Fact-checked by Grok 2 weeks ago

Purcell effect

The Purcell effect is the modification of the rate of an excited atom or quantum emitter to its with a resonant electromagnetic , which alters the density of modes available at the transition frequency. This leads to an enhancement of the emission rate when the cavity mode matches the atomic transition, quantified by the Purcell factor F_p = \frac{3}{4\pi^2} \frac{\lambda^3}{V} [Q](/page/Q), where \lambda is the emission wavelength, V is the cavity mode volume, and [Q](/page/Q) is the cavity quality factor; conversely, suppression can occur in cavities with modes mismatched to the transition. Proposed by Edward M. Purcell in 1946 as a means to accelerate probabilities at radio frequencies by confining electromagnetic fields in a resonant structure, the effect was initially theoretical but gained experimental traction in the 1970s through studies of fluorescence near reflecting surfaces. In the framework of , it demonstrates how the vacuum fluctuations of the —responsible for spontaneous emission in free space—can be engineered by boundaries, enabling reversible atom-photon interactions via Rabi oscillations in high-Q cavities. The Purcell effect has profound implications for controlling light-matter interactions at the quantum level, underpinning phenomena such as inhibited emission (where lifetimes extend dramatically below cavity cutoffs) and enhanced radiation rates exceeding 500-fold in tuned systems. It forms the basis for technologies including single-photon sources, processors, and micromasers, where precise manipulation of emission enhances efficiency in nanoscale optical devices like photonic crystals and cavities.

Fundamentals

Definition and Basic Principles

The spontaneous emission process describes the decay of an excited quantum emitter, such as an or , to its while emitting a into the surrounding electromagnetic . In free space, this decay occurs at a fixed rate determined by the emitter's and the broadband density of photonic modes available across all frequencies and directions. This rate, often quantified by the Einstein A , reflects the uniform vacuum fluctuations that stimulate the emission. When the emitter is placed within a , the structured electromagnetic environment fundamentally alters this process by confining into discrete modes, modifying the available photonic states near the emitter. The Purcell effect specifically denotes the resulting change in the rate, which can be enhanced or suppressed depending on how well the 's resonant modes align with the emitter's transition frequency and orientation. In resonant , where the mode frequency matches the emitter's transition, the emission rate into the mode increases due to the concentration of , while emission into non-resonant directions may be inhibited. This phenomenon arises within the broader framework of (), which explores light-matter interactions in confined spaces. At its core, the Purcell effect stems from the variation in the local density of optical states (LDOS)—the number of accessible modes per unit volume at the emitter's location—which directly governs the rate according to adapted to . In free space, the LDOS is isotropic and frequency-independent in the optical regime, yielding a baseline rate; cavities, however, can amplify the LDOS at specific frequencies by factors tied to the cavity's quality factor and mode volume, thereby accelerating decay when the emitter is positioned at an antinode of the resonant field. Conversely, mismatches in frequency or spatial mode profile lead to a reduced LDOS, slowing and enabling over radiative lifetimes. This modulation of the LDOS provides a foundational principle for engineering emitter dynamics in photonic structures.

Purcell Factor

The Purcell factor F_p quantifies the enhancement of an emitter's rate in a resonant relative to free space, serving as a key metric in . It is expressed by the formula F_p = \frac{3}{4\pi^2} \left( \frac{\lambda}{n} \right)^3 \frac{Q}{V}, where \lambda is the of the emitted , n is the of the surrounding medium, Q is the 's quality factor, and V is the effective mode volume of the resonant field. This expression, originally derived for radio-frequency transitions, has been adapted to optical regimes to predict emission rate modifications. The quality factor Q measures the resonance sharpness, defined as Q = \omega_0 / \Delta \omega, where \omega_0 is the resonant and \Delta \omega is the full-width at half-maximum linewidth; a higher Q indicates longer times before , enhancing the interaction duration between the emitter and the cavity . Conversely, the mode V quantifies the spatial confinement of the electromagnetic , typically on the order of (\lambda / n)^3 or smaller in optimized ; reducing V concentrates the field intensity at the emitter's position, amplifying the local density of optical states. The term (\lambda / n)^3 normalizes the enhancement to the cubic wavelength in the medium, ensuring the factor's dimensionless nature as a pure rate ratio. Maximum enhancement via the Purcell factor occurs under the condition of single-mode resonance, where the cavity mode precisely matches the emitter's transition frequency to maximize field overlap and minimize competing decay channels. In practical optical microcavities, such as Fabry-Pérot or whispering gallery resonators, F_p typically achieves enhancements of 10 to 100, scaling with advances in nanofabrication that yield high Q (10^4 to 10^6) and subwavelength V.

