Fact-checked by Grok 2 weeks ago

Rydberg atom

A Rydberg atom is an excited atom with one or more electrons promoted to a state characterized by a very high , n \gg 1, resulting in exaggerated atomic properties such as large orbital radii scaling as n^2, long radiative lifetimes scaling as n^3, and enhanced sensitivity to external fields. These atoms exhibit strong long-range dipole-dipole interactions that scale as n^4, enabling phenomena like the Rydberg blockade, where excitation of one atom prevents nearby atoms from being excited due to energy shifts. The term "Rydberg atom" honors Swedish physicist , who in developed a formula describing the wavelengths of atomic spectral lines, laying the groundwork for understanding highly excited states. Modern studies of Rydberg atoms gained momentum in the 1970s with the advent of tunable lasers, allowing precise excitation and spectroscopy of these states, as reviewed in early works on atomic physics. Key properties include high polarizabilities scaling as n^7, making them extremely responsive to electric fields, and the formation of exotic Rydberg molecules bound by interactions between the Rydberg electron and ground-state atoms. In cold atomic ensembles, Rydberg atoms facilitate collective effects, such as superradiance and spin squeezing, due to their large interaction volumes. Rydberg atoms have become pivotal in quantum technologies, particularly for processing, where the mechanism enables high-fidelity two-qubit gates like CNOT operations with error rates below 10^{-3}. They are also employed in precision sensing of electromagnetic fields, achieving sensitivities down to microvolts per centimeter, and in simulating many-body for studying transitions. Ongoing research explores their use in scalable quantum networks and neutral-atom quantum computers, leveraging advances in optical trapping and state control.

Definition and Formulation

Basic Definition and Properties

A Rydberg atom is defined as a neutral atom in which a single is excited to a state characterized by a large n \gg 1. This excitation results in exaggerated atomic properties compared to ground-state atoms, primarily due to the highly delocalized nature of the Rydberg electron's orbit. The orbital radius scales as n^2 a_0, where a_0 is the , leading to atomic sizes on the order of micrometers for n \approx 50. Additionally, the radiative lifetime of these states scales as n^3, often exceeding microseconds, which enhances their utility in precision measurements. The electronic states of Rydberg atoms are described by the standard atomic s: the principal n, orbital quantum number l (with $0 \leq l < n), magnetic quantum number m_l (with -l \leq m_l \leq l), and spin projection m_s = \pm 1/2. These atoms exhibit key properties such as transition dipole moments that scale as n^4, enabling strong long-range interactions, and scalar polarizabilities that scale as n^7, making them highly susceptible to external electric fields. At large n, Rydberg atoms display semi-classical behavior, where quantum wavefunctions approximate classical electron orbits, bridging microscopic quantum mechanics and macroscopic classical dynamics. Their sensitivity to perturbations, such as blackbody radiation or stray fields, arises from the small energy spacing between nearby levels, which decreases as $1/n^3. The energy levels of Rydberg states follow the Rydberg formula, adapted from hydrogen-like atoms. For hydrogen, the binding energy is E_n = -13.6 \, \mathrm{eV} / n^2. In multi-electron atoms like hydrogen-like ions, this becomes E_n = -13.6 \, \mathrm{eV} \cdot Z^2 / n^2, where Z is the nuclear charge. For alkali atoms, such as rubidium or cesium, the core electrons introduce a quantum defect \delta_{n,l}, modifying the effective principal quantum number to n^* = n - \delta_{n,l}, so the energy is E_{n,l} = -\mathrm{Ry} / (n - \delta_{n,l})^2, where Ry is the Rydberg constant (approximately 13.6 eV); this defect accounts for penetration into the ionic core and is most significant for low l. These scaling laws underscore the tunable nature of Rydberg properties with n.

Historical Development

The discovery of Rydberg atoms traces back to the late 19th century through spectroscopic observations of alkali metals. In the 1880s, Swedish physicist analyzed series spectra from elements like sodium and potassium, identifying regular patterns in emission lines that converged to ionization limits. His empirical formula, relating wavenumbers to principal quantum numbers, introduced the R \approx 109737 cm⁻¹, which quantified the spacing of these high-lying energy levels and laid the groundwork for understanding highly excited atomic states. In the mid-20th century, advancements in experimental spectroscopy revealed more about these high-principal-quantum-number (n) states. During the 1950s, Michael Seaton developed classical theories to describe autoionization processes in Rydberg series, treating the decay of doubly excited states above the ionization threshold using semiclassical approximations for electron correlations. This work provided essential insights into the stability and broadening of high-n levels in complex atoms. By the 1960s, W.R.S. Garton and F.S. Tomkins advanced detection techniques, identifying high-n Rydberg states in through high-resolution electron-impact spectra, which uncovered unexpected structures like quasi-Landau resonances near the ionization limit. Key theoretical contributions in the 1930s also shaped early understanding of Rydberg dynamics. In 1934, estimated the lifetimes of high-lying states in alkali atoms, accounting for collisional broadening in dense vapors through a statistical model that predicted linewidths scaling with density and state energy, influencing subsequent studies of radiative and non-radiative decay. The 1970s marked a pivotal shift with the advent of tunable lasers, enabling selective optical excitation of individual Rydberg levels for the first time, which allowed precise measurements of state properties and interactions previously inaccessible via discharge lamps. By the 1980s, Rydberg atoms emerged as model systems for exploring quantum chaos, driven by their sensitivity to external fields and large orbital radii. Studies of microwave ionization and level statistics in strong magnetic fields demonstrated signatures of classical chaos in quantum spectra, such as level repulsion and scarring, bridging semiclassical theory with experimental observations. This period solidified Rydberg systems as paradigms for non-integrable dynamics. Entering the 2020s, Rydberg atoms have evolved into platforms for quantum simulation, leveraging their strong, tunable interactions to emulate many-body phenomena like and quantum phase transitions in controlled arrays. In 2025, notable advances include the development of a highly accurate Rydberg-based thermometer by and enhanced use of Rydberg atom arrays for simulating molecular dynamics.

