Fact-checked by Grok 2 weeks ago

Slow light

Slow light refers to the propagation of light pulses through a medium at a greatly reduced , often orders of magnitude slower than the in vacuum (c ≈ 3 × 10⁸ m/s), due to strong that increases the group refractive index (ng) while maintaining transparency. The is defined as vg = c / ng, where ng = n + ω(dn/dω*) and n is the phase ; this reduction arises from coherent light-matter interactions or structural resonances that steepen the without significant absorption. The phenomenon was first experimentally demonstrated in 1999 by Lene V. Hau and colleagues, who used (EIT) in an ultracold Bose-Einstein condensate of sodium atoms to slow light to approximately 17 m/s, compressing a pulse spatially while preserving its shape. Subsequent advances included stopping light entirely in 2001 via dynamic EIT in atomic vapors, enabling reversible storage of optical information as atomic spin coherence. Key mechanisms for generating slow light include EIT and coherent population oscillations (CPO) in atomic and solid-state media, stimulated Brillouin and (SBS/SRS) in optical fibers, and photonic bandgap structures such as waveguides and crystals that engineer dispersion through geometry. In waveguides, for instance, SBS has achieved delays up to 25 ns over 25 GHz bandwidths, while EIT in gas-filled hollow-core fibers has demonstrated 800 ps delays at telecom wavelengths. Slow light holds significant promise for applications in and quantum technologies, including optical buffers for all-optical and in , where it compensates for timing mismatches without electronic conversion. It enhances nonlinear optical effects by increasing light-matter interaction time, enabling compact devices for frequency conversion and ; for example, group indices exceeding 108 amplify phase shifts for sensitive and with resolutions down to 15 MHz. In 2022, plasmonic structures with transition metal dichalcogenides realized on-chip slow light with slowdown factors of ~1300 at . As of 2025, advances include AI-accelerated silicon photonic slow-light technology enabling 400 Gbps/λ transmission, paving the way for integrated sensors and topological .

Fundamentals

Definition and basic principles

Slow light refers to the phenomenon in which the of a is dramatically reduced, often to speeds of a few meters per second or less, while propagating through a medium, without violating the principles of that prohibit information transfer faster than the in (). This reduction occurs because the , which determines the speed of the carrying the , can be slowed by the medium's properties, whereas no single photon's speed exceeds . The underlying principle stems from the steep dispersion of the medium's refractive index (n) near absorption lines, creating a narrow transparency window where light can propagate with minimal loss but experiences a sharp variation in n with frequency. In such regions, the frequency-dependent refractive index causes different spectral components of the pulse to travel at varying speeds, effectively compressing the pulse and delaying its overall transit. This dispersion-engineered slowdown enhances light-matter interactions without fundamentally altering the phase velocity of individual wave components. To understand slow light, it is essential to distinguish between (v_p = c / n), which describes the of a wave's constant-phase surfaces and can exceed c in some media, and the (v_g), which governs the transport of the or and is the relevant quantity for slow light effects. While phase velocity changes do not affect information , reductions in v_g arise from the dispersive term in the , ensuring that is preserved as the signal velocity remains below c. A basic example involves a short entering a highly dispersive medium: the frequencies near the window slow more than those at the trailing edge, reshaping the while imposing a measurable delay on its arrival at the output. The conceptual foundations of such dispersion effects trace back to 19th-century studies on wave propagation and by researchers like and , though the modern experimental realization and understanding of emerged in the through advances in .

Group velocity and dispersion relations

The group velocity v_g, defined as the derivative of the \omega with respect to the wave number k, v_g = \frac{d\omega}{dk}, represents the speed at which and propagate through a dispersive medium. In the context of slow light, this velocity is significantly reduced compared to the in vacuum c, arising from the steep slope of the \omega(k) near points of high . Near a resonance frequency, the can be approximated as n(\omega) \approx n_0 + \frac{dn}{d\omega} \Delta \omega, where n_0 is the background and \Delta \omega = \omega - \omega_0 is the detuning from the \omega_0. This linear expansion leads to the expression v_g = \frac{c}{n + \omega \frac{dn}{d\omega}}, where a large positive value of the term \frac{dn}{d\omega} in regions of normal (where \frac{dn}{d\omega} > 0) results in a substantial reduction of v_g. Equivalently, the is given by v_g = \frac{c}{n_g}, with the group index n_g = n + \omega \frac{dn}{d\omega}; a plot of n(\omega) near typically shows a steep positive , corresponding to high n_g and thus slow light . The Kramers-Kronig relations play a crucial role by linking the real part of the (dispersion) to the imaginary part (), ensuring that steep dispersion features are accompanied by narrow absorption lines. These relations, derived from principles, impose constraints that prevent superluminal signaling, as the reduction in slow light reflects a coherent redistribution of rather than a violation of . Consequently, slow light does not imply that energy is stored in the medium or that relativity is breached, since the can exceed c in such dispersive environments without carrying information.