Historical Development

Edward Purcell's Original Contribution

Edward Mills Purcell, an American physicist renowned for his work in (NMR), shared the 1952 with for developing new methods for nuclear magnetic precision measurements. His insights into stemmed from wartime research on technology during World War II, where he led a group at the Massachusetts Institute of Technology's Radiation Laboratory focused on advancing microwave radar systems. This experience with resonant circuits and electromagnetic fields directly informed his later explorations in atomic and nuclear processes. In a seminal 1946 presented at the meeting, titled "Spontaneous Emission Probabilities at Radio Frequencies," Purcell addressed the decay rates of excited nuclear spins in the context of NMR experiments. He proposed that placing a nuclear spin system within a resonant —such as those used in radio-frequency circuits—could dramatically accelerate relaxation through an enhanced reaction field generated by the coil itself. For a typical resonating at the transition frequency with a quality factor of around 100, Purcell calculated that the rate would increase by a factor of approximately 50 compared to free space, effectively modifying the environment's feedback on the emitter. Purcell highlighted this mechanism by noting the role of the "reaction field," stating that "the reaction field of the coil... may increase the probability by a large factor," thereby altering the otherwise fixed rate of spontaneous decay. He further extended this concept by analogy to optical emission, predicting that atomic transitions could similarly experience accelerated decay rates when confined within a resonant tuned to the emission frequency, challenging the notion of as an intrinsic atomic property.

Evolution in Cavity QED

In the 1970s, the Purcell effect experienced a significant revival within , driven by investigations into interactions with microwave resonators. Pioneering work by Philippe Goy, , and collaborators at the demonstrated pulsed operation with low atomic thresholds, revealing strong coupling between highly excited atomic states and cavity modes that enhanced rates in line with Purcell's predictions. This resurgence shifted focus from Purcell's original context to quantum optical analogs, emphasizing the role of cavity boundaries in altering atomic decay dynamics. Central to this evolution was the integration of the into the Jaynes-Cummings model, a quantum framework describing the interaction between a two-level atom and a single quantized cavity mode. Originally formulated in 1963, the model provided a rigorous basis for understanding weak coupling regimes where the manifests as a perturbative modification of the emission rate, while also foreshadowing stronger interactions. By the early , Haroche's group experimentally confirmed these concepts through observations of inhibited and enhanced from Rydberg atoms in high-quality superconducting cavities, achieving up to a 500-fold and linking the effect to the distinction between weak (irreversible decay) and strong (reversible Rabi oscillations) coupling regimes. Key experimental milestones in the 1980s further solidified these theoretical advances. Notably, K. H. Drexhage's 1974 studies on fluorescent molecules near surfaces demonstrated distance-dependent modifications to emission lifetimes, offering classical validation of environmental influences on that paralleled cavity effects. Transitioning to quantum descriptions, these efforts culminated in the observation of vacuum Rabi splitting in 1992, where the normal modes of the coupled atom-cavity system split into distinct frequencies, extending the Purcell effect beyond rate enhancements to coherent, oscillatory energy exchange in the strong coupling limit. The 1990s saw the Purcell effect extend to solid-state systems through microcavity research by Yoshihisa Yamamoto and colleagues, who systematically altered excitonic spontaneous emission in planar dielectric semiconductor structures via applied electric fields, achieving tunable Purcell factors and paving the way for integrated photonic devices. This period marked a full conceptual shift from classical boundary-induced rate changes to quantum-coherent light-matter interactions, with the Jaynes-Cummings Hamiltonian unifying disparate regimes and enabling applications in quantum information processing.