Theoretical Framework

Hydrogenic Potential

The hydrogenic potential governs the behavior of Rydberg atoms, approximating the interaction between the highly excited valence electron and the ionic core as a Coulomb attraction. For a hydrogen atom, the potential takes the form V(r) = -\frac{e^2}{4\pi\epsilon_0 r}, where Z = 1, e is the elementary charge, \epsilon_0 is the vacuum permittivity, and r is the radial distance from the nucleus. In multi-electron atoms such as alkali metals, the inner electrons screen the nuclear charge, resulting in an effective Z \approx 1 for the Rydberg electron in high-n states, as the core acts like a point charge of +1. To describe the quantum states, the time-independent Schrödinger equation is solved for the reduced mass system, where the reduced mass \mu accounts for the finite nuclear mass but approximates the electron mass m_e for heavy atoms. The equation separates into angular and radial parts due to spherical symmetry, yielding the radial Schrödinger equation for the reduced radial wave function u(r) = r R(r): -\frac{\hbar^2}{2\mu} \frac{d^2 u}{dr^2} + \left[ V(r) + \frac{\hbar^2 l(l+1)}{2\mu r^2} \right] u(r) = E u(r), where l is the orbital angular momentum quantum number, and the centrifugal term \frac{\hbar^2 l(l+1)}{2\mu r^2} arises from the angular momentum. This setup captures the hydrogen-like orbitals central to Rydberg states, with solutions depending on the principal quantum number n > l. In high-n Rydberg states, the hydrogenic reveals characteristic scaling behaviors that amplify atomic properties. The average orbital radius scales as \langle r \rangle \propto n^2 [a_0](/page/Bohr_radius), where a_0 is the , leading to orbits extending to micrometer scales for n \sim 100. The levels are E_n = -\frac{13.6 \, \mathrm{[eV](/page/EV)}}{n^2} (in the nuclear limit), so the spacing between adjacent levels \Delta E \approx \frac{27.2 \, \mathrm{[eV](/page/EV)}}{n^3} decreases rapidly, enabling fine control via external fields. These scalings highlight why Rydberg states exhibit exaggerated responses compared to ground-state atoms. Deviations from the pure hydrogenic model occur in alkali Rydberg atoms due to the valence electron's penetration into the ionic core, parameterized by the orbital-dependent quantum defect \delta_l. This modifies the energy levels to E_n = -\frac{13.6 \, \mathrm{eV}}{(n - \delta_l)^2}, where \delta_l is significant for low l (e.g., \delta_0 \approx 3 for s-states) but approaches zero for high l, restoring hydrogen-like behavior. The quantum defect accounts for core polarization and exchange effects without altering the long-range form.

Quantum-Mechanical Description

The quantum-mechanical treatment of Rydberg atoms focuses on the dynamics of the highly excited in the field of the ionic core, described by the time-independent H \psi = E \psi. The non-relativistic for this single active approximation is H = \frac{\mathbf{p}^2}{2\mu} + V(r), where \mu is the of the -core system, \mathbf{p} is the , and V(r) is the effective radial potential that asymptotically approaches the form -Z/r (in , with nuclear charge Z) for large interparticle separations r. For multi-electron atoms such as metals, V(r) incorporates short-range core effects via quantum defects, but the long-range behavior remains hydrogenic. Relativistic corrections, including spin-orbit coupling H_{\rm SO} = \xi(r) \mathbf{L} \cdot \mathbf{S} (with radial function \xi(r) \propto 1/r^3), are treated perturbatively for high principal quantum numbers n. Due to the spherical of the central potential V(r), both (even or odd under spatial inversion) and total \mathbf{J} = \mathbf{L} + \mathbf{S} (or \mathbf{L} in the absence of effects) are conserved quantum numbers, leading to good quantum numbers l (orbital ) and m_l (its projection). Electric dipole transitions between Rydberg states, mediated by the Hamiltonian H' = - \mathbf{d} \cdot \mathbf{E} (with dipole operator \mathbf{d} = -e \mathbf{r}), obey strict selection rules derived from the matrix elements \langle \psi_f | \mathbf{r} | \psi_i \rangle \neq 0: \Delta l = \pm 1, \Delta m_l = 0, \pm 1, and \Delta n arbitrary for large n. These rules arise from the vector nature of the dipole operator under rotations and parity transformations. Unlike low-lying states, the allowance of arbitrary \Delta n connects a given initial state to numerous nearby Rydberg levels, resulting in broad linewidths and absorption spectra that span wide frequency ranges, often on the order of GHz for n \sim 50. For sufficiently high n (typically n \gtrsim 30), the correspondence principle bridges quantum and classical descriptions: the discrete quantum states with definite n, l, and eccentricity-like quantum numbers map onto classical Keplerian elliptical orbits, where the electron's radial and angular motion precesses slowly compared to the orbital period \propto n^3. This semi-classical limit is evident in the scaling of expectation values, such as the orbital radius \langle r \rangle \propto n^2, aligning quantum probability distributions with classical trajectory densities. Fine-structure effects, encompassing spin-orbit coupling, relativistic kinetic energy corrections, and the Darwin term, introduce splittings that scale universally as $1/n^3 relative to the gross Rydberg energy E_n \propto -1/n^2, yielding shifts on the MHz scale for n \sim 100. Additionally, the Lamb shift—a quantum electrodynamic vacuum fluctuation effect—contributes a comparable $1/n^3 correction to s-states and nearby levels, becoming measurable in precision spectroscopy of high-n Rydberg states and lifting residual degeneracies beyond Dirac theory.

Electron Wavefunctions

The electron wavefunction for a Rydberg atom in a state characterized by principal quantum number n, orbital angular momentum quantum number l, and magnetic quantum number m is expressed in spherical coordinates as \psi_{nlm}(r, \theta, \phi) = R_{nl}(r) Y_{lm}(\theta, \phi), where R_{nl}(r) is the radial part and Y_{lm}(\theta, \phi) are the . This separable form arises from the quantum-mechanical solution for highly excited states, which are nearly hydrogenic due to the large orbital radius of the Rydberg . The radial wavefunction R_{nl}(r) for these hydrogen-like states takes the form R_{nl}(r) \propto \rho^l e^{-\rho/2} L_{n-l-1}^{2l+1}(\rho), where \rho = 2r / (n a_0) with a_0 the Bohr radius, and L_{n-l-1}^{2l+1}(\rho) denotes the associated Laguerre polynomial of degree n-l-1. The full normalized expression includes a prefactor ensuring \int_0^\infty r^2 |R_{nl}(r)|^2 dr = 1, but the proportional form highlights the polynomial structure that governs the oscillatory behavior. This radial function features exactly n - l - 1 nodes, corresponding to regions of zero probability density along the radial direction, which increase with n for fixed l and reflect the classical turning points of the electron's orbit. The angular dependence is provided by the spherical harmonics Y_{lm}(\theta, \phi), which determine the orbital shape and orientation, such as the azimuthal symmetry for m = 0 or the toroidal distribution for higher |m|. These functions are independent of n and ensure the total angular momentum quantization, with l ranging from 0 to n-1. The probability density |\psi_{nlm}|^2 = |R_{nl}(r)|^2 |Y_{lm}(\theta, \phi)|^2 thus concentrates most of the electron's probability at large radii, with the radial maximum occurring near r \approx n^2 a_0, scaling quadratically with n and underscoring the extended spatial extent of Rydberg orbitals. In non-hydrogenic Rydberg atoms, such as those in alkali metals, the quantum defect \delta_l introduces perturbations to this ideal form, particularly for low l where the wavefunction penetrates the ionic core. This penetration modifies the effective principal quantum number to n^* = n - \delta_l, shifting the energy levels inward and altering the inner region's amplitude, which enhances interactions with the core electrons and influences ionization thresholds. For high l (near-circular orbits), \delta_l is small, preserving the hydrogenic character, whereas for l = 0 (s-states), significant penetration leads to larger \delta_l and greater deviation from pure Coulomb behavior.

