Fact-checked by Grok 2 weeks ago

Electromagnetically induced transparency

Electromagnetically induced transparency (EIT) is a quantum effect in which a strong renders an otherwise opaque medium transparent to a weak at resonant frequencies by creating a coherent that suppresses . This phenomenon occurs in a three-level system, typically in a (Λ) configuration, where the couples one to an , and the links the other to the same , resulting in destructive that decouples the atoms from the . The transparency window is narrow, with a linewidth determined by the intensity and decoherence rates, often expressed as \gamma_{EIT} = \gamma_{12} + |\Omega_c|^2 / \gamma_{13}, where \Omega_c is the and \gamma terms represent rates. The effect was first theoretically predicted by Kocharovskaya and Khanin in 1988, and proposed by S. E. Harris and colleagues in 1990 as a method to enable nonlinear optical processes without . EIT builds on earlier concepts like coherent population trapping from the . It was first experimentally observed in 1991 by Boller, Imamoglu, and Harris using hot vapor with pulsed lasers, demonstrating near-perfect transparency in an optically thick medium. EIT has profound implications for quantum optics and photonics, enabling phenomena such as ultraslow light propagation (group velocities reduced to meters per second), reversible storage of light pulses as atomic coherences for quantum memories, and lasing without population inversion. Practical implementations often use thermal vapors of alkali atoms like rubidium or cesium, where Doppler effects and optical depth must be managed to achieve high-contrast EIT. Analogs of EIT extend the effect beyond systems, appearing in solid-state quantum dots, superconducting circuits, and resonators, broadening its utility in integrated quantum technologies and precision sensing applications like clocks and magnetometers.

Fundamentals

Definition and Basic Principles

Electromagnetically induced transparency (EIT) is a quantum that enables coherent of light-matter interactions, creating a narrow transparency window within an otherwise absorbing medium at the resonance frequency of a weak probe field. This phenomenon arises when a strong laser field modifies the medium's optical , suppressing and for the probe light through destructive between quantum pathways. The fundamental setup for EIT involves a three-level system in a (Λ) configuration, featuring two long-lived states denoted as |1⟩ and |2⟩, and a short-lived |3⟩. A weak probe field drives the transition from |1⟩ to |3⟩, which would normally be absorbed, while a strong control field the transition from |2⟩ to |3⟩, establishing a coherent between the states. In this system, the control field induces Autler-Townes splitting of the |3⟩ state into dressed components separated by the control field's , but the key to transparency lies in the formation of a —a stationary coherent superposition of |1⟩ and |2⟩ with no admixture of |3⟩. This dark state decouples the atoms from the probe field, as the excitation amplitude via the probe interferes destructively with that induced by the control, preventing population transfer to the absorbing and resulting in transmission without loss or phase shift at line center. EIT was first experimentally observed in 1991 by Boller, Imamoğlu, and Harris using hot vapor, where the control field created a clear window for the resonant beam in an optically thick sample.

Historical Development

The foundational concepts leading to electromagnetically induced (EIT) emerged from earlier studies of coherent optical interactions in atomic s. In the , self-induced transparency was theoretically predicted and experimentally observed, describing how short, coherent light pulses could propagate through resonant media without absorption by forming solitons that match the medium's response time. Building on this, coherent population trapping (CPT) was first demonstrated in 1976, where atoms in a three-level were optically pumped into a non-absorbing superposition state, reducing under bichromatic excitation in sodium vapor. These phenomena highlighted quantum effects in multilevel atoms, laying the groundwork for manipulating light-matter interactions without dissipation. The theoretical framework for EIT was proposed in 1990 by S. E. Harris, J. E. Field, and A. Imamoğlu, who predicted that a strong coupling could induce in an otherwise absorbing probe within a three-level atomic system, accompanied by anomalous dispersion near . This work extended CPT principles to enable coherent control of optical susceptibility in dense media. Experimental confirmation followed in 1991 by K.-J. Boller, A. Imamoğlu, and S. E. Harris, who observed near-perfect and steep dispersion in an optically thick vapor using counterpropagating probe and coupling fields on an autoionizing . During the 1990s, EIT was extended to diverse systems, enhancing its versatility. In 1994, ultraslow optical dephasing was measured in europium-doped yttrium orthosilicate crystals, revealing coherence times exceeding 100 μs that would later prove essential for realizing EIT in solid-state rare-earth systems. By 1999, EIT was demonstrated in ultracold atomic ensembles, with L. V. Hau and colleagues achieving light propagation at reduced group velocities of 17 m/s in a Bose-Einstein condensate of sodium atoms, showcasing control over photonic propagation in quantum degenerate gases. These advances solidified EIT's role in .