Theoretical Derivation

Heuristic Approach

The heuristic approach to understanding the Purcell effect relies on applying to the spontaneous emission rate of an excited emitter, emphasizing the role of the modified density of electromagnetic states in a resonant cavity. According to , the transition rate \Gamma from an excited state to the ground state is given by \Gamma = \frac{2\pi}{\hbar} |\mu|^2 \rho(\omega), where \mu is the magnitude of the dipole matrix element, \hbar is the reduced Planck's constant, and \rho(\omega) is the density of photon states at the transition frequency \omega. This expression highlights that the emission rate is proportional to the available density of final states into which the emitter can decay. In free space, the density of states is broadband and given by \rho_0(\omega) = \frac{\omega^2}{\pi^2 c^3} (per unit volume in vacuum), leading to the familiar Einstein A coefficient for spontaneous emission. Within a resonant cavity, however, the electromagnetic environment confines the modes, causing \rho(\omega) to peak sharply at the cavity resonance frequency. For a high-quality-factor (Q) cavity supporting a single dominant mode, this peak can be heuristically approximated near resonance as a Lorentzian lineshape, with the on-resonance value \rho(\omega) \approx \frac{\omega^2 Q}{\pi c^3 V}, where V is the effective mode volume and Q = \omega / \Delta\omega is the quality factor, with \Delta\omega the resonance linewidth. This approximation arises from integrating the narrow Lorentzian contribution of the cavity mode over the frequency range, effectively concentrating the states within the small volume V and enhancing the local density by a factor proportional to Q / V. The Purcell factor F_p, defined as the ratio of the cavity-enhanced emission rate to the free-space rate (\Gamma / \Gamma_0 = F_p), then follows from the ratio of densities, adjusted for the enhanced per in the confined geometry and the alignment of the with the . Assuming the emitter is oriented parallel to the and the transition frequency matches the cavity resonance, a step-by-step yields the expression F_p = \frac{3 Q \lambda^3}{4\pi^2 n^3 V}, where \lambda = 2\pi c / \omega is the free-space and n is the of the medium filling the . The factor of 3 accounts for the averaging over random orientations in free space versus perfect alignment in the cavity . This assumes the weak-coupling , where the emitter-cavity is perturbative (coupling rate g \ll \kappa, \gamma_0, with \kappa = \omega / Q the cavity decay rate and \gamma_0 the free-space linewidth), and positions the point-like emitter at a field antinode for maximum enhancement.

