Fact-checked by Grok 2 weeks ago

Light-emitting diode physics

A (LED) is a that emits via , a process in which forward-biased current across a p-n junction causes electrons from the n-type region to recombine with holes in the p-type region, releasing photons with equal to the material's . The fundamental physics relies on direct semiconductors, such as (GaAs) or (GaN), where the recombination occurs efficiently without significant involvement, enabling the conversion of to optical with high efficiency. In operation, applying a forward voltage reduces the depletion region's potential barrier, allowing carrier injection and diffusion into the , where dominates light production. The emitted light's wavelength, and thus color, is precisely tuned by the band gap energy—ranging from (e.g., ~1.4 eV for GaAs at 900 nm) to (e.g., <400 nm for certain InGaN alloys)—with common materials like AlGaInP for red-orange and InGaN for blue-green emissions. This band gap engineering allows LEDs to span the visible spectrum, from violet (~400 nm) to red (~760 nm). Efficiency in LEDs is characterized by internal quantum efficiency, which measures the fraction of injected carriers that result in photons via radiative recombination, often approaching 0.75–0.95 in optimized , though external efficiency is lower (~20–80% as of 2025) due to total internal reflection at the semiconductor-air interface, with modern extraction techniques significantly mitigating these losses. Advances in heterostructures and quantum wells enhance carrier confinement and reduce non-radiative losses, enabling luminous efficacies that surpass traditional sources like incandescent bulbs. White light, essential for general illumination, is achieved not through direct emission but via phosphor conversion of blue LEDs or color mixing of red, green, and blue variants.

Basic Principles

p-n Junction Operation

A p-n junction is formed by joining a p-type semiconductor, doped with acceptor impurities that create holes as majority carriers, to an n-type semiconductor, doped with donor impurities that provide electrons as majority carriers. Upon contact, electrons from the n-region diffuse into the p-region, and holes from the p-region diffuse into the n-region, leading to recombination near the interface and the exposure of fixed ionized donors (positive charge) on the n-side and ionized acceptors (negative charge) on the p-side. This charge separation establishes an electric field that opposes further diffusion, resulting in a depletion region—a carrier-depleted zone around the junction where mobile charges are minimized. The built-in potential V_{bi} across the junction in equilibrium arises from this charge separation and is given by V_{bi} = \frac{kT}{q} \ln \left( \frac{N_A N_D}{n_i^2} \right), where k is Boltzmann's constant, T is temperature, q is the elementary charge, N_A and N_D are the acceptor and donor doping concentrations, and n_i is the intrinsic carrier concentration. Energy band diagrams illustrate this: in equilibrium, the Fermi levels align across the junction, bending the conduction and valence bands to form a potential barrier of height qV_{bi}, with the depletion region showing a linear band slope due to the electric field. Under reverse bias, applying a negative voltage to the p-side (relative to the n-side) widens the depletion region and increases the barrier height, enhancing the electric field and suppressing carrier diffusion. In forward bias, a positive voltage to the p-side reduces the barrier, narrowing the depletion region and allowing majority carriers to overcome it. Under forward bias exceeding V_{bi}, minority carriers are injected across the junction: electrons from the n-region into the p-region and holes from the p-region into the n-region. These injected minorities then diffuse into the respective quasi-neutral regions, where they recombine with majority carriers over a characteristic diffusion length, typically on the order of micrometers, governed by the minority carrier lifetime and diffusivity. The resulting current-voltage characteristics follow the Shockley diode equation: I = I_s \left( e^{qV / kT} - 1 \right), where I_s is the reverse saturation current, dependent on material parameters like doping and lifetime, describing the exponential increase in forward current due to enhanced carrier injection and diffusion. In p-n junctions, carrier dynamics differ between direct and indirect bandgap semiconductors: in direct bandgap materials, momentum conservation allows efficient radiative recombination of injected carriers without phonon involvement, while in indirect bandgap materials, phonon-assisted processes are required for momentum matching, leading to lower recombination probabilities and predominantly non-radiative pathways. This distinction influences the overall efficiency of carrier transport and recombination in the junction, with direct bandgap semiconductors exhibiting faster radiative dynamics essential for light emission.