Theoretical Framework

Atomic System and Quantum Interference

Electromagnetically induced transparency (EIT) arises fundamentally from the quantum mechanical dynamics of a three-level atomic system in a lambda (Λ) configuration, consisting of two ground states |1⟩ and |2⟩ (where |2⟩ is typically a metastable state) and a common |3⟩. The probe field couples the transition |1⟩ ↔ |3⟩ with Rabi frequency Ω_p, while the control field couples |2⟩ ↔ |3⟩ with Ω_c. Spontaneous decay occurs from |3⟩ to |1⟩ at rate γ_{31} and to |2⟩ at rate γ_{32}, with the ground-state coherence between |1⟩ and |2⟩ decaying at rate γ_{12} due to processes. This system enables coherent control of atomic excitations through quantum interference, suppressing absorption of the probe field under resonant conditions. The time evolution of the system is described by the formalism, governed by the incorporating the and dissipative terms. In the rotating frame and , the is given by H = -\hbar \left( \Omega_p |3\rangle\langle 1| + \Omega_c |3\rangle\langle 2| + \text{h.c.} \right), assuming resonant fields for simplicity (detunings can be included generally). The for the elements ρ_{ij} include coherent driving terms from the fields and relaxation: for example, the ρ_{31} evolves as \dot{\rho}_{31} = -(\gamma_{31}/2 + i\Delta_p) \rho_{31} + i \Omega_p (\rho_{11} - \rho_{33}) + i \Omega_c \rho_{21}, where Δ_p is the probe detuning, and similar expressions hold for other elements like ρ_{21} and ρ_{32}. The control field induces a Raman ρ_{12} = ρ_{21}^* between the ground states, which plays a crucial role in the process. In the , these equations reveal how the presence of Ω_c modifies the probe . A key feature of this dynamics is the formation of a coherent |D⟩, a eigenstate of the that is orthogonal to the bright state and decoupled from the optical fields: |D⟩ = (Ω_c |1⟩ - Ω_p |2⟩) / √(|Ω_p|^2 + |Ω_c|^2). This state has zero eigenvalue and contains no admixture of the decaying |3⟩, rendering it immune to and long-lived, limited only by γ_{12}. The atomic population is driven into |D⟩ by the coherent fields, preventing excitation to |3⟩ and thus eliminating . The mechanism stems from destructive quantum between excitation pathways. Without the field, the induces a ρ_{31} with imaginary part Im(ρ_{31}) proportional to . The field generates the Raman ρ_{12}, which contributes an interfering term in the equation for ρ_{31}, effectively canceling Im(ρ_{31}) at (Δ_p = 0). This redirects the interaction away from dissipative channels, resulting in a window in the . In the strong field limit (|Ω_c| ≫ γ_{31}), near-perfect is achieved on over a wide spectral window of width approximately |Ω_c|, with the residual determined by the ground-state γ_{12}, allowing for high-contrast features even in warm ensembles where γ_{12} is finite but small compared to other .