Rigorous Formulation

The spontaneous emission rate \Gamma of an excited two-level atom with transition dipole moment \vec{\mu} located at position \vec{r}_0 in a structured electromagnetic environment is rigorously derived within macroscopic quantum electrodynamics (QED) as \Gamma = \frac{2}{\hbar} \operatorname{Im} [\vec{\mu}^* \cdot \vec{E}(\vec{r}_0, \omega)], where \omega is the transition frequency and \vec{E}(\vec{r}_0, \omega) is the electric field at the atom's position produced by the dipole itself, satisfying the appropriate boundary conditions of the environment. This expression arises from the Fermi's golden rule applied to the interaction Hamiltonian, accounting for vacuum fluctuations via the field's positive-frequency part, and reduces to the free-space rate \Gamma_0 = \frac{\omega^3 |\vec{\mu}|^2}{3\pi \epsilon_0 \hbar c^3} in the absence of boundaries. The field \vec{E}(\vec{r}_0, \omega) is obtained by solving the classical Helmholtz equation for a point dipole source, incorporating material dispersion and losses through the permittivity \epsilon(\vec{r}, \omega). A powerful framework for computing this field employs the dyadic Green's function \mathbf{G}(\vec{r}, \vec{r}', \omega), which propagates the field from source to observation point while enforcing boundary conditions: \vec{E}(\vec{r}_0, \omega) = \frac{\omega^2 \mu_0}{\epsilon_0} \mathbf{G}(\vec{r}_0, \vec{r}_0, \omega) \cdot \vec{\mu}. Substituting yields the decay rate \Gamma = \frac{2 \omega^2}{\hbar \epsilon_0 c^2} \vec{\mu} \cdot \operatorname{Im} [\mathbf{G}(\vec{r}_0, \vec{r}_0, \omega)] \cdot \vec{\mu}^*, where the imaginary part captures the local response of the environment. This connects directly to the projected local density of electromagnetic states (LDOS) \rho(\vec{r}_0, \omega) for dipole orientation \hat{u}, defined as \rho(\vec{r}_0, \omega) = \frac{6\omega}{\pi c^2} \operatorname{Im} [G_{ii}(\vec{r}_0, \vec{r}_0, \omega)], with the factor of 6 arising from averaging over isotropic orientations (trace over diagonal components). The LDOS quantifies the available photon modes at \vec{r}_0 and frequency \omega, modifying \Gamma = \Gamma_0 \frac{\rho(\vec{r}_0, \omega)}{\rho_0(\omega)} relative to the free-space value \rho_0(\omega) = \frac{\omega^2}{\pi^2 c^3}. In resonant cavities, the Green's function is expanded over the cavity eigenmodes \vec{e}_n(\vec{r}) satisfying \nabla \times \nabla \times \vec{e}_n = \frac{\omega_n^2}{c^2} \epsilon(\vec{r}) \vec{e}_n, normalized such that \int |\vec{e}_n(\vec{r})|^2 dV = V (effective volume): \mathbf{G}(\vec{r}, \vec{r}', \omega) = \sum_n \frac{\vec{e}_n(\vec{r}) \vec{e}_n^*(\vec{r}')}{\omega_n^2 - \omega^2 - i \omega_n \gamma_n}, where \gamma_n = \omega_n / Q_n is the mode linewidth. Near a single \omega \approx \omega_m, the dominant contribution simplifies the Purcell factor F_p = \Gamma / \Gamma_0 to F_p = \frac{6\pi c^3}{\omega^3} \frac{|\vec{E}(\vec{r}_0)|^2}{V} Q, with |\vec{E}(\vec{r}_0)|^2 = |\vec{e}_m(\vec{r}_0)|^2 and Q = \omega_m / \gamma_m, assuming normalized modes where the maximum field intensity relates to the volume. This form highlights the enhancement scaling with quality factor and inversely with volume, derived from boundary-induced mode confinement. Extensions to non-ideal cavities incorporate losses through complex \epsilon(\vec{r}, \omega) in the mode equation, reducing Q via absorption (\operatorname{Im} \epsilon > 0) and radiation leakage, as captured in the pole structure of \mathbf{G}. Multi-mode effects arise naturally from the full sum in the Green's expansion, requiring summation over all resonant contributions when mode spacing is comparable to linewidths, leading to broadband modifications of \rho(\vec{r}_0, \omega). In the bad-cavity limit (Q \ll \omega / \Delta \omega, where \Delta \omega is mode spacing), the enhancement approaches the heuristic density-of-states picture but remains rigorously tied to the imaginary part of the multi-mode Green's function.

Experimental Realizations

Early Demonstrations

The first experimental demonstration of the Purcell effect was reported in 1983 by Goy et al., who observed enhanced from Rydberg atoms of sodium interacting with a high-quality-factor superconducting . In their setup, atoms in the Rydberg state |23P_{3/2}> were prepared and passed through the cavity tuned to the atomic transition frequency, resulting in a measured shortening of the emission lifetime by a factor of up to 20 compared to free space, consistent with the predicted Purcell enhancement due to the increased density of electromagnetic modes in the cavity. As a precursor to cavity-based demonstrations, Drexhage's 1970 experiments explored modifications to fluorescence decay rates of organic molecules positioned at varying distances from interfaces, such as substrates or metallic mirrors. These studies revealed oscillatory variations in the emission lifetime as a function of emitter-mirror separation, attributable to effects that alter the local density of optical states, laying groundwork for understanding Purcell-like enhancements near boundaries without resonant cavities. In the , optical microcavity experiments with organic dyes provided further confirmations at visible wavelengths. Yokoyama et al. in 1991 reported a fourfold increase in the rate of 6G dye molecules within a Fabry-Pérot microcavity formed by mirrors, where the mode volume and quality factor were tuned to match the dye's around 590 nm. This enhancement, quantified via time-resolved measurements, directly illustrated the Purcell factor's role in directing emission into the mode, enabling low-threshold oscillation in the dye solution. Early demonstrations faced significant technical challenges, including the need for cryogenic temperatures to achieve and maintain high Q factors in superconducting cavities, as thermal noise at room temperature degraded performance. Additionally, optical cavities often suffered from relatively low Q values due to material and losses, limiting achievable enhancements, while precise measurement of lifetime shortening required advanced pulsed excitation and detection techniques to resolve sub-nanosecond dynamics amid .