Historical development

Early theoretical foundations

The theoretical foundations of slow light originated in 19th-century investigations into optical , which provided the essential framework for understanding how material resonances influence propagation speeds. In 1871, Wilhelm Sellmeier derived a dispersion formula relating the square of the n(\omega) to the \omega of , expressed as n^2(\omega) = 1 + \sum_i \frac{B_i}{1 - (\omega / \omega_i)^2}, where B_i are coefficients and \omega_i represent resonance frequencies of the material. This equation modeled the frequency-dependent refractive index near electronic resonances, establishing that strong dispersion could significantly alter the group velocity v_g = (dk/d\omega)^{-1}, where k is the wave number, thereby enabling reductions below the in vacuum. Early 20th-century advancements built on this by examining interactions in dispersive environments. In 1922, theoretically described the inelastic scattering of by acoustic phonons, predicting frequency shifts and delays in pulse propagation through with thermal density fluctuations. His 1926 analysis further explored scattering processes and the dynamics of delayed pulses in dispersive , illustrating how phonon-mediated interactions could mimic acoustic wave behavior and prolong signal transit times. Independently, Leonid Mandelstam proposed in 1926 that scattering from acoustic waves in crystals would produce similar effects, coining the basis for Mandel'shtam-Brillouin scattering as an early conceptual precursor to controlled light slowing via coherent wave couplings. Brillouin's broader contributions emphasized "acoustic-like" light propagation, where optical waves in periodic or resonant structures exhibit behaviors analogous to sound waves, including reduced effective velocities due to bandgap effects and scattering. During the 1960s, emerging quantum optics research on coherent forward scattering in atomic systems began to predict group velocity reductions through constructive interference of scattered fields, extending classical dispersion ideas to quantum-coherent regimes without absorption losses. Although the explicit term "slow light" emerged only in the late 20th century, these electromagnetism and wave theory foundations demonstrated the inherent feasibility of achieving sub-luminal group velocities via tailored dispersion. In the 1970s, detailed studies of pulse propagation in absorbing media quantified such effects, with Garrett and McCumber's 1970 analysis of Gaussian light pulses near anomalous dispersion lines predicting delays corresponding to group velocity reductions up to three orders of magnitude slower than in vacuum, attributable to resonant absorption reshaping the pulse envelope.

Key experimental milestones

A foundational experimental advance for slow light was the demonstration of electromagnetically induced transparency (EIT) by Boller, İmamoğlu, and Harris in 1991, who rendered an optically thick hot rubidium vapor transparent to a probe laser, creating a window of steep dispersion without absorption that enabled subsequent group velocity reductions. The first observation of slow light using EIT followed in 1995, when Kwong, Shore, and Harris reported propagation dynamics in lead vapor, measuring reduced group velocities due to the induced transparency and dispersion. In 1999, Lene Hau's group at Harvard University achieved a groundbreaking slowdown of light pulses to a group velocity of 17 m/s in an ultracold gas of sodium atoms using EIT, representing a reduction by a factor of approximately 10^7 from the in vacuum; this experiment, published in , highlighted the potential for further manipulations such as stopping entirely and garnered widespread recognition for the field. Building on these advances, Kash et al. demonstrated ultraslow light propagation in a room-temperature rubidium vapor cell in 1999, achieving a group velocity of 90 m/s through coherent driving of a three-level atomic system, proving that extreme slowdowns were possible without cryogenic cooling. In 2001, Hau's extended these techniques to halt light pulses completely in an ultracold sodium Bose-Einstein condensate, effectively storing the optical information as a spin coherence before releasing it unchanged, with effective velocities approaching zero during the storage phase; this work explored applications in Bose-Einstein condensates, where light speeds as low as ~10 m/s were realized in subsequent refinements using similar coherent control methods. Shifting from atomic media to solid-state structures, in 2002, Thomas Krauss and coworkers reported the first observation of slow light in photonic crystal waveguides, where engineered band structures near the Brillouin zone edge reduced group velocities by factors of up to 20 in gallium arsenide-based devices, paving the way for compact integrated slow-light platforms. By 2005, demonstrations of coupled resonator optical waveguides () achieved slowdown factors exceeding 10^4, as shown in experiments with arrays of high-Q microrings or microspheres, where pulse delays reached hundreds of picoseconds over millimeter scales due to the cumulative phase shifts across resonators. By 2009, integrated silicon slow-light devices had matured, with photonic crystal waveguides on silicon-on-insulator platforms demonstrating group delays of up to 50 nanoseconds over centimeter-length propagation paths, corresponding to slowdown factors of around 1500 while maintaining low propagation losses below 10 dB/cm.