Susceptibility and Transparency Conditions

In the three-level lambda system, the linear \chi(\omega_p) for a weak interacting with the from level |1\rangle to |3\rangle is given by \chi(\omega_p) = \frac{N |\mu_{31}|^2}{\epsilon_0 \hbar} \frac{i (\Delta_p - i \gamma_{31}) }{ (\Delta_p - i \gamma_{31})(\Delta_p - \Delta_c - i \gamma_{21}) - |\Omega_c|^2 / 4 }, where N is the atomic density, \mu_{31} is the dipole matrix element, \Delta_p and \Delta_c are the and coupling detunings from resonance, \gamma_{31} and \gamma_{21} are the coherence decay rates for the respective off-diagonal elements, and \Omega_c is the of the on the |2\rangle to |3\rangle . This expression arises from the steady-state solution of the equations under the weak approximation, capturing the coherent interaction between the and coupling fields. Transparency occurs when the real part of the refractive index remains near unity while absorption is minimized, specifically when \operatorname{Re}[\chi(\omega_p)] \approx 0 and \operatorname{Im}[\chi(\omega_p)] \approx 0 at \Delta_p = \Delta_c = 0. This condition is satisfied in the limit of strong coupling, |\Omega_c| \gg \gamma_{31}, where the denominator becomes dominated by the coupling term, suppressing the probe absorption to nearly zero at line center. The resulting transparency window has a width on the order of |\Omega_c|^2 / \gamma_{31} in the weak coupling limit, which widens as the coupling strength increases. This enables high-fidelity transmission over a controllable spectral bandwidth. In the strong coupling limit, the window width approaches the Rabi splitting \sim |\Omega_c|. Within this transparency window, the dispersion exhibits an anomalous positive slope, \frac{d \operatorname{Re}[\chi]}{d \omega_p} > 0, arising from the steep variation in the real part of the susceptibility near \Delta_p = 0. This positive group index, n_g = 1 + \frac{\omega_p}{2} \frac{d \operatorname{Re}}{d \omega_p} with n \approx 1 + \operatorname{Re}[\chi]/2, contrasts with typical negative dispersion in absorbing media and underpins reduced group velocities without significant loss. Dephasing effects, particularly the decay of ground-state \gamma_{12} between levels |1\rangle and |2\rangle, modify \gamma_{21} = \gamma_{12} + \gamma_{\rm sp} (where \gamma_{\rm sp} is the small contribution), broadening the transparency window and reducing its depth. Increased \gamma_{12} due to environmental interactions diminishes the efficiency, leading to residual even at line center and a less pronounced slope. In multi-level extensions such as V-type or ladder systems, the susceptibility formula undergoes modifications to account for additional couplings or decay pathways, potentially yielding multiple transparency windows or altered profiles while preserving the core mechanism.

Experimental Aspects

Medium Requirements and Setup

To observe electromagnetically induced transparency (EIT), the medium must provide sufficient (OD), typically in the range of 1 to 100, to ensure significant interaction between the light fields and atoms while allowing measurable changes. The OD is given by OD = N σ L, where N is the atomic density, σ is the resonant absorption cross-section (on the order of λ²/2π for optical transitions, with λ the ), and L is the medium path length (often several cm in vapor cells). For effective EIT, N must exceed 10^{12} cm^{-3}, which corresponds to vapor pressures achievable in cells at moderate heating. Common media employ a lambda-type three-level , where the ground states |1\rangle and |2\rangle are long-lived (e.g., hyperfine-split levels), and |3\rangle is an coupled to both. In vapors such as (Rb) or cesium (Cs), the D2 transition is widely used, with |1\rangle as the 5S_{1/2} F=1 hyperfine level, |2\rangle as 5S_{1/2} F=2, and |3\rangle as a 5P_{3/2} level (typically F'=1 or F'=2). Solid-state alternatives, such as praseodymium-doped (Pr:YSO) crystals cooled to cryogenic temperatures (around 3-5 K), offer narrow linewidths and long coherence times due to the ionic environment, using states within the 4I_{9/2} and 4I_{11/2} manifolds for the lambda configuration. The configuration involves a weak field (power on the order of nanowatts to avoid ) resonant with the |1\rangle to |3\rangle and a strong control field (milliwatts, with Ω_c / 2π ~ 1-10 MHz) resonant with |2\rangle to |3\rangle. These fields are typically co-propagating to satisfy phase-matching conditions and minimize walk-off, with matching linear polarizations (or orthogonal circular for specific Zeeman selections) and near-zero two-photon detuning for optimal . Environmental control is essential to suppress decoherence sources. For vapor cells, temperatures below 100°C (e.g., 40-80°C) maintain the required while limiting , whose width is Δω_D = (ω / c) \sqrt{k_B T / m} (with ω the transition , m the , k_B Boltzmann's , and T ), which can otherwise smear the EIT resonance over hundreds of MHz. Buffer gases (e.g., or at 10-100 ) or trapping cold atoms (via magneto-optical traps at μK temperatures) reduce atomic collisions and transit-time effects, enhancing coherence. Key challenges include transit-time broadening, where atoms traverse the laser in time τ = w_0 / v_th (w_0 beam waist, v_th ~200 m/s), yielding linewidths Γ_tt ~ 10-100 kHz that limit EIT resolution in dilute beams. Phase matching demands precise alignment ( overlap <1 mrad) to prevent dephasing from atomic motion in the Doppler-broadened ensemble. These factors tie into the theoretical susceptibility, where medium dispersion and absorption vanish under ideal EIT conditions.