Production Techniques

Electron Impact Excitation

Electron impact excitation produces Rydberg atoms through inelastic scattering processes in which an incident collides with a ground-state atom and transfers energy to one of its bound , elevating it to a Rydberg state with high n. The energy threshold for this corresponds to the difference between the ground-state energy and the Rydberg level energy, which decreases toward the atomic ionization potential as n increases, allowing excitation to very high n states near the ionization threshold. Cross sections for electron impact excitation to high Rydberg states are typically modest for the initial promotion from the ground state, with measurements in helium yielding an absolute cross section of (9 ± 5) × 10^{-17} / n^3 cm² for populating states with principal quantum number n using 100 eV incident electrons. The total for the high-n manifold (n ≥ 15) is on the order of 10^{-18} cm². These cross sections peak near the excitation threshold and exhibit an n distribution that maximizes around n \approx 25 under these conditions, reflecting the kinematics of energy transfer in the collision. For subsequent interactions, such as l-changing collisions between Rydberg states induced by low-energy electrons, the cross sections are substantially larger and scale proportionally to n^4 owing to the extended size of the Rydberg orbital, which presents a geometrically large target for the incident electron; values reach approximately $10^6 Ų at n = 20. Experiments employ collimated electron beams with energies of 10–100 directed into a low-pressure gas or crossed with an atomic beam to achieve controlled collisions, minimizing background interactions. The resulting Rydberg atoms are detected via field ionization, applying a pulsed or swept to strip the loosely bound Rydberg and collect the ions with a detector, enabling measurement of populated n and l distributions. Seminal work in the late 1960s and 1970s, including studies on by Schiavone et al., utilized such setups to characterize Rydberg state populations and advance early spectroscopic investigations of these highly excited species. This technique suffers from limitations inherent to collisional processes, including spectral broadening due to multiple -atom interactions in the target gas, which populates a of s rather than isolating specific levels. Compared to optical methods, offers lower selectivity, often yielding broad distributions in both n and orbital l, though its nature facilitates studies of Rydberg manifolds.

Charge Exchange Excitation

Charge exchange excitation involves the resonant transfer of an from a atom to an incident , resulting in the formation of a Rydberg atom from the . For example, a proton (H⁺) colliding with a atom such as cesium () or () captures an , yielding a in a high state, H(n), and leaving the atom ionized as Cs⁺ or K⁺. This process is resonant when the of the compensates for the difference between the potential of the atom and the of the target Rydberg state, enabling efficient single- capture without significant excitation of . The rate coefficients for these resonant charge exchange reactions depend on the relative collision velocity and the principal n, with cross sections increasing for higher n due to the larger orbital of the Rydberg , which enhances the range. For ions at velocities around 10³–10⁴ m/s, typical rate coefficients exceed 10^{-7} cm³/s, reflecting the near-Langevin capture rates modified by the resonant condition. The extended wavefunction in Rydberg states further facilitates capture by overlapping effectively with the incoming ion's trajectory during close approaches. Quantitative measurements, such as cross sections for n=20–60 states in 20 keV H⁺ + collisions, demonstrate efficiencies as approximately 1/n³ for the state distribution, peaked around n ≈ 30–40 for optimal . In laboratory applications, this method is utilized in fast beams passed through vapor cells to generate controlled beams of Rydberg atoms for and collision studies, as well as in ion trap experiments where tuned energies enable selective . It also plays a key role in astrophysical modeling, where charge exchange contributes to Rydberg atom populations in ionized plasmas, influencing recombination rates and emissions in nebulae and stellar atmospheres. A primary is the efficient production of high-n states (n > 50) without requiring complex systems, allowing access to weakly bound electrons in regimes where optical is inefficient or impractical.

Optical Excitation

Optical excitation of Rydberg atoms typically involves laser-based schemes to promote atoms from the to high-lying Rydberg levels through stepwise or direct processes, enabling precise control over the excitation dynamics. In alkali atoms like , a common two-photon excitation pathway proceeds via an , such as 5S_{1/2} \to 5P_{3/2} \to nS_{1/2} or nD_{3/2/5/2}, using tunable lasers operating near 780 nm for the first step and around 480 nm for the second step to access principal quantum numbers n up to several hundred. This method leverages the large transition dipole moments in the upper steps, allowing efficient with pulse durations on the order of microseconds. Multi-step schemes with more than two photons are employed for species like or , where direct single-photon access is challenging due to wavelength constraints, but two-photon processes remain prevalent for their coherence and selectivity. The application of electric dipole selection rules, which dictate \Delta l = \pm 1 per transition (where l is the orbital angular momentum quantum number), ensures state-specific excitation and high purity of the Rydberg population, often exceeding 99% in controlled environments like magneto-optical traps. These rules, rooted in the quantum-mechanical description of atomic transitions, restrict accessible states and suppress unwanted excitations to nearby levels, facilitating the isolation of pure n, l manifolds essential for coherent manipulation. For instance, in rubidium experiments, careful polarization control of the lasers aligns with \Delta m_l = 0, \pm 1 rules to further enhance fidelity. A key challenge in optical excitation arises from AC Stark shifts induced by the intense fields, which can detune the and broaden the linewidth, alongside off-resonant excitations that populate unwanted states and reduce . These effects are particularly pronounced for high n states due to the enhanced , potentially shifting by several MHz. Mitigation strategies include the use of chirped pulses, where the frequency is swept adiabatically to follow the time-dependent Stark shifts, maintaining and achieving transfer efficiencies above 90% while minimizing decoherence from in the . Modern techniques extend optical to Rydberg dressing, where weak, off-resonant laser fields couple the ground or low-lying states to Rydberg levels with only a small (typically 1-10%) of Rydberg character, avoiding full while imparting tunable interactions. This partial dressing, often via two-photon schemes detuned by tens of MHz, creates effective potentials for quantum simulation and sensing, with the dressed state's scaling as n^7 for interaction strengths on the order of kHz at micrometer separations. Such approaches have been demonstrated in ultracold gases, enabling coherent control without the losses associated with pure Rydberg states.