Modern Implementations

In the , plasmonic nanocavities emerged as a key platform for achieving exceptionally high Purcell factors, often exceeding 1000, by tightly confining light around metal nanoparticles coupled to quantum dots. These structures leverage polaritons to dramatically increase the local density of optical states, accelerating rates from embedded emitters such as colloidal quantum dots at . For instance, hybrid plasmonic nanodisks with radii around 1000 nm have demonstrated Purcell factors up to 1827 at telecommunication wavelengths, enabling ultrafast emission on timescales without significant losses. Similarly, plasmonic nanoantennas integrated with quantum dots have realized factors near 940, facilitating single-photon emission with lifetimes shortened to approximately 10 ps. These advancements highlight the role of precise nanoparticle geometry and emitter positioning in maximizing enhancement while mitigating ohmic losses inherent to metallic systems. Photonic crystals and nanowires have provided another avenue for demonstrating the Purcell effect through precise control of emission inhibition and enhancement via engineered defects in structures. In a seminal experiment by Baba et al., slow-light propagation in line-defect enabled modulation of group velocities, leading to enhanced light-matter interactions that inhibit or accelerate depending on the photonic band structure. This work showcased Purcell factors varying by orders of magnitude, with inhibition reducing emission rates by factors of up to 10 in band-gap regions and enhancement boosting them in defect modes, laying groundwork for integrated nanoscale . Subsequent realizations in nanowire-embedded further refined this, achieving directional emission control through symmetry-breaking defects that funnel light into specific modes. In (QED), superconducting qubits coupled to microwave cavities have realized strong light-matter interactions since the mid-2000s, yielding Purcell factors exceeding $10^4. A 2007 demonstration by Blais and collaborators established strong dispersive in qubits integrated with resonators, where the qubit-cavity interaction strength g approached 100 MHz, resulting in emission rates into the cavity mode enhanced by over four orders of magnitude relative to dissipative channels. Ultra-strong regimes, with ratios g/\omega > 0.1, have been achieved in systems, suppressing unwanted Purcell decay while enabling coherent control of photon emission, as evidenced by vacuum Rabi splitting exceeding 200 MHz. Such systems have since become benchmarks for scalable quantum processors, with high-fidelity single-photon generation. More recent advancements, as of 2023, include Purcell enhancements exceeding $10^5 for color centers in hexagonal coupled to nanophotonic cavities, enabling efficient room-temperature quantum emitters for integrated . Contemporary measurements of the Purcell effect rely on advanced techniques to quantify lifetime modifications and emission statistics. , often via streak cameras or time-correlated single-photon counting, directly probes the shortened decay times of emitters in cavities, allowing computation of the Purcell factor as the ratio of enhanced to free-space lifetimes. Complementarily, Hanbury Brown-Twiss assesses directionality and quantum by correlating arrivals at two detectors, revealing antibunching (g^{(2)}(0) < 0.5) in Purcell-enhanced single-photon sources and confirming mode-selective emission into cavity waveguides. These methods ensure precise characterization of enhancement in nanostructured environments, distinguishing Purcell contributions from other radiative processes.