Key Demonstrations and Measurements

The first experimental demonstration of was conducted by Boller, Imamoglu, and Harris in 1991 using hot strontium vapor in a heat pipe, with a probe laser at 460.7 nm on the 5s² ¹S₀ – 5s5p ¹P₁ transition and a coupling laser at 689 nm on the 5s5p ¹P₁ – 5s5d ¹D₂ transition in a lambda-type three-level system. Transmission spectra revealed an increase in probe transmittance from approximately exp(-20) (effectively zero) without the coupling laser to exp(-1) (about 37%) with it, confirming the EIT window through destructive quantum interference that suppressed absorption. This measurement highlighted the potential for coherent control of optical properties in atomic vapors. A landmark demonstration in cold atomic ensembles came in 1999 with the work of Hau et al., who observed in a of sodium atoms cooled to 450 nK. Using a and stimulated Raman coupling, they propagated a coherent light pulse through the condensate, achieving an optical depth of around 4 but demonstrating high-fidelity transparency with pulse transmission efficiencies exceeding 50%. This experiment underscored 's role in reducing group velocity while maintaining signal integrity in ultracold media. Subsequent cold-atom realizations, such as those in rubidium ensembles, have reported optical depths up to 100 with transmission efficiencies approaching 85%, as measured by probe pulse attenuation and recovery in . Direct measurements of dispersion in EIT have quantified the dramatic slowing of light via the group velocity, given by v_g = \frac{c}{1 + \frac{\omega}{2} \frac{d \operatorname{Re}(\chi)}{d\omega}}, where c is the speed of light in vacuum, \omega is the probe frequency, and \chi is the susceptibility. In early experiments, Field, Hahn, and Harris (1995) observed pulse propagation at v_g \approx c/165 (about 1.8 \times 10^6 m/s) in , verified by time-of-flight measurements of delayed pulses with 55% transmission. Later cold-atom studies extended this to v_g \sim 17 m/s (c / 1.76 \times 10^7) in , with dispersion slopes d \operatorname{Re}(\chi)/d\omega on the order of $10^8 to $10^{10} contributing to slowdown factors of $10^3 to $10^6 times slower than c. These results were obtained through interferometric pulse delay profiling and phase-sensitive detection. Efficiency metrics in EIT experiments emphasize near-complete absorption suppression and precise phase control. Absorption has been reduced to less than 1% at the transparency window center in optimized cold rubidium ensembles, quantified via calibrated probe power transmission and spectral fitting of the imaginary susceptibility. Phase shifts, indicative of the steep dispersion, have been measured using heterodyne detection techniques, revealing nonlinear phase accumulation up to several radians without significant loss in high-optical-depth media (OD > 50). These metrics establish EIT's viability for coherent , with efficiencies scaling with and . Recent advances up to 2025 have extended EIT to room-temperature platforms, bypassing cryogenic requirements. In vapors, Lin et al. (2024) reported room-temperature Rydberg EIT in vapor cells with buffer gas, enabling applications in sensing, as quantified by high-resolution . As of 2025, further advances include EIT analogs in metamaterials achieving tunable transparency for optical applications and EIT-based storage for microwave quantum memories.

Applications and Extensions

Slow and Stopped Light

Electromagnetically induced transparency (EIT) enables the dramatic reduction of light's through the steep near the transparency window, allowing propagation speeds far below the vacuum while maintaining low absorption. In a typical Λ-type system, the group velocity v_g of the probe pulse is approximated by v_g \approx \frac{|\Omega_c|^2}{g^2 N}, where \Omega_c is the of the control field, g is the atom-light , and N is the density. By tuning \Omega_c to small values, v_g can be reduced to as low as a few meters per second, as demonstrated in ultracold gases where slowdown factors exceeding $10^7 have been achieved. This coherent slowing arises from the storage of probe field information in the atomic ground-state , preserving the pulse's quantum properties during propagation. Stopped light extends this control by effectively halting pulse propagation through dynamic EIT, where the control field is adiabatically turned off, mapping the probe pulse onto a long-lived atomic spin coherence within the medium. Upon turning the control field back on, the coherence is transferred back to the optical field, releasing the stored with high fidelity. This process relies on the dark-state formed in EIT, which transitions from a photonic to a purely as \Omega_c approaches zero. The first experimental realization of stopped light occurred in 2001 using a Bose-Einstein condensate (BEC) of sodium atoms, where a was stored for up to 1 ms before retrieval. Subsequent experiments in the extended stopped light to solid-state systems, achieving storage times over one minute in rare-earth-ion-doped crystals such as praseodymium-doped yttrium orthosilicate (Pr:YSO), including coherent storage of light pulses and images. In erbium-doped materials, coherent storage of telecom-wavelength pulses has been demonstrated, leveraging the long spin coherence times of solid hosts at cryogenic temperatures. These advancements highlight EIT's versatility beyond ultracold gases, enabling practical implementations in compact, room-temperature-compatible media. The achievable slowdown and storage durations in EIT are fundamentally limited by decoherence processes, such as atomic relaxation, and nonlinear effects like forward , which introduce noise and distortion. The maximum slowdown factor is typically on the order of the medium's (OD), as higher OD enhances but also amplifies decoherence rates, capping storage efficiency below unity for large delays. Unlike incoherent slow-light techniques, such as those based on cavity-enhanced absorption or stimulated Brillouin scattering, EIT-based slowing preserves the quantum coherence and phase information of the light pulse due to its reliance on reversible atomic coherences rather than dissipative processes. This coherence preservation distinguishes EIT for applications requiring intact photonic quantum states, such as in quantum networks.