Response to External Fields

Stark and Zeeman Effects

In Rydberg atoms, the describes the perturbation of energy levels due to an external , with the response depending on the degree of degeneracy in the principal n manifold. For hydrogenic systems, where states within a given n are highly degenerate, the linear Stark effect dominates, arising from the first-order perturbation that lifts the degeneracy. The energy shift for these states is given by \Delta E = -\frac{3}{2} n (n_1 - n_2) e E a_0, where n_1 and n_2 are the parabolic quantum numbers, e is the elementary charge, E is the electric field strength, and a_0 is the Bohr radius. This linear shift was first experimentally observed in high-n Rydberg states of barium through high-resolution absorption spectroscopy, revealing a regular manifold structure attributed to parabolic quantization. In contrast, for non-hydrogenic Rydberg atoms like alkali metals, the presence of quantum defects breaks the degeneracy, resulting in a quadratic Stark effect for low angular momentum states (\ell \lesssim 4), where the energy shift scales as \Delta E \propto -\alpha E^2 / 2 and the polarizability \alpha is enhanced, reaching values up to \sim n^7 a_0^3. The in Rydberg atoms involves the splitting of energy levels in a due to the interaction of the atomic with the field. The linear Zeeman shift is \Delta E = \mu_B B m_j g_L, where \mu_B is the , B is the strength, m_j is the projection of the total , and g_L \approx 1 is the orbital for states dominated by orbital motion. In high-\ell Rydberg states, where \ell \approx n-1, this splitting is significantly enhanced because the maximum |m_j| \approx \ell \sim n, leading to total manifold splittings on the order of n \mu_B B, far larger than in ground-state atoms. External fields induce significant state mixing within the Rydberg n manifold through the , as the couples states with \Delta \ell = \pm 1 and \Delta m = 0, producing hybrid states with permanent electric dipole moments up to \sim n^2 e a_0. These dipoles, oriented along or against the field, result in avoided level crossings at specific field strengths, where the energy gap is determined by the coupling matrix element, altering transition probabilities and pathways. In the presence of both electric and magnetic fields, or combined with , this mixing promotes classical , manifesting as irregular spectral features and enhanced dynamical processes. experiments on Rydberg atoms in such fields have demonstrated the onset of through broadened resonances, recurrence spectra, and increased rates, providing a direct probe of the classical-to-quantum transition.

Diamagnetic Effects

In Rydberg atoms, diamagnetic effects stem from the quadratic interaction with magnetic fields, introducing a perturbative term to the Hamiltonian that confines the outer electron's orbit perpendicular to the field direction. The relevant Hamiltonian contribution is H_d = \frac{e^2 B^2}{8 m_e} r_\perp^2, where e is the electron charge, B is the magnetic field strength, m_e is the electron mass, and r_\perp denotes the cylindrical radius perpendicular to the field axis (assumed along z). This term arises from the \mathbf{p} \cdot \mathbf{A} coupling in the minimal substitution for the vector potential \mathbf{A} = \frac{1}{2} \mathbf{B} \times \mathbf{r}, expanded to second order. For high principal quantum numbers n, the expectation value \langle r_\perp^2 \rangle \propto n^4 a_0^2 (with a_0 the Bohr radius), yielding an energy shift \Delta E_d \propto B^2 n^4. This scaling exaggerates the response in Rydberg states compared to ground-state atoms, where such shifts are negligible, enabling the diamagnetic regime at laboratory fields of order 1–10 T for n \gtrsim 30. Classically, the diamagnetic term imposes a harmonic confinement on the electron motion in the plane perpendicular to \mathbf{B}, with frequency \omega_d = e B / (2 m_e). In the presence of crossed electric and magnetic fields, the electron trajectories form bounded diamagnetic orbits characterized by an \mathbf{E} \times \mathbf{B} drift superimposed on cyclotron oscillations, limiting the orbital excursion to a finite radius r \approx m_e c |\mathbf{E}| / (e B^2) (in cgs units). These orbits contrast with the unbounded hyperbolic trajectories in pure electric fields, stabilizing the Rydberg electron against ionization for certain initial conditions and field orientations. The classical dynamics reveal a transition from integrable Kepler-like motion at low fields to non-separable, potentially chaotic paths as the diamagnetic confinement strengthens. Quantum manifestations of these effects appear prominently in the high-n manifolds, where the diamagnetic perturbation lifts the degeneracy of hydrogenic levels and induces level repulsion, creating avoided crossings that reorder the spectrum within the n shell. Additionally, quantum scarring emerges, with eigenstates concentrating probability density along unstable classical periodic orbits, enhancing recurrence signals in time-dependent wave packet evolution. These features were first experimentally resolved in the 1980s through high-resolution spectroscopy of alkali Rydberg atoms, such as lithium and sodium, in fields up to 0.6 T, revealing irregular spacings and scar-like modulations in photoabsorption spectra. The onset of in the diamagnetic regime occurs when the characteristic energy shift from H_d surpasses the unperturbed level spacing \Delta E \approx 1/(2 n^3) (in Rydberg units), roughly satisfying B^2 n^4 / n^3 \sim B n \gtrsim 1 (in scaled where of proportionality depends on the specific ). This marks the of perturbative treatments, leading to full n-mixing, exponential proliferation of periodic orbits, and a crossover in level statistics from Poissonian (integrable) to Wigner-Dyson () distributions. Early 1980s experiments on and Rydberg series confirmed this threshold through deviations from regular quasi-Landau resonances into absorption.

Applications

Precision Measurements

Rydberg atoms are particularly valuable for precision metrology due to their long radiative lifetimes and large elements, which enable extended coherent interrogation times and high-fidelity state manipulation. Trapping these atoms in optical s minimizes and environmental perturbations, facilitating measurements with exceptional stability. Theoretical proposals suggest that Rydberg excitation can mediate strong interactions to generate spin-squeezed states in optical clocks, potentially surpassing the standard and enhancing performance for tests of fundamental symmetries and timekeeping. The pronounced sensitivity of Rydberg states to (BBR) shifts arises from their enhanced polarizabilities, which scale as n^7 for scalar components and introduce dynamic Stark shifts proportional to the thermal photon field. Rydberg atoms are thus sensitive probes for ambient radiation effects in . High-n Rydberg provides a pathway to refine the by accessing nearly -like states where quantum defects—deviations from the potential due to core penetration—are minimized. In alkali atoms like cesium, precision measurements of transitions to circular Rydberg states (maximal , l = n-1) suppress these defects, yielding effective principal quantum numbers with relative uncertainties around 10^{-10}. Such experiments, often performed in cold atomic ensembles, have directly supported updates to the value, reducing its overall CODATA uncertainty from prior limitations.