Electroluminescence Mechanism

Electroluminescence in light-emitting diodes (LEDs) arises from the recombination of injected electrons and holes in a semiconductor p-n junction under forward bias, converting electrical energy into photons. This process was first observed in 1907 by , who noted a yellow glow from a silicon carbide (SiC) crystal when applying a voltage, marking the initial discovery of solid-state electroluminescence. The practical realization of visible electroluminescence came in 1962 with 's demonstration of a red-emitting GaAsP diode, which relied on direct bandgap radiative recombination to achieve visible light emission. The core mechanism of light emission is radiative recombination, where an electron from the conduction band recombines with a hole in the valence band, releasing a photon with energy approximately equal to the semiconductor's bandgap energy E_g. In direct bandgap materials like , the conduction band minimum and valence band maximum occur at the same wavevector k = 0 in the Brillouin zone, allowing momentum conservation during the transition without additional phonon assistance; thus, the process is efficient, with the emitted photon's energy given by E \approx E_g. In contrast, indirect bandgap semiconductors such as have band extrema at different k-values, requiring phonon emission or absorption to conserve momentum, which suppresses radiative efficiency and favors non-radiative paths, making them unsuitable for efficient LEDs. Non-radiative recombination competes with the radiative process, reducing light output by dissipating energy as heat. Shockley-Read-Hall (SRH) recombination occurs via defect or impurity states within the bandgap, trapping carriers and enabling thermal re-emission or non-radiative annihilation, with rates dependent on defect density and capture cross-sections. Auger recombination involves three carriers: an electron-hole pair recombines radiatively, but the energy excites a third carrier to a higher state, which then relaxes non-radiatively via phonons; this process becomes dominant at high carrier densities in narrow-bandgap materials. Surface recombination further contributes at interfaces, where dangling bonds act as efficient traps, accelerating non-radiative decay and necessitating passivation techniques. The internal quantum efficiency (IQE) quantifies the fraction of injected carriers that result in radiative recombination, defined as \text{IQE} = \frac{R_{\text{rad}}}{R_{\text{rad}} + R_{\text{SRH}} + R_{\text{Auger}} + R_{\text{surface}}} where R_{\text{rad}}, R_{\text{SRH}}, R_{\text{Auger}}, and R_{\text{surface}} are the respective recombination rates; high IQE (>80%) is achievable in direct bandgap III-V semiconductors with low defect densities. Recombination processes and emission properties exhibit temperature dependence. Radiative rates in direct bandgap materials show weak temperature dependence, while non-radiative SRH rates may decrease if traps become ionized; rates, however, rise strongly with temperature owing to higher carrier energies. The emission \lambda shifts redward with increasing temperature as the bandgap narrows, following \lambda = \frac{hc}{E_g(T)}, where E_g(T) decreases approximately linearly for many semiconductors, leading to a typical shift of 0.1–0.3 nm/K in GaN-based LEDs.

Device Structure and Materials

Semiconductor Materials

Light-emitting diodes (LEDs) primarily utilize III-V compound semiconductors due to their direct or tunable bandgaps, which facilitate efficient radiative recombination for light emission. Gallium arsenide (GaAs) is a direct-bandgap material with an energy gap of 1.42 eV, enabling infrared emission around 870 nm and serving as a foundational material for early LED devices. Gallium phosphide (GaP), an indirect-bandgap semiconductor with a bandgap of 2.26 eV, has been employed for green light emission at approximately 550 nm, though its indirect nature results in lower efficiency compared to direct-bandgap alternatives. Indium gallium nitride (InGaN), a direct-bandgap alloy, offers tunable bandgap energies from approximately 0.7 eV (InN) to 3.4 eV (GaN) by varying the indium composition, enabling emission from near-infrared to ultraviolet wavelengths. High-efficiency visible LEDs are primarily achieved in the blue to ultraviolet range, with recent advances (as of 2024) enabling high-efficiency green and red emissions for full-color displays. Alloying III-V compounds, such as aluminum gallium arsenide (AlGaAs) and indium gallium phosphide (InGaP), is crucial for bandgap engineering while maintaining lattice matching to substrates like GaAs to minimize defects and dislocations that degrade performance. AlGaAs exhibits near-perfect lattice matching to GaAs (mismatch <0.2%), promoting high-quality epitaxial growth with reduced non-radiative recombination centers. In InGaP alloys, compressive or tensile strain is intentionally introduced to adjust the bandgap for red to yellow emission, but careful control is required to avoid phase separation and strain-induced defects that increase with indium content. Achieving p-type doping in wide-bandgap III-V materials like gallium nitride (GaN) presents significant challenges, primarily due to the high activation energy of magnesium (Mg) acceptors, which ranges from 200 to 265 meV, limiting hole concentration and conductivity at room temperature. This requires high-temperature annealing to activate Mg and remove hydrogen passivation, yet solubility limits and compensation by defects often result in suboptimal p-type layers essential for p-n junctions. The development of LED semiconductors evolved from infrared GaAs devices in the early 1960s to visible-light emitters, culminating in efficient blue InGaN LEDs that enabled white light generation through phosphor conversion. This breakthrough, recognized by the 2014 Nobel Prize in Physics awarded to Isamu Akasaki, Hiroshi Amano, and Shuji Nakamura, overcame longstanding challenges in growing high-quality GaN-based materials. Despite their advantages, LED semiconductors face limitations including toxicity in cadmium-based II-VI compounds like CdSe used historically for green and yellow emission, high production costs for GaN due to complex epitaxial growth on lattice-mismatched substrates, and thermal stability issues where elevated temperatures can accelerate defect formation and reduce efficiency in wide-bandgap materials.