Quantum Memory and Information Processing

Electromagnetically induced transparency (EIT) enables quantum memory through dynamic control of light propagation in atomic ensembles, where a probe photon carrying quantum information is mapped onto a collective atomic coherence for storage and later retrieval. In the standard light storage protocol, a weak probe field resonant with the |1⟩ to |3⟩ transition co-propagates with a strong control field resonant with the |2⟩ to |3⟩ transition in a Λ-type three-level system, creating a dark-state polariton that propagates slowly within the EIT window. By adiabatically reducing the control field intensity, the polariton is compressed, transferring the probe's quantum state into a long-lived spin coherence between the ground states |1⟩ and |2⟩, effectively stopping the light while preserving its quantum properties. Retrieval is achieved by reapplying the control field in the reverse direction, which reconstructs the probe pulse with high temporal and spectral fidelity. This process, first theoretically proposed and experimentally demonstrated in rubidium vapor, forms the basis for reversible quantum state transfer in EIT-based memories. Early demonstrations achieved storage efficiencies and fidelities exceeding 90% for classical coherent pulses, as reported by the Lukin group using optimized control in warm atomic vapors, enabling near-unity retrieval for pulses up to microseconds long. For quantum applications, EIT memories have stored heralded single photons generated via the Duan-Lukin-Cirac-Zoller (DLCZ) protocol, which creates correlated photon-atom excitations through spontaneous in atomic ensembles; subsequent EIT storage yields efficiencies around 70% while maintaining quantum coherence, as achieved in cold cesium atoms during the 2010s. These milestones highlight EIT's capability to handle nonclassical light states essential for tasks. In quantum for long-distance networks, EIT-based memories facilitate entanglement distribution by storing photon-matter entanglement and enabling operations, where Bell-state measurements on stored excitations from adjacent extend entanglement over lossy channels without direct transmission. The original DLCZ relies on EIT to map photonic qubits onto ensembles for purification and , achieving heralded entanglement with low error rates in proof-of-principle experiments. Integration with (QED) enhances coupling efficiency by confining the probe and fields in a high-finesse , increasing the effective and enabling deterministic single-photon from cavity-QED sources, with theoretical efficiencies approaching unity for weak excitations. Key challenges in EIT quantum memories include from , which decoheres the stored spin state, and (FWM) processes driven by the control field, generating spurious photons that degrade retrieval fidelity. limits storage times to milliseconds in warm vapors, while FWM , arising from third-order nonlinearities, can overwhelm weak quantum signals during readout. Solutions involve impedance-matched cavities, where the cavity decay rate is tuned to match the atomic absorption, minimizing losses and suppressing by balancing input-output , thereby boosting overall to over 80% even in low-optical-depth media. In the , hybrid approaches combining EIT with solid-state systems like diamond nitrogen-vacancy () centers have advanced room-temperature quantum memories, leveraging the long times of electron and spins under ambient conditions to store optical qubits with minimal cryogenic requirements.