Techniques for achieving slow light

Electromagnetically induced transparency in

Electromagnetically induced transparency (EIT) is a quantum observed in , where a weak probe laser field interacts with a three-level system in a Lambda configuration, consisting of two ground states and one excited state. A strong control laser field couples one ground state to the excited state, leading to quantum interference that suppresses absorption of the probe field at resonance, creating a narrow transparency window in the absorption spectrum. This window is accompanied by a steep positive dispersion, resulting in a dramatically reduced group velocity for the probe light pulse. The underlying mechanism involves the control field dressing the atomic states, which splits the excited state absorption line into an Autler-Townes doublet due to the AC Stark effect. Near the center of the transparency window, the refractive index slope reaches values of \frac{dn}{d\omega} \approx 10^8 s, enabling significant slowdown of light propagation through the medium. This steep dispersion arises from the coherent superposition of atomic states, where the probe and control fields induce a dark state that does not couple to the excited state, minimizing decoherence and absorption. The group delay experienced by a probe pulse of length L in the medium is given by \tau_g = \frac{L}{v_g} = \frac{n_g L}{c}, where v_g is the , c is the in , and the group index n_g \approx \omega \frac{dn}{d\omega} near the transparency window for optical frequencies \omega. In experiments with (Rb) or cesium (Cs) atomic vapors, this has resulted in slowdowns to velocities of 10–100 m/s, corresponding to group delays on the order of microseconds for centimeter-scale cells. For instance, in ultracold sodium gas, a velocity of 17 m/s was achieved, demonstrating the scalability to warm vapors like Rb and Cs with similar reductions. Typical experimental setups employ co-propagating probe and control laser fields tuned to the D2 line transitions in Rb or Cs vapor cells at room temperature, often with buffer gases to reduce transit-time broadening. The EIT transparency window has a bandwidth of approximately 1–10 MHz, with residual absorption minimized to less than 1%, allowing high-fidelity transmission of probe pulses. These configurations use extended cavities or free-space propagation to overlap the beams precisely, enabling observation of the dispersion-induced pulse compression and delay. A distinctive feature of EIT in atomic media is the ability to realize "stored light," where turning off the control field maps the propagating probe pulse onto a long-lived atomic spin coherence via coherent population trapping (CPT) in the dark state. This process halts light propagation entirely, storing the optical information as a collective atomic excitation that can be retrieved by reapplying the , with storage times limited by atomic decoherence rates.

Dispersion in photonic structures

Photonic crystals are artificial dielectric structures with a periodic variation in refractive index on the scale of the optical wavelength, creating photonic bandgaps analogous to electronic bandgaps in semiconductors. Within these structures, slow light arises near the edges of the photonic bands, where the dispersion relation \omega(k) becomes nearly flat, leading to a group velocity v_g = d\omega / dk \to 0. This flattening occurs particularly at the boundary, where the wavevector k approaches \pi / a (with a the lattice period), and the band structure can be approximated as \omega(k) \approx \omega_0 + \alpha (k - \pi/a)^2, resulting in vanishing slope and thus reduced v_g. The slowdown factor, defined as the group index n_g = c / v_g (where c is the speed of light in vacuum), can reach values of 100 to 1000 in optimized designs, enabling significant pulse delays over compact lengths. The mechanism relies on classical dispersion engineering through the periodic modulation, which modifies the effective medium properties and local density of optical states without requiring active quantum effects. A key early experimental demonstration occurred in 2002 using GaAs-based two-dimensional photonic crystal slabs, where a group velocity reduction to approximately c/40 (corresponding to n_g \approx 40) was observed at room temperature via time-of-flight measurements near the band edge. These structures operate passively at ambient conditions and have been integrated with silicon photonics platforms using CMOS-compatible fabrication, allowing for on-chip slow-light devices with tunable dispersion via thermal or electro-optic means. Metamaterials, often incorporating plasmonic elements or negative-index components within photonic-like periodic arrays, achieve slow light by enhancing the local density of states and tailoring dispersion through subwavelength resonances. For instance, in structures with coupled defect cavities embedded in two-dimensional photonic lattices, light propagation is slowed by resonant coupling that flattens the effective dispersion curve, similar to band-edge effects but with added control over negative refractive indices. A notable theoretical proposal involves adiabatically tapering negative-index metamaterial waveguides to trap and store light as a "rainbow" spectrum, potentially stopping different wavelengths at distinct positions. A critical figure of merit for practical slow-light devices in these structures is the product of the slowdown factor S = n_g and the operational bandwidth BW, denoted as S \times BW, which quantifies the achievable delay while minimizing losses from dispersion-induced broadening or scattering. Optimization of this metric, often targeting values exceeding 50 in photonic crystal waveguides, balances enhanced light-matter interactions against propagation losses, enabling applications in integrated optics.