Role in Plasmas

Rydberg atoms form in plasmas through processes such as impact and radiative or recombination, particularly in low-density regimes where high (high-n) states are populated without rapid . In partially ionized gases, free electrons collide with ground-state atoms, exciting them to Rydberg levels, while recombination of ions and electrons can directly cascade into these extended orbital states. These mechanisms are prominent in ultracold plasmas, where initial Rydberg evolves into a phase through -Rydberg , leading to further and heating. In thermal equilibrium, the population fraction of Rydberg states follows the Saha-Boltzmann distribution, given by f_n \propto n^2 \exp(-E_n / kT), where n is the principal quantum number, E_n is the binding energy, k is Boltzmann's constant, and T is the electron temperature. This distribution indicates that Rydberg states become significantly populated at temperatures around 1 eV, as the exponential term balances the n^2 degeneracy factor for high-n levels near the ionization threshold. In low-density plasmas, where collisional de-excitation is minimal, this equilibrium allows substantial fractions of atoms to occupy Rydberg configurations, influencing overall plasma dynamics. Rydberg atoms serve as effective diagnostics in plasma environments, particularly through the analysis of broadening, which reveals local and . High-n transitions exhibit pronounced Stark broadening due to their large orbital radii and to electric microfields from charged particles, enabling measurements in devices like tokamaks. For instance, broadening of Balmer lines from Rydberg states provides non-intrusive probes of edge conditions, with linewidths scaling with and correlating to temperatures via established models. Interactions among Rydberg atoms in s involve collisions that can trigger Förster resonances, where resonant between pairs modifies state populations and enhances . These dipole-dipole exchanges, tuned by external fields or conditions, lead to efficient state mixing in dense Rydberg gases transitioning to s. Additionally, screening via effects reduces the effective range of Rydberg-Rydberg interactions, altering collision cross-sections and stabilizing high-n states against premature in screened environments.

Astrophysical Relevance

Rydberg atoms play a crucial role in astrophysical environments, particularly within H II regions where hot stars ionize , leading to the formation of partially ionized plasmas. In these nebulae, electrons recombine with protons to form highly excited atoms in Rydberg states with principal quantum numbers n > 100, which are dominant at electron temperatures around 10^4 due to the thermal population distribution favoring large n. These atoms emit radio recombination lines (RRLs) during cascades from high-n levels, serving as key diagnostics for the physical conditions of the . RRLs primarily arise from Δn = 1 transitions, known as α lines, in atoms within H II regions. The frequency of these lines follows the approximate relation \nu \propto \frac{1}{n^2 (n+1)}, placing higher-n transitions in the radio domain for n ≳ 30. Linewidths are broadened by thermal Doppler effects from velocities and in the , typically yielding widths of several km/s that reflect the kinematic structure of the emitting gas. Observations of RRLs act as analogs to the 21 cm hyperfine line of neutral but probe ionized gas at higher n, enabling maps of proton density (H⁺) and temperature. For instance, multiline RRL studies toward the have derived densities ranging from 10^3 to 10^5 cm⁻³ and temperatures of 7000–9000 across its H II regions, revealing spatial variations in ionization structure. A major challenge in RRL observations is free-free absorption by the thermal bremsstrahlung continuum in the H II , which can optically thicken at lower frequencies and mask line emission. This effect is mitigated through modeling with escape probability approximations, accounting for non-local conditions and photon trapping in dense nebulae.

Strongly Interacting Systems

In dense ensembles of Rydberg atoms, long-range interactions become dominant, leading to quantum many-body effects that enable the realization of strongly correlated . The primary interaction is the van der Waals potential, given by V(r) = \frac{C_6}{r^6}, where the coefficient C_6 scales as n^{11} with the principal n, resulting in interaction strengths that grow rapidly with excitation level. This scaling arises from the large dipole moments of Rydberg states and the near-resonant coupling between pair states. These van der Waals interactions can be modified near Förster resonances, where the energy defect (Förster defect) between pair states is small, allowing resonant dipole exchange and altering the from repulsive to attractive or oscillatory forms. Such defects lead to enhanced blockade radii scaling approximately as n^4 [a_0](/page/Bohr_radius), where [a_0](/page/Bohr_radius) is the Bohr radius, due to the transition to a dipole-dipole regime with stronger near-field coupling. A key consequence is the Rydberg blockade, where the strong interaction shifts the double-excitation energy far from , preventing simultaneous excitation of multiple atoms within a blockade volume of approximately \mu \sim 10 \, \mu \mathrm{m}^3. This effect, first proposed for implementing fast quantum gates via controlled phase shifts between atoms, forms the basis for entangling operations in neutral-atom platforms. In larger ensembles, these interactions give rise to complex many-body states, such as effective models mediated by processes that map Rydberg excitations to flips. For instance, in optical lattices, the and terms simulate Ising chains, enabling the study of transitions and correlated dynamics in one- and two-dimensional arrays. Experimental realizations rely on ultracold atomic gases, typically or cesium, excited to Rydberg states using lasers within or lattices to control positions and densities. Breakthroughs in the demonstrated collective effects like , where synchronized emission from Rydberg ensembles enhances decay rates and reveals infinite-range interactions, as observed in elongated clouds with enhanced output directional along the excitation axis. As of 2025, Rydberg atoms continue to advance quantum technologies, with scalable neutral-atom arrays demonstrating error-corrected logical qubits and improved fidelities in multi-qubit gates for quantum and .

Advanced Modeling and Research

Classical Simulations

Classical simulations of Rydberg atoms model the 's motion as Newtonian within the potential of the augmented by external fields, such as electric or magnetic perturbations, leading to dynamics when these non-integrable interactions disrupt regular motion. In this framework, the follows classical trajectories governed by Hamilton's equations, where emerges from the to initial conditions in systems like the exposed to fields, enabling the study of thresholds without quantum wavefunctions. This approach contrasts with full quantum treatments by focusing on ensemble averages of trajectories to capture diffusive energy spread and . A key technique in these simulations is the use of Poincaré surfaces of section, which intersect the to reveal structures like invariant KAM tori in integrable regimes; their breakdown under perturbations visualizes the onset of , as seen in Rydberg electrons in combined electric and magnetic fields. By plotting and at fixed energy intervals, these sections highlight the transition from quasi-periodic motion on tori to ergodic filling of , providing insights into the classical analog of quantum scarring or level statistics in Rydberg systems. Such simulations find applications in predicting photoionization rates under intense laser fields and field ionization in static electric fields, where classical trajectories accurately reproduce quantum results for principal quantum numbers n > 30 due to the correspondence principle, as the de Broglie wavelength becomes negligible compared to orbital scales. For instance, in microwave-driven ionization, classical models capture the scaling of rates with field strength and frequency, matching experimental thresholds for high-n states. However, these methods fail to account for quantum tunneling in barrier penetration or coherent effects at low n, where wave-like interference dominates; nonetheless, they remain valuable for exploring high-field diamagnetism, simulating electron orbits in strong magnetic fields that align with quantum predictions of permanent dipole moments.