Cooling Techniques and Other Uses

Electromagnetically induced transparency (EIT) enables sub-Doppler of neutral atoms by exploiting quantum interference to trap atoms in non-absorbing dark states, which suppresses and minimizes recoil heating from photon absorption and re-emission. In this process, atoms are optically pumped into velocity-selective dark states where the cooling mechanism relies on the detuning between the and fields, creating momentum-dependent that preferentially cools atoms near zero . Theoretical analyses predict temperatures significantly below the Doppler limit of approximately 140 μK for rubidium-87 atoms using Λ-type three-level systems. The cooling rate in EIT schemes is governed by the expression \gamma_{\text{cool}} \approx \frac{\hbar k^2}{m} \cdot \frac{\Omega_c}{\gamma}, where \hbar is the reduced Planck's constant, k is the wave number of the probe light, m is the atomic mass, \Omega_c is the Rabi frequency of the control field, and \gamma is the excited-state decay rate; this rate reflects the balance between recoil-induced diffusion and dark-state pumping efficiency. By optimizing the control field intensity and detunings, the technique can reduce temperatures to subrecoil levels while maintaining high cooling efficiency, making it suitable for preparing ultracold ensembles for precision measurements. Beyond cooling, EIT enhances Kerr nonlinearities in atomic media, yielding giant values of the nonlinear n_2 \sim 10^{-5} cm²/W, which is orders of magnitude larger than in conventional materials due to the coherent to dark states that amplifies shifts without significant . This nonlinearity facilitates all-optical switching at low light intensities, enabling compact devices for where a weak probe beam's is modulated by a beam via cross-phase modulation. In integrated , EIT has been realized in waveguides coupled to ring resonators, allowing tunable transparency windows and slow-light effects for on-chip optical buffers and modulators. EIT-based sensing leverages shifts in the transmission spectrum linewidth or resonance position due to external fields, particularly for high-sensitivity magnetometry in vapors. By monitoring Zeeman-induced splitting of EIT resonances in , high sensitivities in the fT/√Hz range have been achieved in compact vapor cells, surpassing traditional fluxgate sensors while operating without cryogenic shielding. These magnetometers exploit the narrow EIT linewidth (sub-kHz) for precise field detection over a wide , with applications in geophysical and biomedical . Recent extensions of EIT include control of pulses in coherent media, where a strong control pulse induces for an isolated extreme-ultraviolet probe, enabling manipulation of pulse absorption, emission, and delay on sub- timescales. In topological , post-2020 implementations in metamaterials, such as two-dimensional photonic crystals, demonstrate EIT-like combined with topological edge states, providing robust, backscattering-free light propagation for routing in disordered environments. As of , advances in integrated EIT devices have enabled scalable quantum memories for photonic quantum networks.