Heterostructures and Quantum Confinement

Heterostructures in light-emitting diodes (LEDs) enhance performance by providing effective confinement of charge carriers and photons within the active region, enabling higher efficiency and brighter emission compared to homojunction devices. The foundational concept of the double heterostructure (DH), proposed in 1963, involves sandwiching a narrower-bandgap active layer between wider-bandgap p-type and n-type cladding layers, which creates discontinuities in the conduction and valence band edges, denoted as ΔE_c and ΔE_v, respectively. These band offsets confine electrons and holes to the active region, reducing non-radiative recombination losses outside it. A classic example is the GaAs active layer bounded by AlGaAs cladding layers, where the aluminum content in AlGaAs creates sufficient band offsets (typically ΔE_c ≈ 0.3–0.6 eV depending on composition) to achieve efficient carrier injection and radiative recombination at wavelengths around 870 nm. Experimental comparisons of such DH LEDs with single heterostructure counterparts demonstrated internal quantum efficiencies exceeding 20% and reduced operating voltages, highlighting the role of symmetric carrier confinement in minimizing current spreading. Building on DH designs, quantum wells (QWs) introduce two-dimensional (2D) carrier confinement by using ultrathin active layers (typically 2–10 nm thick) sandwiched between thicker barrier layers, which quantize the energy states and modify the density of states. In InGaN/GaN multiple quantum well (MQW) structures, for instance, alternating InGaN wells and GaN barriers enable blue emission at ~450 nm with enhanced wavefunction overlap between electrons and holes. This confinement increases the oscillator strength and radiative recombination rate, where the inverse lifetime 1/τ_rad scales with the square of the electron-hole wavefunction overlap |ψ|^2, leading to faster emission and higher efficiency. High-brightness blue QW LEDs achieved external quantum efficiencies up to 9% at 20 mA drive current, a significant improvement over bulk active layers. These structures offer key advantages, including reduced threshold currents for lasing applications (though LEDs benefit from lower drive currents for equivalent output) and elevated internal quantum efficiency (IQE) due to suppressed carrier leakage. However, in wurtzite III-nitride materials like InGaN/GaN grown along the c-axis, spontaneous and piezoelectric polarization fields induce an internal electric field (~1–3 MV/cm), causing the (QCSE). This red-shifts the emission wavelength and reduces |ψ|^2 by separating electron and hole wavefunctions, thereby decreasing IQE at longer wavelengths such as green. To address QCSE, non-polar and semi-polar GaN orientations have been developed, reducing internal fields and enabling higher efficiency in green and yellow LEDs, as demonstrated in recent epitaxial structures (as of 2023). Epitaxial growth of these heterostructures and QWs relies primarily on metal-organic chemical vapor deposition (MOCVD), which allows precise control of layer thickness and composition at atmospheric or low pressure, essential for achieving sharp interfaces and low defect densities in materials like and . Further dimensionality reduction to quantum wires (1D confinement) and quantum dots (0D confinement) modifies the density of states into step-like or delta functions, potentially enhancing gain and efficiency, though these are explored in detail in advanced nanocrystal-based LEDs.