Coupled resonator and waveguide methods

Coupled resonator optical waveguides (CROWs) consist of a one-dimensional chain of identical high-quality-factor resonators, such as microrings or microspheres, linked by evanescent coupling between nearest neighbors. This discrete structure enables light propagation through successive tunneling of electromagnetic energy from one resonator to the next, analogous to in a tight-binding model from The propagation characteristics are governed by the derived from the tight-binding approximation: \omega(k) = \omega_0 - \kappa \cos(ka) where \omega_0 is the resonant frequency of an isolated resonator, \kappa is the inter-resonator coupling coefficient, k is the Bloch wave number, and a is the distance between adjacent resonators. The group velocity v_g = d\omega/dk follows as v_g = \kappa a \sin(ka), which varies with k and approaches zero near the band edges (k = 0 or k = \pi/a), resulting in significant slowdown where v_g \ll c/n (with c the speed of light in vacuum and n the refractive index). This band-edge slowdown can achieve group velocity reductions by factors of $10^3 to $10^4, as demonstrated in experiments with chains of microring resonators and coupled microspheres. An experimental demonstration of a CROW occurred in 2006 using polymer-based microring resonators to verify the predicted dispersion and pulse delay. CROWs are compatible with CMOS fabrication processes, enabling compact, chip-scale devices integrated into silicon photonics platforms, particularly for telecom wavelengths around 1550 nm, with advancements prominent in the 2010s. Compared to electromagnetically induced transparency methods, CROWs offer broader operational bandwidths on the order of GHz, supporting higher data rates while maintaining low dispersion over short propagation distances. A related configuration is the side-coupled integrated spaced sequence of resonators (SCISSOR), where a linear waveguide is periodically side-coupled to a sequence of resonators, inducing strong dispersion through resonant feedback into the bus waveguide. Like CROWs, SCISSORs exploit evanescent coupling to reduce group velocity near resonance, but the side-coupling geometry allows for easier input/output coupling and reduced intrinsic losses in integrated formats. Experimental realizations in silicon-on-insulator platforms have shown effective slow-light propagation with velocity reductions comparable to CROWs, suitable for on-chip applications. Waveguide-based methods for slow light include stimulated Brillouin scattering (SBS) in optical fibers, where a pump wave interacts with a counter-propagating signal via electrostrictive coupling to acoustic phonons, creating a narrowband gain resonance that modifies the dispersion. This phonon-mediated process induces a tunable group delay proportional to the pump intensity, with demonstrations achieving delays up to tens of nanoseconds in standard single-mode fibers at telecom wavelengths under modest pump powers (around 1 W). Longer delays, extending to microseconds, have been reported using high-nonlinearity fibers or recirculating configurations to enhance the effective interaction length. SBS offers advantages in fiber-compatible systems, with bandwidths limited by the acoustic phonon lifetime (typically ~10 MHz) but scalable through multi-pump schemes.

Coherent population oscillations and stimulated Raman scattering

Coherent population oscillations (CPO) provide another mechanism for slow light in atomic and solid-state media, particularly at room temperature. In CPO, a strong pump field modulates the population in a two-level system, creating a refractive index grating that leads to dispersion without requiring a third level as in . This results in group velocities reduced to ~100 m/s in ruby crystals, with applications in passive optical buffers. Stimulated Raman scattering (SRS) in optical fibers or gases generates slow light through nonlinear interaction between pump, Stokes, and probe fields, producing gain and dispersion similar to SBS but via optical phonons or molecular vibrations. SRS has achieved pulse delays of several pulse widths in hydrogen-filled hollow-core fibers, offering broader bandwidths (~GHz) compared to SBS for telecommunication applications.