Applications

In Quantum Optics and Photonics

In quantum optics and photonics, the Purcell effect plays a pivotal role in engineering light-matter interactions at the single-photon level, enabling enhanced control over quantum emitters for advanced photonic technologies. By modifying the local density of optical states within cavities or resonators, the effect accelerates or suppresses spontaneous emission rates, thereby improving the efficiency and coherence of photon generation and manipulation. This capability is essential for developing scalable quantum systems that operate with high fidelity and low decoherence. For single-photon sources, the Purcell effect enhances the indistinguishability of emitted photons through cavity-modified emission dynamics in semiconductor quantum dots. In photonic crystal cavities, the Purcell enhancement factor can exceed 10, shortening the excited-state lifetime and reducing timing jitter, which boosts two-photon interference visibility to over 90% at telecom wavelengths. For instance, InAs quantum dots embedded in GaAs membranes coupled to ring resonators have demonstrated near-unity indistinguishability (>95%) under resonant excitation, facilitating applications in and . In quantum information processing, Purcell-enhanced strengthens atom-photon interfaces, crucial for that extend entanglement distribution over long distances. By integrating quantum dots with high-Q resonators, the effect amplifies the spin-photon rate, achieving Purcell factors above 100 to enable efficient two-qubit gates between and spins. This enhancement supports the creation of heralded entanglement between distant nodes via photon-mediated interactions, with demonstrated fidelities exceeding 80% in solid-state implementations. Neutral atoms trapped in fiber cavities have also shown sixfold Purcell broadening of emission lines, improving the efficiency of photon collection for protocols. Optical switching and transistors in the 2010s leveraged cavity-enhanced nonlinearities to realize low-power, single-photon-level operations. In waveguide-coupled quantum dots, resonant enables giant optical nonlinearities where a single control switches a signal with transmission contrast up to 30%. Demonstrations using resonantly driven quantum dots in photonic-crystal waveguides achieved switching at powers corresponding to below 1 per pulse, paving the way for energy-efficient all-optical quantum gates. Solid-state quantum memories further exploited this for transistor-like behavior, with input s modulating output by factors of 10 while consuming sub-femtowatt power. Integration with hybrid systems, such as diamond nitrogen-vacancy (NV) centers in optical resonators, exploits the for room-temperature . Coupling NV centers to microcavities enhances emission rates by factors up to 20, enabling coherent spin-photon interfaces with coherence times over 100 μs at ambient conditions. This facilitates platforms combining solid-state spins with photonic circuits for quantum sensing and networking, where Purcell-modified supports single-shot readout with >90% efficiency. Such systems bridge cryogenic and room-temperature regimes, advancing practical quantum technologies. As of 2025, further advancements include Purcell-enhanced single-photon sources at wavelengths for scalable quantum networks.

In Nanoscale Devices

The Purcell effect plays a crucial role in enhancing rates within nanoscale devices, particularly in vertical-cavity surface-emitting lasers (VCSELs) and light-emitting diodes (LEDs). By integrating emitters into high-quality-factor microcavities, the effect reduces lasing thresholds and boosts through increased coupling of into the lasing mode. For instance, in GaN-based VCSELs with shortened cavity lengths, the Purcell factor scales inversely with cavity volume, leading to factors up to 2.9 × 10⁻² and slope improving by over an (e.g., from 1 to 15 mW/A), which demonstrates threshold reductions from 6.3 to 1.2 mJ/cm². Similarly, in 2D -activated VCSELs using WS₂ positioned at the cavity antinode, the Purcell enhancement achieves a coupling factor of 0.77, enabling room-temperature lasing thresholds as low as 0.44 W/cm² and significant gains in ultrathin structures. In plasmonic nanoscale lasers, the Purcell effect facilitates the realization of spasers (surface plasmon amplification by stimulated emission of radiation), which enable coherent light generation at subwavelength scales. A seminal demonstration involved 44-nm gold-core nanoparticles coated with a dye-doped silica shell, where optical gain from the dye molecules compensated for plasmonic losses, achieving stimulated emission of surface plasmons at 531 nm and outcoupling to photonic modes for nanolasing. This approach leverages the high local density of states in plasmonic resonators to amplify emission, paving the way for compact, thresholdless lasers integrated into nanoscale circuits. The Purcell effect also modifies emission rates in plasmonic nanostructures for biosensing and photovoltaic applications, enhancing and light harvesting efficiency. In biosensors, fluorophores to plasmonic nanocavities or arrays boosts via the Purcell factor, enabling substantial enhancement with high directionality for single-molecule detection. For , plasmonic resonances in metal -semiconductor hybrids increase the local density of optical states, enhancing and ; plasmonic arrays coupled to dye-sensitized cells improve and efficiency by modifying emitter decay rates. Despite these advances, implementing the Purcell effect in nanoscale devices faces significant challenges, including quenching losses from nonradiative recombination in metallic structures and difficulties in scaling to integrated circuits. In plasmonic nanocavities, proximity to metal surfaces (e.g., <2 nm) can quench emission by up to 25% via nonradiative channels, though dielectric spacers like Al₂O₃ mitigate this to achieve quantum yields up to 62% in excitons with Purcell factors of 180. Scalability issues arise from fabrication tolerances in subwavelength cavities, thermal management, and maintaining high quality factors amid losses, limiting integration into dense photonic circuits as device sizes approach fundamental limits where Purcell enhancement saturates.

