greene, and E. Luc-Koenig, Multichannel Rydberg spectroscopy of complex atoms, Rev. Mod. Phys. 68, 1015 (1996).
  • Robicheaux et al. [2018] F. Robicheaux, D. W. Booth, and M. Saffman, Theory of long-range interactions for Rydberg states attached to hyperfine-split cores, Phys. Rev. A 97, 022508 (2018).
  • Robicheaux [2019] F. Robicheaux, Calculations of long range interactions for 87Sr Rydberg states, J. Phys. B: At. Mol. Opt. Phys. 52, 244001 (2019).
  • Liao et al. [1980] P. Liao, R. Freeman, R. Panock, and L. Humphrey, Hyperfine-induced singlet-triplet mixing in 3He, Optics Commun. 34, 195 (1980).
  • Beigang et al. [1981] R. Beigang, E. Matthias, and A. Timmermann, Influence of singlet-triplet mixing on the hyperfine structure of 5snd5𝑠𝑛𝑑5snd5 italic_s italic_n italic_d Rydberg states in Sr87superscriptSr87{}^{87}\mathrm{Sr}start_FLOATSUPERSCRIPT 87 end_FLOATSUPERSCRIPT roman_SrPhys. Rev. Lett. 47, 326 (1981).
  • Petrosyan et al. [2017] D. Petrosyan, F. Motzoi, M. Saffman, and K. Mølmer, High-fidelity Rydberg quantum gate via a two-atom dark state, Phys. Rev. A 96, 042306 (2017).
  • Walker and Saffman [2008] T. G. Walker and M. Saffman, Consequences of Zeeman degeneracy for the van der Waals blockade between Rydberg atoms, Phys. Rev. A 77, 032723 (2008).
  • Ryabtsev et al. [2010] I. I. Ryabtsev, D. B. Tretyakov, I. I. Beterov, and V. M. Entin, Observation of the Stark-tuned Förster resonance between two Rydberg atoms, Phys. Rev. Lett. 104, 073003 (2010).
  • Tretyakov et al. [2014] D. B. Tretyakov, V. M. Entin, E. A. Yakshina, I. I. Beterov, C. Andreeva, and I. I. Ryabtsev, Controlling the interactions of a few cold Rb Rydberg atoms by radio-frequency-assisted Förster resonances, Phys. Rev. A 90, 041403 (2014).
  • Beterov and Saffman [2015] I. I. Beterov and M. Saffman, Rydberg blockade, Förster resonances, and quantum state measurements with different atomic species, Phys. Rev. A 92, 042710 (2015).
  • Paris-Mandoki et al. [2016] A. Paris-Mandoki, H. Gorniaczyk, C. Tresp, I. Mirgorodskiy, and S. Hofferberth, Tailoring Rydberg interactions via Förster resonances: state combinations, hopping and angular dependence, J. Phys. B: At. Mol. Opt. Phys. 49, 164001 (2016).
  • Weber et al. [2017] S. Weber, C. Tresp, H. Menke, A. Urvoy, O. Firstenberg, H. P. Büchler, and S. Hofferberth, Tutorial: Calculation of Rydberg interaction potentials, J. Phys. B: At. Mol. Opt. Phys. 50, 133001 (2017).
  • Seaton [1966] M. J. Seaton, Quantum defect theory I. General formulation, Proc. Phys. Soc. 88, 801 (1966).
  • Fano [1970] U. Fano, Quantum defect theory of l𝑙litalic_l uncoupling in H2subscriptH2{\mathrm{H}}_{2}roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT as an example of channel-interaction treatment, Phys. Rev. A 2, 353 (1970).
  • Robaux and Aymar [1982] O. Robaux and M. Aymar, A program for analysing the Rydberg series of highly excited discrete spectra by M.Q.D.T., Comp. Phys. Comm. 25, 223 (1982).
  • Vaillant et al. [2014] C. L. Vaillant, M. P. A. Jones, and R. M. Potvliege, Multichannel quantum defect theory of strontium bound Rydberg states, J. Phys. B: At. Mol. Opt. Phys. 47, 155001 (2014).
  • Polizzi [2009] E. Polizzi, Density-matrix-based algorithm for solving eigenvalue problems, Phys. Rev. B 79, 115112 (2009).
  • Nvidia [2024] Nvidia, CUDA toolkit (2024).
  • Foundation [2024] D. Foundation, DuckDB (2024).
  • Bezanson et al. [2017] J. Bezanson, A. Edelman, S. Karpinski, and V. B. Shah, Julia: A fresh approach to numerical computing, SIAM Review 59, 65 (2017).
  • Ireland et al. [2024] P. M. Ireland, D. M. Walker, and J. D. Pritchard, Interspecies Förster resonances for Rb-Cs Rydberg d𝑑ditalic_d-states for enhanced multi-qubit gate fidelities, Phys. Rev. Res. 6, 013293 (2024).
  • Anand et al. [2024] S. Anand, C. E. Bradley, R. White, V. Ramesh, K. Singh, and H. Bernien, A dual-species Rydberg array (2024), arXiv:2401.10325 [quant-ph] .
  • Pagano et al. [2022] A. Pagano, S. Weber, D. Jaschke, T. Pfau, F. Meinert, S. Montangero, and H. P. Büchler, Error budgeting for a controlled-phase gate with strontium-88 Rydberg atoms, Phys. Rev. Res. 4, 033019 (2022).
  • Peper et al. [2024] M. Peper, Y. Li, D. Y. Knapp, M. Bileska, S. Ma, G. Liu, P. Peng, B. Zhang, S. P. Horvath, A. P. Burgers, and J. D. Thompson, Spectroscopy and modeling of 171Yb Rydberg states for high-fidelity two-qubit gates (2024), arXiv:2406.01482 [physics.atom-ph] .