References

  1. [1]
    Observation of electromagnetically induced transparency
    We report the first demonstration of a technique by which an optically thick medium may be rendered transparent.
  2. [2]
    A practical guide to electromagnetically induced transparency in ...
    Jan 2, 2025 · This tutorial introduces the theoretical and experimental basics of electromagnetically induced transparency (EIT) in thermal alkali vapors. We ...
  3. [3]
    Nonlinear optical processes using electromagnetically induced ...
    Mar 5, 1990 · Nonlinear optical processes using electromagnetically induced transparency. S. E. Harris, J. E. Field, and A. Imamoğlu. Edward L. Ginzton ...
  4. [4]
    [PDF] Electromagnetically induced transparency - PhysLab
    When a strong coupling laser field is used to drive a resonant transition in a three-level atomic system, the absorption of a weak probe laser field.
  5. [5]
    Electromagnetically induced transparency in rubidium - AIP Publishing
    Feb 1, 2009 · Electromagnetically induced transparency (EIT) is a novel phenomenon used to modify the optical response of an atomic medium to a resonant laser field.
  6. [6]
    Electromagnetically induced transparency: Optics in coherent media
    Jul 12, 2005 · This article reviews these advances and the new possibilities they offer for nonlinear optics and quantum information science. As a basis for ...
  7. [7]
    Self-Induced Transparency by Pulsed Coherent Light | Phys. Rev. Lett.
    Self-Induced Transparency by Pulsed Coherent Light. SL McCall and EL Hahn. Department of Physics, University of California, Berkeley, California.Missing: 1960s | Show results with:1960s
  8. [8]
    Ultraslow optical dephasing in E u 3 + : Y 2 S i O 5
    We report the measurement of what we believe to be the narrowest known optical resonance in a solid. This was made on the 7 F 0 → 5 D 0 transition of E u 3 + : ...Missing: EIT | Show results with:EIT
  9. [9]
    The Nobel Prize in Physics 2022 - NobelPrize.org
    The Nobel Prize in Physics 2022 was awarded jointly to Alain Aspect, John F. Clauser and Anton Zeilinger for experiments with entangled photons.
  10. [10]
  11. [11]
  12. [12]
    [2205.10959] A practical guide to electromagnetically induced ...
    May 22, 2022 · This tutorial introduces the theoretical and experimental basics of Electromagnetically Induced Transparency (EIT) in thermal alkali vapors.
  13. [13]
    Cold atomic media with ultrahigh optical depths | Phys. Rev. A
    Nov 5, 2014 · We present an experimental study to achieve ultrahigh optical depths for cold atomic media with a two-dimensional magneto-optical trap (MOT) of cesium.
  14. [14]
    Electromagnetically induced transparency over spectral hole ...
    We experimentally observed EIT in an optically thick, spectral hole-burning solid of Pr:YSO at temperatures higher than needed for spectral hole burning.
  15. [15]
    Highly efficient optical quantum memory with long coherence time in ...
    An ideal optical fiber loop has a loss of 50% in 100 μs, and until now no universal quantum memory has beaten this time efficiency limit. Here, we report ...
  16. [16]
    Rydberg electromagnetically induced transparency of vapor in a cell ...
    We have observed Rydberg EIT in a vapor cell containing 5 Torr of Ne buffer gas. Results obtained at low probe power have revealed frequency shifts of the EIT ...
  17. [17]
    Optimizing Slow and Stored Light via EIT in Alkali Vapor - Optica ...
    Even under idealized conditions of negligible spin decoherence there is a fundamental limit on storage efficiency due to the finite optical depth of an ...
  18. [18]
    Storage of Light in Atomic Vapor | Phys. Rev. Lett.
    Jan 29, 2001 · We report an experiment in which a light pulse is effectively decelerated and trapped in a vapor of Rb atoms, stored for a controlled period of time, and then ...Missing: seminal | Show results with:seminal
  19. [19]
    Coherent Optical Memory with High Storage Efficiency and Large ...
    The classical fidelity of the recalled pulse is better than 90% and nearly independent of the storage time, as confirmed by the direct measurement of phase ...Missing: group | Show results with:group
  20. [20]
    EIT-based quantum memory for single photons from cavity-QED
    May 11, 2011 · We investigate the feasibility of implementing an elementary building block for quantum information processing.
  21. [21]
    Optical quantum memory based on electromagnetically induced ...
    Electromagnetically induced transparency (EIT) is a promising approach to implement quantum memory in quantum communication and quantum computing applications.<|control11|><|separator|>
  22. [22]
    Impedance-matched cavity quantum memory | Phys. Rev. A
    Aug 11, 2010 · In conclusion, we have shown that impedance matching to an optical cavity allows the implementation of highly efficient quantum memories for an ...Missing: EIT | Show results with:EIT
  23. [23]
    Electromagnetically Induced Transparency in a Diamond Spin ...
    May 22, 2013 · In this Letter, we report EIT with low-defect-density diamond at zero magnetic field. Using electron-irradiation and annealing techniques [39]Missing: 1994 | Show results with:1994
  24. [24]
    Sub-Doppler laser cooling using electromagnetically induced ...
    May 9, 2017 · We propose a sub-Doppler laser-cooling mechanism that takes advantage of the unique spectral features and extreme dispersion generated by the phenomenon of ...Missing: μK | Show results with:μK
  25. [25]
    Temperature limits in laser cooling of free atoms with three-level ...
    Atoms are optically pumped into a stationary dark state at near zero velocity, leading to subrecoil temperatures. EIT cooling has been demonstrated using Λ ...
  26. [26]
    [1609.09105] Sub-Doppler Laser Cooling using Electromagnetically ...
    Sep 28, 2016 · We propose a sub-Doppler laser cooling mechanism that takes advantage of the unique spectral features and extreme dispersion generated by the phenomenon of ...Missing: 1999 Rb 1 μK
  27. [27]
    Tunable electromagnetically induced transparency in integrated ...
    In this paper, we propose a novel method towards tunable EIT in a fully integrated silicon photonics circuits consisting of a silicon ring resonator with two ...
  28. [28]
    Absorption and emission of single attosecond light pulses in an ...
    The EIT condition is controlled by the time delay between the two pulses. (b) The spectrum for t 0 = 10 fs shows the largest separation in the Autler-Townes ...
  29. [29]