References

  1. [1]
  2. [2]
    Spontaneous emission in micro- or nanophotonic structures - PhotoniX
    Sep 16, 2021 · According to Purcell's theory, the SE of quantum emitters is controlled by the optical modes of the cavity [1]. To engineer the optical modes in ...<|control11|><|separator|>
  3. [3]
  4. [4]
    An antenna model for the Purcell effect | Scientific Reports - Nature
    Aug 10, 2015 · The Purcell effect is defined as a modification of the spontaneous emission rate of a quantum emitter at the presence of a resonant cavity.
  5. [5]
  6. [6]
    The Nobel Prize in Physics 1952 - NobelPrize.org
    The Nobel Prize in Physics 1952 was awarded jointly to Felix Bloch and Edward Mills Purcell for their development of new methods for nuclear magnetic precision ...
  7. [7]
    Edward Purcell - Magnet Academy - National MagLab
    Edward Mills Purcell was an American physicist who received half of the 1952 Nobel Prize for Physics for his development of a new method of ascertaining the ...
  8. [8]
  9. [9]
    [PDF] EM Purcell "Spontaneous emission probabilities at radio frequencies"
    Page 1. E. M. Purcell "Spontaneous emission probabilities at radio frequencies" Phys. Rev. 69, 681 (1946)
  10. [10]
    [PDF] Serge Haroche - Nobel Lecture: Controlling Photons in a Box and ...
    A simple theoretical model, introduced by Jaynes and Cummings [3] in the early days of laser physics, applies to both situations and contributes to unify. Page ...
  11. [11]
  12. [12]
  13. [13]
  14. [14]
    Observation of Cavity-Enhanced Single-Atom Spontaneous Emission
    Jun 13, 1983 · It has been observed that the spontaneous-emission lifetime of Rydberg atoms is shortened by a large ratio when these atoms are crossing a high- Q ...
  15. [15]
    Spontaneous emission and laser oscillation properties of ...
    Jun 10, 1991 · An experimental study on spontaneous and stimulated emission properties of planar optical microcavities confining an organic dye solution is reported.
  16. [16]
    Subwavelength hybrid plasmonic nanodisk with high Q factor and ...
    Apr 28, 2011 · For a hybrid plasmonic nanodisk with the radius of 1000 nm, a Q factor of 819 and a Purcell factor of 1827 are achieved at the telecommunication wavelength of ...<|separator|>
  17. [17]
    Ultrafast Room-Temperature Single Photon Emission from Quantum ...
    Nov 25, 2015 · We demonstrate the regime of ultrafast spontaneous emission (∼10 ps) from a single quantum emitter coupled to a plasmonic nanocavity at room temperature.
  18. [18]
    Ultrafast spontaneous emission source using plasmonic ... - Nature
    Jul 27, 2015 · ... Purcell factors of <145 have been demonstrated. Higher Purcell factors of up to 1,000 have been obtained for molecules ...
  19. [19]
    Single quantum-dot Purcell factor and 𝛽 factor in a photonic crystal ...
    May 24, 2007 · A theoretical formalism to calculate the spontaneous emission rate enhancement (Purcell factor) and propagation mode β factor from single quantum dotsMissing: nanowires | Show results with:nanowires
  20. [20]
    Circuit quantum electrodynamics | Rev. Mod. Phys.
    May 19, 2021 · The prospect of realizing strong coupling of superconducting qubits to photons stored in high-quality coplanar waveguide resonators, together ...
  21. [21]
    (PDF) Ultra-strong coupling regime of cavity QED with flux qubits
    PDF | With improved dephasing rate and coupling strength, the transmon qubit has recently been used to reach the strong coupling regime of cavity QED.