Applications

Optical buffering and delay lines

Optical buffering refers to the temporary storage of optical signals using slow light techniques, enabling all-optical memory without the need for optoelectronic conversion, which is essential for high-speed data processing in photonic systems. In electromagnetically induced transparency (EIT) media, light pulses can be stored for durations ranging from nanoseconds to microseconds by mapping the optical information onto atomic coherences in the medium. A seminal demonstration involved storing a light pulse in a vapor cell for up to 0.5 ms, showcasing the potential for compact, low-power buffers in future optical (RAM). This approach leverages the reversible mapping of light onto long-lived atomic states, allowing retrieval on demand with high fidelity. Recent advances include AI-accelerated silicon photonic slow-light technology enabling 400 Gbps/λ transmission and beyond in telecommunications as of 2025. Delay lines based on slow light provide tunable temporal control of optical signals through engineered dispersion, facilitating applications in packet switching and signal synchronization. By reducing the group velocity via structured media, these lines compress propagation times into much shorter physical lengths compared to conventional fiber delays. For instance, in silicon photonic crystal waveguides, a 2005 experiment achieved a group velocity reduction to approximately c/300 over sub-millimeter lengths, resulting in effective delays on the order of tens of nanoseconds while maintaining low loss, suitable for on-chip optical packet routing. Such compact devices enable tunable delays by dynamically adjusting the dispersion, as demonstrated in coupled resonator optical waveguides (CROWs), where arrays of resonators induce slow light propagation for buffering in telecommunications networks. In telecommunications, slow light delay lines support synchronization of optical packets in wavelength-division multiplexing systems, reducing contention and improving network efficiency without electronic buffering. Fiber-based stimulated Brillouin scattering (SBS) techniques offer another avenue, achieving delays up to tens of nanoseconds in standard single-mode fibers at telecom wavelengths, with potential extension to longer durations in extended fiber lengths for applications like radar signal processing. A key performance metric for practical optical buffers is the delay-bandwidth product, ideally exceeding 100 to support broadband signals without significant distortion; challenges such as pulse broadening are mitigated through linear dispersion engineering to preserve signal integrity.

Enhanced light-matter interactions

Slow light significantly enhances light-matter interactions by prolonging the time photons spend within a nonlinear medium, thereby increasing the effective interaction length and amplifying third-order nonlinear optical effects governed by the susceptibility \chi^{(3)}. This prolongation arises because the reduced group velocity v_g extends the photon lifetime, leading to an effective length L_\mathrm{eff} = L \cdot (c / v_g), where L is the physical length of the medium and c is the in vacuum. In engineered structures such as waveguides with group indices n_g = c / v_g exceeding 100, this can result in nonlinearity enhancements by factors up to $10^4, as the interaction strength scales with n_g^2 for processes like . One key application is all-optical switching, where slow light enables efficient phase modulation at milliwatt or lower powers. For instance, in electromagnetically induced transparency (EIT) media using rubidium atoms, coherent preparation allows a control pulse to induce switching of a signal pulse with an efficiency of 55%, using ultralow energies of approximately 0.3 pJ per pulse (corresponding to roughly $10^3 photons). This demonstrates the potential for π-phase shifts in compact systems, far below typical thresholds required in conventional nonlinear media. In sensing applications, slow light in photonic crystal structures boosts refractive index detection by enhancing the overlap between the evanescent field and the coupled with high quality factors (Q-factors) on the order of $10^5. This results in sensitivities approaching $1.9 \times 10^5 nm/RIU, enabling precise measurements of biomolecular changes in liquids or gases. Such high sensitivities stem from the slowed propagation, which amplifies resonance shifts due to index perturbations. Slow light also improves Raman amplification and related nonlinear processes. In silicon photonic crystal waveguides, enhanced stimulated Raman scattering has been observed, with gain coefficients increased by factors proportional to the group index slowdown. Demonstrations in the 2010s further showed slow-light-enhanced four-wave mixing for efficient wavelength conversion, achieving up to 12 dB improvement in conversion efficiency near band edges with group indices around 80. The nonlinear phase shift in these systems follows \phi_\mathrm{NL} \propto n_2 I L_\mathrm{eff} / \lambda, where n_2 is the nonlinear refractive index, I is the intensity, and \lambda is the wavelength, underscoring how L_\mathrm{eff} drives the amplification.