Emerging Research Directions

Recent advancements in Rydberg atom research have focused on leveraging their strong interactions for scalable quantum technologies, particularly in . Neutral atom arrays using Rydberg-mediated have demonstrated high-fidelity operations, with two-qubit controlled-phase achieving approximately 99% . As of 2025, these developments enable programmable quantum processors supporting thousands of physical qubits encoded in logical states, including arrays of up to 6,100 qubits and continuous operation of 3,000-qubit systems, paving the way for fault-tolerant computation. Rydberg molecules continue to reveal novel binding mechanisms, with ultralong-range structures formed by embedding ground-state atoms within the Rydberg electron orbital, extending bond lengths to micrometers. Post-2020 studies have explored vibronic states in heteronuclear systems like Cs-RbCs, where nonadiabatic couplings influence molecular stability and partial wave amplitudes. Such molecules hold potential for realizing Rydberg polarons, where coherent facilitates excitation transport in ultracold environments. Hybrid integrations of Rydberg atoms with other quantum platforms are emerging to enable quantum networks. Coupling Rydberg ensembles to optical cavities supports photon-mediated entanglement for , integrating atomic processors with fiber-optic links. Interfaces with superconducting qubits via resonators allow transduction between microwave and optical domains, achieving high-fidelity state transfer. Despite these progresses, key challenges persist in Rydberg systems. Mitigating decoherence from motional effects, such as van der Waals-induced in arrays, remains critical for maintaining in large-scale simulators. Precise control at high principal quantum numbers (n > 200) demands advanced trapping and state preparation to counter increased sensitivity to fields and . Additionally, simulating condensed matter phenomena like Hubbard models requires overcoming interaction range limitations to accurately capture strong correlations in fermionic systems. Ongoing 2025 research includes probing quantum floating phases in arrays of up to 92 qubits and entangling Rydberg superatoms via single-photon interference.