  22. [22]
    Continuous-wave versus time-resolved measurements of Purcell ...
    The Purcell effect using QDs as emitters has first been observed when coupled to pillar type microcavities in the late 1990s. 13. Moreover, when its ...
  23. [23]
    Strong Purcell Effect on a Neutral Atom Trapped in an Open Fiber ...
    Oct 23, 2018 · A strong Purcell enhancement in such open-cavity platforms can only be realized by using microresonators with reduced mode volume [19] or ...
  24. [24]
    Purcell effect in photonic crystal microcavities embedding InAs/InP ...
    The spontaneous emission rate and Purcell factor of self-assembled quantum wires embedded in photonic crystal micro-cavities are measured at 80 K by using micro ...2.1. Qwr Epitaxy · 5. Discussion · 5.2 Polarization Mismatch
  25. [25]
    Purcell-enhanced single photons at telecom wavelengths from a ...
    Feb 23, 2024 · Their suitability for integration into photonic structures allows for enhanced brightness through the Purcell effect, supporting efficient ...Missing: Vlasov | Show results with:Vlasov
  26. [26]
    [2007.12920] Purcell enhanced and indistinguishable single-photon ...
    Jul 25, 2020 · This integrated single-photon source can be readily scaled up, promising a realistic pathway for scalable on-chip linear optical quantum ...
  27. [27]
    [PDF] Quantum repeaters based on individual electron spins and nuclear ...
    Nov 2, 2021 · Purcell factors in excess of 100 are required for a high-performance implementation of the two- qubit gates, and as such, the QD should only be.
  28. [28]
    A single-photon switch and transistor enabled by a solid-state ...
    Jul 6, 2018 · Single-photon switches and transistors generate strong photon-photon interactions that are essential for quantum circuits and networks.
  29. [29]
    Single-photon non-linear optics with a quantum dot in a waveguide
    Oct 23, 2015 · Here we show that a single quantum dot in a photonic-crystal waveguide can be used as a giant non-linearity sensitive at the single-photon level.
  30. [30]
    Cavity-assisted resonance fluorescence from a nitrogen-vacancy ...
    Nov 7, 2024 · ... Purcell factor of ~1.8 increases the zero-phonon line fraction to ... optical transitions20,21,22,23,24,25,26,27,28,29. A spin-photon ...
  31. [31]
    Photophysics of single nitrogen-vacancy centers in nanodiamonds ...
    Nov 22, 2020 · Here we report on modifying the internal quantum efficiency of a single NV center in a nanodiamond coupled to a 1D photonic crystal cavity.
  32. [32]
    On the importance of cavity-length and heat dissipation in GaN ...
    Apr 15, 2015 · Where Fp is the Purcell factor and Vc is the effective optical volume of the laser. ... These results revealed that the performance of the VCSELs ...
  33. [33]
    Room-temperature 2D semiconductor activated vertical-cavity ...
    Sep 14, 2017 · Moreover, introducing Purcell effect by precisely positioning a 2D semiconductor at the antinode of the cavity standing wave along with the ...
  34. [34]
    Demonstration of a spaser-based nanolaser - Nature
    ### Summary of Key Findings on Spaser-Based Nanolaser (Noginov et al., 2009)
  35. [35]
    Purcell-enhanced quantum yield from carbon nanotube excitons ...
    Nov 10, 2017 · Here we demonstrate SWCNT excitons coupled to plasmonic nanocavity arrays reaching deeply into the Purcell regime with Purcell factors (F P) up ...Results · Time-Integrated Measurements · Time-Resolved Measurements