Challenges and future prospects

Fundamental limitations

Slow light phenomena are constrained by fundamental physical principles, including the requirements of causality and , which prohibit information transfer faster than the in , c. Although the v_g can be reduced dramatically in dispersive media, the —governing the propagation of information—remains bounded by c. This is ensured by the presence of Sommerfeld and Brillouin precursors, which are low-amplitude forerunners traveling at c ahead of the main pulse; these precursors carry negligible energy but maintain the causal structure, with approximately 99% of the pulse energy arriving within a time window corresponding to propagation at c. Absorption imposes another key limitation, as dispersive transparency windows enabling slow light, such as those in electromagnetically induced transparency (EIT), exhibit residual loss with absorption coefficient \alpha > 0. This attenuates the pulse over propagation distance L, limiting the maximum delay to T_\text{del} \approx \frac{L}{c} (n_g - 1), where n_g is the group index, and constraining L \leq 1/\alpha for detectable transmission. Compensation using gain media can offset this loss but introduces quantum noise, degrading signal fidelity and setting a practical floor on achievable slowdown factors. The bandwidth of slow light is fundamentally limited by dispersion relations derived from the Kramers-Kronig theorem, which links refractive index changes to absorption profiles. The slowdown factor S = c / v_g \approx \omega \, dn/d\omega (for large ) implies a bandwidth \Delta \omega \approx 1 / (dn/d\omega), such that the product S \cdot \Delta \omega remains roughly constant, typically on the order of the linewidth. This restricts broadband operation; for instance, in EIT systems, the delay-bandwidth product is bounded by atomic coherence times, preventing simultaneous large delays and wide spectral support without distortion. Theoretically, v_g > 0 in any medium, but practical realizations approach a floor around 10 m/s in ultracold atomic gases via EIT, as demonstrated by reducing light speed to 17 m/s in a sodium Bose-Einstein condensate. However, decoherence processes in EIT, such as atomic collisions or spontaneous emission, limit light storage times to milliseconds, further bounding utility for long delays. Higher-order dispersion introduces pulse distortion, particularly group velocity dispersion (GVD, d^2 n / d\omega^2) and third-order dispersion (TOD), which broaden and asymmetrically reshape pulses in slow-light regimes. For example, in photonic crystal waveguides with v_g \approx 0.1c, GVD values of -10^6 ps²/km cause significant trailing-edge extension, while TOD up to $10^5 ps³/km exacerbates asymmetry, necessitating compensation techniques to preserve pulse integrity.

Recent advances and ongoing research

In the 2020s, significant progress has been made in integrating slow light into silicon photonic platforms, particularly for high-speed modulators. Researchers have demonstrated silicon-based slow-light electro-optic modulators achieving bandwidths exceeding 110 GHz, enabling compact devices with modulation lengths as short as 124 μm while maintaining low drive voltages around 1 V. These advancements leverage photonic crystal structures to enhance light-matter interactions without excessive propagation losses, supporting applications in data centers and telecommunications. A notable 2024 experiment utilized cavity enhancement to achieve substantial slowing of light in solid-state systems, demonstrating storage of light pulses in a chip-scale resonator with exceptional points induced by nonlinear Brillouin scattering, resulting in delays on the order of nanoseconds over micrometer scales. This approach highlights the potential for room-temperature operation in integrated transparent media, addressing previous limitations in dispersion engineering. In quantum applications, slow light techniques have advanced quantum information processing, including a 2025 demonstration of stored light using electromagnetically induced transparency in a superconducting qubit-resonator system, where light pulses were slowed and stored with group velocities reduced by factors exceeding 100 in a Λ-type artificial atom configuration. This work paves the way for hybrid quantum circuits. Emerging developments include hybrid systems combining electromagnetically induced transparency (EIT) with photonic chips, such as cavity-coupled two-level systems that enhance slow light via gain mechanisms in photonic molecules, achieving tunable transparency windows with minimal absorption. Furthermore, AI-optimized metamaterials have enabled broadband slow light, with deep reinforcement learning designs for photonic crystal waveguides yielding figure-of-merit products S \times \Delta \omega > 10^4 (where S is the slowdown factor and \Delta \omega the bandwidth), supporting applications in wideband signal processing. These structures hold potential for neuromorphic computing, where slow light enhances nonlinearities in optical neural networks, enabling ultrafast processing with latencies under 1 ps and energy efficiencies surpassing electronic counterparts in pattern recognition tasks. Looking ahead, scalable quantum memories leveraging slow light in rare-earth-doped solids promise integration into photonic networks for fault-tolerant computing, while their role in 6G optical interconnects could enable terabit-per-second links with reduced latency through dispersion-managed fibers. Ongoing research focuses on challenges like loss reduction to below 1 dB/ns delay, via advanced nanofabrication and hybrid material integrations to minimize scattering and absorption in high-index platforms.