References

  1. [1]
    [PDF] Quantum information with Rydberg atoms - UW–Madison Physics
    Aug 18, 2010 · The advances of the last decade are reviewed, covering both theoretical and experimental aspects of Rydberg-mediated quantum information ...
  2. [2]
    Review of Cold Rydberg Atoms and Their Applications
    In this review article, we provide an introductory overview of the most important topics in this fast evolv- ing young research field of cold Rydberg atoms. We ...Missing: paper | Show results with:paper
  3. [3]
    [PDF] Cold and ultracold Rydberg atoms in strong magnetic fields
    Oct 14, 2009 · Rydberg atoms, by virtue of their special properties, are of fascination in atomic and optical physics [1]. These atoms whose size scales as the ...
  4. [4]
    Atomic Spectroscopy - Term Series, Quantum Defects | NIST
    Oct 3, 2016 · A spectral-line series results from either emission or absorption transitions involving a common lower level and a series of successive (core)nl upper levels.
  5. [5]
    Transitions - Hydrogen Energy Levels - NAAP - UNL Astronomy
    Absorbing Photons​​ The formula defining the energy levels of a Hydrogen atom are given by the equation: E = -E0/n2, where E0 = 13.6 eV (1 eV = 1.602×10-19 ...
  6. [6]
    Hydrogen energies and spectrum - HyperPhysics Concepts
    Hydrogen Energy Level Plot. The basic structure of the hydrogen energy levels can be calculated from the Schrodinger equation. The energy levels agree with ...
  7. [7]
    Rydberg Physics
    ... quantum defect δ For alkali atoms in a Rydberg state, the energy of the Rydberg level described by the principal quantum number n is given by. En = I - Ry ...
  8. [8]
    Rydberg Macrodimers: Diatomic Molecules on the Micrometer Scale
    Scaling Laws. Many properties of Rydberg atoms can be calculated using their characteristic dependence on the effective principal quantum number n* = n – δ(n, ...
  9. [9]
  10. [10]
    Rydberg constant | Definition, Formula, Value, & Facts - Britannica
    Sep 12, 2025 · Fundamental constant of atomic physics that appears in the formulas developed (1890) by the Swedish physicist Johannes Rydberg.
  11. [11]
    Michael John Seaton - 16 January 1923 — 29 May 2007 - Journals
    Pengelly and Seaton (13) developed Mike's semi-classical impact parameter method to treat the collisions with ions, h+, he+ and he2+, which cause l-changing ...
  12. [12]
    [PDF] Rydberg Atoms in Strong Fields: A Testing Ground for Quantum Chaos
    The unperturbed energy levels E,,l of alkali Rydberg levels can be computed from the known quantum defects. Using these energies the radial equation is ...
  13. [13]
    Quantum technologies with Rydberg atoms - Frontiers
    Jul 11, 2024 · We present a brief review of recent progress in the development of quantum technologies using Rydberg atoms.Quantum information... · Quantum simulation of many... · Single-photon source for...Missing: 2020s | Show results with:2020s
  14. [14]
    An Introduction to Rydberg atoms with ARC
    Rydberg states are highly excited states of the outer valence electron where the properties can be scaled in terms of the principal quantum number, n.
  15. [15]
    An open-source library for calculating properties of alkali Rydberg ...
    We present an object-oriented Python library for the computation of properties of highly-excited Rydberg states of alkali atoms.
  16. [16]
    [PDF] Chapter 1: Quantum Defect Theory
    Rydberg atoms are excited states of atoms with a large principle quantum number, where the Rydberg electron is only weakly bound to the ionic core.
  17. [17]
    Rydberg Atoms - Cambridge University Press & Assessment
    This book provides a comprehensive description of the physics of Rydberg atoms ... quantum defect theory in the study of Rydberg atomic systems is discussed.
  18. [18]
    [PDF] Hydrogen Radial Wavefunctions - MIT OpenCourseWare
    Since the differential equation depends on и, R(r) must also depend on и, thus. Rnи(r) is the radial part of ψ, and it will generally be an explicit function of.
  19. [19]
    Electron-impact excitation of helium: Cross sections, n, and l ... - OSTI
    ... n = 4 that agrees with the optically measured n = 4 cross section. The 108-..mu..sec time of flight requires that the high Rydberg He atoms have lifetimes ...
  20. [20]
    Electron Impact Excitation - an overview | ScienceDirect Topics
    Kocher and Smith used electron impact excitation to excite Li atoms in a beam to Rydberg states [10]. As shown by Figure 2, they crossed the thermal Li beam ...
  21. [21]
    Cross sections, $n$, and $l$ distributions of high Rydberg states
    Jul 1, 1977 · Helium atoms in highly excited Rydberg states are produced by electron impact and are detected by electric field ionization.
  22. [22]
    Ultra-cold ion–atom collisions: near resonant charge exchange
    Mar 17, 2008 · The huge value of the rate constant that we obtain for the charge exchange in DT+ of. 1.6×10−8 cm3 s−1 is a value more typical of ionic ...
  23. [23]
    Production of Rydberg Atoms - ScienceDirect
    Rydberg atoms, those in states of high principal quantum number n, have an excited valence electron which is weakly bound in a large orbit.
  24. [24]
    Uncertainties in Atomic Data for Modeling Astrophysical Charge ...
    Besides, the charge exchange process not only produces X-ray lines, but also generates lines in the ultraviolet and optical band as the Rydberg levels populated ...
  25. [25]
    Two-Photon Laser Excitation of Rb Rydberg Atoms in the Magneto ...
    Oct 27, 2023 · In this paper, we report our first experimental results on the two-photon laser excitation 5S1/2→5P3/2→nS with a homemade 480 nm laser in the ...
  26. [26]
    Quantum information with Rydberg atoms | Rev. Mod. Phys.
    Aug 18, 2010 · Rydberg atoms with principal quantum number n ⪢ 1 have exaggerated atomic properties including dipole-dipole interactions that scale as n 4 ...Rydberg Atoms and Their... · Rydberg Gates · Collective Effects in Rydberg...
  27. [27]
    Excitation of Rydberg states in rubidium with near infrared diode lasers
    In this work, we demonstrate a diode laser system for use in populating Rydberg states of rubidium atoms in a magneto-optical trap (MOT). As a test of our laser ...
  28. [28]
    [PDF] Two-Photon Rydberg Excitation of Neutral Atoms in Optical Dipole ...
    May 29, 2017 · The signal used to detect Rydberg excitations is the recapture rate of the atoms in a dipole trap. This is suitable as atoms in the Rydberg ...
  29. [29]
    Enhancement of Rydberg Atom Interactions Using ac Stark Shifts
    The ac Stark effect was used to induce resonant energy transfer between translationally cold 8 5 R b Rydberg atoms. When a 28.5 GHz dressing field was set ...Missing: chirped | Show results with:chirped
  30. [30]
    Creation of quantum entangled states of Rydberg atoms via chirped ...
    At the resonance, the chirp of both pulses gets turned off to prevent further ... a.c. Stark shifts and linearly chirped pulses. It is assumed for a ...
  31. [31]
    Creation of quantum entangled states of Rydberg atoms via chirped ...
    Jun 21, 2021 · At the resonance, the chirp of both pulses gets turned off to ... a.c. Stark shifts and linearly chirped pulses. It is assumed for a ...
  32. [32]
    Tuning Ultracold Chemical Reactions via Rydberg-Dressed ...
    Jul 8, 2014 · The v = 1 level is particularly sensitive to a weak amount of Rydberg dressing, with the resonance having a larger magnitude and moving to much ...
  33. [33]
    Rydberg-Stark deceleration of atoms and molecules
    Mar 1, 2016 · This is achieved by exploiting the selection rules for electric dipole transitions in the photoexcitation process to control the absolute value ...Introduction · Rydberg States In Electric... · Electrostatic Trapping...<|separator|>
  34. [34]
    Diamagnetic Zeeman Effect and Magnetic Configuration Mixing in Long Spectral Series of BA I
    - **Title:** Diamagnetic Zeeman Effect and Magnetic Configuration Mixing in Long Spectral Series of Ba I
  35. [35]
  36. [36]
    Page not found | Nature
    Insufficient relevant content.
  37. [37]
    [PDF] CODATA recommended values of the fundamental physical constants
    Sep 26, 2016 · This paper gives the 2014 self-consistent set of values of the constants and conversion factors of physics and chemistry recommended by the ...Missing: blackbody | Show results with:blackbody
  38. [38]
    Measuring the Rydberg constant using circular Rydberg atoms in an ...
    A method for performing a precision measurement of the Rydberg constant R ∞ using cold circular Rydberg atoms is proposed. These states have long lifetimes ...
  39. [39]
    Microwave spectroscopy of the Yb , , and transitions | Phys. Rev. A
    Apr 15, 2019 · The Yb + ion is of great interest for use in an optical frequency standard, the study of parity nonconservation (PNC), and the study of the ...