References

  1. [1]
    [PDF] “Slow” and “fast” light
    A negative group velocity corresponds to the case when the peak of the pulse transmitted through an optical material emerges before the peak of the incident.
  2. [2]
    [PDF] Fundamentals and Applications of Slow Light by Zhimin Shi
    delay of a data packet through a slow-light medium is to define it in terms of the best- decision-time (BDT) topt in the eye-diagram when highest data ...
  3. [3]
  4. [4]
    Researchers now able to stop, restart light - Harvard Gazette
    Jan 24, 2001 · Albert Einstein theorized that light cannot travel faster than 186,282 miles per second. No one has proved him wrong, but he never said that it ...Missing: history | Show results with:history
  5. [5]
    [PDF] Slow Light in Optical Waveguides - Duke Physics
    In particular, we describe how slow light can be achieved by stimulated Brillouin and stimulated Raman scattering (SBS and SRS, respectively) in transparent ...
  6. [6]
    Slow light - Latest research and news | Nature
    Slow light is an electromagnetic wave with a group velocity reduced by optical resonances in the medium through which it is travelling.
  7. [7]
  8. [8]
    On the Group Front and Group Velocity in a Dispersive Medium ...
    Hamilton formulated the concept of group velocity as early as 1839 2, which was further developed by Stokes 3, Rayleigh 4, and Havelock 5.
  9. [9]
    Slow light in various media: a tutorial - Optica Publishing Group
    I consider the physical basics of slow light propagation in atomic media, photonic structures, and optical fibers. I show similarities and differences ...
  10. [10]
    refractive index, Sellmeier equation, dispersion formula - RP Photonics
    Sellmeier equations are very useful, as they make it possible to describe fairly accurately the refractive index in a wide wavelength range.Missing: William | Show results with:William
  11. [11]
    100 years of Brillouin scattering: Historical and future perspectives
    Nov 10, 2022 · We have reviewed the progress in science and applications of Brillouin light scattering over the past 100 years to date. Here, we look at ...
  12. [12]
    Propagation of a Gaussian Light Pulse through an Anomalous ...
    Feb 1, 1970 · The propagation of a Gaussian light pulse through a medium having a positive or negative absorption line is examined.
  13. [13]
    Observation of electromagnetically induced transparency
    May 20, 1991 · We report the first demonstration of a technique by which an optically thick medium may be rendered transparent.
  14. [14]
    Light speed reduction to 17 metres per second in an ultracold atomic ...
    Feb 18, 1999 · Here we report an experimental demonstration of electromagnetically induced transparency in an ultracold gas of sodium atoms.Missing: pulse | Show results with:pulse
  15. [15]
    Slow light in photonic crystal waveguides - IOPscience
    The physical principles behind the phenomenon of slow light in photonic crystal waveguides, as well as their practical limitations, are discussed and put into ...
  16. [16]
    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 ...
  17. [17]
    Producing slow light in warm alkali vapor using electromagnetically ...
    We present undergraduate-friendly instructions on how to produce light pulses propagating through warm Rubidium vapor with speeds less than 400 m/s.
  18. [18]
    A practical guide to electromagnetically induced transparency in ...
    Mar 7, 2023 · This tutorial introduces the theoretical and experimental basics of electromagnetically induced transparency (EIT) in thermal alkali vapors.
  19. [19]
    Electromagnetically induced transparency in rubidium - AIP Publishing
    Feb 1, 2009 · We have described an EIT experiment using rubidium atoms that is suitable for the advanced undergraduate laboratory. The theoretical absorption ...
  20. [20]
    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
  21. [21]
  22. [22]
    Observation of small group velocity in two-dimensional AlGaAs ...
    Mar 12, 2002 · The characteristics of the small group velocity 𝜈 g were experimentally studied with special concern to a flat band in AlGaAs-based ...
  23. [23]
    Coupled-resonator optical waveguide:?a proposal and analysis
    Jun 1, 1999 · In this Letter we propose a new type of waveguide based on coupling of optical resonators, the coupled-resonator optical waveguide (CROW).
  24. [24]
  25. [25]
    235-ring Coupled-Resonator Optical Waveguides
    Silicon microring coupled-resonator optical waveguides (CROWs) with up to 235 unit cells are fabricated and measured. These are the longest CROW structures made ...Missing: telecom | Show results with:telecom
  26. [26]
  27. [27]
    [PDF] Slow light in silicon microring resonators
    Double channel and single channel side- coupled integrated spaced sequence of resonators. (SCISSOR) devices were fabricated with electron beam lithography and ...
  28. [28]
    Active control of slow light on a chip with photonic crystal waveguides - Nature
    ### Summary of Key Experimental Results on Slow Light Delay in Silicon Photonic Crystal Waveguide
  29. [29]
    Tunable All-Optical Delays via Brillouin Slow Light in an Optical Fiber
    In this Letter, we demonstrate optically controllable slow-light delays at telecommunication wavelengths via stimulated Brillouin scattering in a conventional ...Missing: Mandel' shtam- precursor
  30. [30]
    Slow light, induced dispersion, enhanced nonlinearity, and optical ...
    Mar 6, 2002 · Such a structure leads to exotic optical characteristics, including ultraslow group velocities of propagation, enhanced optical nonlinearities, ...
  31. [31]
    Slow Light Enhanced Nonlinear Optics in Silicon Photonic Crystal ...
    Sep 12, 2025 · ... 2002, he was a Manager and. Senior Scientist with JDS Uniphase ... ics. Thomas F. Krauss received the Dipl.-Ing. degree. in photographic ...
  32. [32]
  33. [33]
    Refractive index gas sensor based on the Tamm state in a ... - Nature
    Jun 16, 2020 · The optimised structure of the proposed sensor achieves ultra-high sensitivity (S = 1.9×105 nm/RIU) and a low detection limit (DL = 1.4×10−7 RIU) ...Sensor Design · Effect Of Ag Layer Thickness · Sensor Analysis Under...
  34. [34]
    Enhanced stimulated Raman scattering in slow-light photonic crystal waveguides
    ### Summary of Enhanced Stimulated Raman Scattering in Slow-Light Silicon Photonic Crystal Waveguides
  35. [35]
    Observation of four-wave mixing in slow-light silicon photonic crystal ...
    In the specific case of slow-light enhancement of FWM in photonic crystal waveguides, the conversion efficiency has been theoretically suggested to be enhanced ...
  36. [36]
    [PDF] Fast Light, Slow Light and Optical Precursors: What Does It All Mean?
    Jan 5, 2007 · One aspect of Sommerfeld and Brillouin's result that can lead to confusion is the possible situation wherein one or more of these wave packets ...
  37. [37]
    [PDF] Limits on the Time Delay Induced by Slow-Light Propagation
    material contribution to the group delay Tdel = Tg − L/c, which is the difference between the group delay and the delay experienced upon propagation ...
  38. [38]
    [PDF] Slow- and fast-light: fundamental limitations - University of Rochester
    Dec 7, 2007 · This paper is devoted to the examination of some fundamental aspects of slow- and fast-light propagation. In particular, we make a careful ...Missing: review | Show results with:review
  39. [39]
    Delay-bandwidth product of electromagnetically induced ...
    May 3, 2007 · The imaginary part of the dispersion is connected to the real part through the Kramers-Kronig relation; it describes the absorption-gain ...
  40. [40]
    The effect of higher-order dispersion on slow light propagation in photonic crystal waveguides
    ### Summary of Pulse Distortion Due to Higher-Order Dispersion in Slow Light
  41. [41]
    Slow-light silicon modulator with 110-GHz bandwidth - Science
    Oct 20, 2023 · Silicon modulators are key components to support the dense integration of electro-optic functional elements for various applications.
  42. [42]
    Slow light silicon modulator beyond 110 GHz bandwidth - arXiv
    Feb 7, 2023 · We theoretically propose and experimentally demonstrate a design strategy for silicon modulators by employing the slow light effect.Missing: 2020s | Show results with:2020s
  43. [43]
    Storing light near an exceptional point | Nature Communications
    Sep 16, 2024 · Here we experimentally observe room-temperature storing light near an exceptional point induced by nonlinear Brillouin scattering in a chip-scale 90-μm-radius ...Results · Observation Of Storing Light... · Light Storage
  44. [44]
    Slow and stored light via electromagnetically induced transparency ...
    Jan 21, 2025 · In this Letter we present the experimental realization of EIT slow and stored light phenomena in superconducting waveguide quantum ...
  45. [45]
    Cavity gain enhanced slow light in a hybrid photonic molecule system
    Sep 29, 2025 · We propose a hybrid photonic molecule system, which includes a two-level system coupled to two optical cavities and the two cavities ...
  46. [46]
    Exploring high-performance photonic crystal slow light waveguides ...
    Oct 15, 2024 · Here, deep reinforcement learning is proposed for the first time for the design of complex slow-light photonic crystal waveguides.
  47. [47]
    Ultrafast neuromorphic computing with nanophotonic optical ... - arXiv
    Jan 28, 2025 · On-chip photonic neural networks offer a promising platform that leverage high bandwidths and low propagation losses associated with optical ...
  48. [48]
    [2508.06675] A scalable photonic quantum interconnect platform
    Aug 8, 2025 · Many quantum networking applications require efficient photonic interfaces to quantum memories which can be produced at scale and with high ...Missing: future slow light 6G reduction
  49. [49]
    Quantum for 6G communication: A perspective - Wiley Online Library
    May 31, 2023 · The paper presents how quantum technologies will enable future 6G and what are the future research directions in this area.Missing: prospects | Show results with:prospects
  50. [50]
    Lighting the way forward: The bright future of photonic integrated ...
    This review explores recent advancements in optical platforms, highlighting ongoing research and developments that are driving technological progress in this ...