  40. [40]
    Rydberg Atom Formation in Ultracold Plasmas: Small Energy ...
    Jun 5, 2008 · We present extensive Monte Carlo calculations of electron-impact-induced transitions between highly excited Rydberg states and provide accurateMissing: via | Show results with:via
  41. [41]
    Evolution dynamics of a dense frozen Rydberg gas to plasma
    Oct 21, 2004 · The predominant scattering process in the formation of the plasma is electron scattering from Rydberg atoms. Initially one would expect ℓ - ...
  42. [42]
    Transition rates for a Rydberg atom surrounded by a plasma
    Apr 27, 2016 · Rydberg atoms, as ubiquitous states formed due to the recombination of electrons and positive ions in laser-produced plasmas [14] , play a ...
  43. [43]
    Populations of Rydberg states of atoms in nebulae
    Then the equilibrium population of the level n will be expressed by the Saha-Boltzmann ... S. A. GULYAEV and S. A. NEFEDOV: Populations of Rydberg states ...
  44. [44]
    Dielectronic recombination data for dynamic finite-density plasmas
    the Saha–Boltzman populations and ionization fractions in the high-density ... We do note again that high Rydberg states in a finite density plasma are.
  45. [45]
    Review of Rydberg Spectral Line Formation in Plasmas - MDPI
    Nov 10, 2004 · In the present review, we have considered Rydberg hydrogen atoms (hydrogen-like ions) in a low-density plasma, demonstrating several methods for ...
  46. [46]
    Using Laser-Induced Rydberg Spectroscopy diagnostic for direct ...
    Aug 6, 2018 · We discuss a novel diagnostic allowing direct measurements of the local electric field in the edge region of NSTX/NSTX-U. This laser based ...
  47. [47]
    Interplay between thermal Rydberg gases and plasmas | Phys. Rev. A
    Apr 19, 2019 · The electrons produced by the ionization of the Rydberg atoms become an additional source for collisions with Rydberg atoms leading to even more ...
  48. [48]
    The role of the Debye screening of circular Rydberg states of ...
    Jun 18, 2020 · We show in detail how the screening decreases the value of the critical electric field required for the ionization at different values of the ...
  49. [49]
    Dense plasmas, screened interactions, and atomic ionization
    The Saha–Boltzmann equation can be used to show that, at fixed temperature, the degree of ionization z̄ increases as n=∑znz decreases. And, if one accounts for ...
  50. [50]
    Radio recombination lines from the largest bound atoms in space
    In this paper, we report the detection of a series of radio recombination lines (RRLs) in absorption near 26 MHz arising from the largest bound carbon atoms ...
  51. [51]
    Multiline observations of hydrogen, helium, and carbon radio ...
    We present a study of hydrogen, helium, and carbon millimeter-wave radio-recombination lines (RRLs) toward 10 representative positions throughout the Orion ...<|control11|><|separator|>
  52. [52]
    Radio Recombination Lines in Galactic H II Regions - NASA ADS
    We report radio recombination line (RRL) and continuum observations of a sample of 106 Galactic H II regions made with the NRAO 140 Foot (43 m) radio telescope.Missing: Rydberg | Show results with:Rydberg
  53. [53]
    A pilot study of Galactic radio recombination lines using FAST
    Transitions with Δn = 1 are called α lines, transitions with Δn = 2 β lines, and so on. For hydrogen, the frequencies of α lines with n ≳ 30 are in the ...Missing: formula | Show results with:formula
  54. [54]
    Discovery of Radio Recombination Lines from Proplyds in the Orion ...
    Apr 9, 2025 · Our study demonstrates that radio recombination lines are readily detectable in ionized photoevaporating disks, providing a new way to measure ...
  55. [55]
    An nl-model with a full radiative transfer treatment for level ...
    The escape probability approximation was used to estimate the effect of the radiation field in previous models for calculating hydrogen-level populations. The ...Missing: challenges | Show results with:challenges
  56. [56]
    Radio and infrared recombination studies of the southern massive ...
    Free–free emission is dominated by interactions of free electrons with the most abundant element, singly ionized hydrogen. Observations of the radio free–free ...
  57. [57]
    van der Waals interaction potential between Rydberg atoms near ...
    Dec 29, 2017 · An exception is the van der Waals interaction between highly excited (Rydberg) atoms that, due to the scaling of the C 6 coefficient as n 11 ...Missing: C6 | Show results with:C6
  58. [58]
    [PDF] Chapter 2: Interacting Rydberg atoms
    The Hamiltonian for this quantum many-body system can be written ... [2] H. A. Bethe and E. E. Salpeter, Quantum Mechanics of One- And Two-Electron Atoms.
  59. [59]
    Highly Resolved Measurements of Stark-Tuned Förster Resonances ...
    Mar 14, 2012 · We report on experiments exploring Stark-tuned Förster resonances between Rydberg atoms with high resolution in the Förster defect.
  60. [60]
    Rydberg-Rydberg interaction strengths and dipole blockade radii in ...
    Oct 19, 2023 · The strength of the long-range interaction between two atoms is quantified by the van der Waals coefficient C 6 . A larger C 6 value ...
  61. [61]
    Fast Quantum Gates for Neutral Atoms | Phys. Rev. Lett.
    Sep 4, 2000 · Fast Quantum Gates for Neutral Atoms. D. Jaksch, J. I. Cirac, and P. Zoller · S. L. Rolston · R. Côté1 and M. D. Lukin2. Institut für ...
  62. [62]
    Simulating quantum spin models using Rydberg-excited atomic ...
    At large separations and away from Förster resonances, Rydberg–Rydberg interactions can be treated perturbatively leading to van der Waals (vdW) ...<|control11|><|separator|>
  63. [63]
    Coherent Many-Body Spin Dynamics in a Long-Range Interacting ...
    Dec 14, 2017 · ... Rydberg-dressed Ising spin chains of atoms trapped in an optical lattice. We identify partial many-body revivals at up to about ten times ...<|separator|>
  64. [64]
    Many-body interferometry of a Rydberg-dressed spin lattice - Nature
    Aug 1, 2016 · Here we report the realization of coherent Rydberg dressing to implement a two-dimensional synthetic spin lattice. Our single-atom-resolved ...Missing: chains | Show results with:chains
  65. [65]
    Super-radiance reveals infinite-range dipole interactions through a ...
    Nov 30, 2017 · Super-radiance can be measured as an enhanced decay rate at short times. Both effects can provide quantitative experimental evidence of ...
  66. [66]
    Chaotic ionization of highly excited hydrogen atoms
    The experimental and theoretical studies of the microwave ionization of highly excited hydrogen atoms provide a unique opportunity to explore two new ...
  67. [67]
    Effects of Classical Resonances on the Chaotic Microwave ...
    Effects of Classical Resonances on the Chaotic Microwave Ionization of Highly Excited Hydrogen Atoms, R V Jensen. ... Ionization of Rydberg hydrogen atoms ...
  68. [68]
    [PDF] The Kicked Rvdberg Atom: Regular and Stochastic Motion ... - OSTI
    We have investigated the dynamics of a three-dimensional classical Rydberg atom ... The KAM tori do not topologically separate different regions of chaotic. Page ...
  69. [69]
    Classical view of the properties of Rydberg atoms - AIP Publishing
    Apr 1, 1992 · Classical view of the properties of Rydberg atoms: Application of the correspondence principle Available. T. P. Hezel;. T. P. Hezel.
  70. [70]
    High-fidelity parallel entangling gates on a neutral-atom ... - Nature
    Oct 11, 2023 · Generically, with these global pulses and Rydberg blockade, one can natively realize symmetric, diagonal gates (for example, CPHASE gates as ...
  71. [71]
    Logical quantum processor based on reconfigurable atom arrays
    Dec 6, 2023 · Here we report the realization of a programmable quantum processor based on encoded logical qubits operating with up to 280 physical qubits.
  72. [72]
    Ultralong-range Rydberg molecules - IOPscience
    Oct 9, 2024 · Ultralong-range Rydberg molecules (ULRMs) comprise a Rydberg atom in whose electron cloud are embedded one (or more) ground-state atoms that are weakly-bound.
  73. [73]
    Ultralong-range Cs-RbCs Rydberg molecules: Nonadiabaticity of ...
    Jul 15, 2024 · ... ultralong-range Rydberg molecules. ... Additionally, the partial wave amplitudes of the vibronic states are also affected by the nonadiabatic ...
  74. [74]
    Quantum Computing with Circular Rydberg Atoms
    Aug 6, 2021 · In this work, we propose a novel approach to Rydberg-atom arrays using long-lived circular Rydberg states in optical traps.
  75. [75]
    Hubbard physics with Rydberg atoms: Using a quantum spin ...
    Taking as an example a Rydberg-based analog quantum processor to solve the interacting spin model, we avoid the challenges of variational algorithms or ...Article Text · INTRODUCTION · THEORY · RESULTS