Optical Physics

Refractive Index Mismatch

In light-emitting diodes (LEDs), the significant difference between the refractive index of the semiconductor material and that of the surrounding medium, typically air, leads to substantial optical losses through total internal reflection (TIR). Semiconductor materials commonly used in LEDs, such as gallium arsenide (GaAs) with a refractive index n \approx 3.5 or gallium nitride (GaN) with n \approx 2.4, contrast sharply with air's n = 1. When photons generated by recombination in the active region propagate toward the chip's surface, those incident at angles exceeding the critical angle \theta_c = \sin^{-1}(n_{\text{out}}/n_{\text{in}}) undergo TIR and remain confined within the device. For GaN-based LEDs, this critical angle is approximately 24°, meaning approximately 91% of possible emission directions result in reflection back into the semiconductor (with ~9% within the escape cone). The escape cone defined by this critical angle severely restricts the external quantum efficiency by limiting the solid angle through which light can exit the LED chip. For isotropic emission from the active region, the fraction of light escaping into air without enhancement is given by the geometric factor $1/(2n^2), representing the portion of the hemispherical solid angle within the escape cone. In high-index materials like GaAs (n \approx 3.5), this yields only about 4% extraction efficiency in the simple model, while for GaN (n \approx 2.4), it is around 9%; however, accounting for chip geometry and multiple interfaces often reduces the effective value to approximately 2% for unoptimized planar devices. This confinement arises because recombination in the active region produces photons uniformly in all directions, but only those nearly perpendicular to the surface avoid TIR. Under the ray optics model, emitted photons are traced as rays that bounce multiple times within the LED chip due to TIR at the semiconductor-air interfaces, eventually leading to reabsorption or escape from sidewalls. These guided rays travel laterally through the high-index layers, undergoing successive reflections that increase the path length and probability of parasitic absorption by free carriers, defects, or metallic contacts. In typical LED chips, this trapping dominates optical losses, as the rays' trajectories form closed paths or waveguiding modes parallel to the surface, preventing direct extraction. The refractive index mismatch is further influenced by wavelength-dependent dispersion, where n varies with photon energy according to relations like the Sellmeier equation, n^2(\lambda) = 1 + \sum_i B_i \lambda^2 / (\lambda^2 - C_i), capturing resonant contributions from electronic transitions. In III-V semiconductors, shorter wavelengths (e.g., blue light in ) correspond to higher n near the band edge, resulting in a smaller \theta_c and more pronounced TIR compared to longer red wavelengths. This dispersion exacerbates light trapping in ultraviolet or blue LEDs, where extraction fractions drop below those in red devices. The mismatch profoundly impacts emission patterns in LED chips, favoring confinement for isotropic sources while permitting higher escape for directed emission aligned normal to the interfaces. In practice, the active region's recombination yields largely isotropic radiation, promoting lateral waveguiding and reducing perpendicular output, which contrasts with ideal directed emitters that could exploit the full escape cone more effectively.

Light Extraction Techniques

One primary method to enhance light extraction in LEDs involves encapsulation with materials that reduce the refractive index contrast at the semiconductor-air interface. By surrounding the LED chip with an epoxy resin having a refractive index of approximately 1.5, the critical angle for total internal reflection increases, enlarging the escape cone and allowing more light to exit the device. This approach typically improves extraction efficiency by 1.5 to 2 times compared to air-encapsulated devices. Further enhancement is achieved through hemispherical dome shapes in the epoxy, which redirect rays that would otherwise undergo multiple internal reflections, effectively shaping the emission pattern and boosting output power. Surface texturing techniques scatter light to redirect photons beyond the critical angle, significantly increasing extraction. Microroughening the p- surface, for instance, has been shown to double the light output power in InGaN-based LEDs by promoting diffuse reflection and reducing Fresnel losses. Photonic crystals, formed by periodic nanopatterns on the surface, further amplify this effect through diffraction and bandgap engineering, achieving 2-3 times higher extraction in LEDs by coupling guided modes to free space. Transition coatings employ thin multilayer films to create a graded refractive index profile, minimizing reflections at interfaces. Structures such as SiO₂/TiO₂ multilayers approximate an antireflection coating, where the reflection coefficient R = \left| \frac{n_1 - n_2}{n_1 + n_2} \right|^2 is reduced across layers, enabling up to 30% improvement in light output for GaN devices. These coatings provide a smooth index transition from the high-index semiconductor (n ≈ 2.5) to air (n = 1), suppressing waveguiding losses. Chip shaping modifies the device geometry to optimize ray paths and minimize trapping. Hemispherical or truncated pyramid designs in GaN LEDs redirect light toward the surface, increasing extraction by 20-50% depending on the aspect ratio, as steeper angles reduce the probability of total internal reflection for oblique rays. Post-2010 advances have integrated nanorod arrays and advanced photonic crystals to achieve extraction efficiencies exceeding 50% in GaN LEDs. Nanorod structures, such as ZnO or GaN arrays, scatter light laterally and vertically, enhancing output by over 100% in some green emitters through improved photon escape from sidewalls. Similarly, optimized photonic crystal patterns on ITO or p-GaN layers have demonstrated light extraction up to 70% by precisely controlling mode coupling, representing a key step toward high-brightness applications. As of 2025, experimental demonstrations in advanced GaN microLEDs with nanorod arrays and photonic crystals have achieved light extraction efficiencies exceeding 60%, with simulations suggesting potential up to 80% under optimized conditions.

Performance Metrics

Efficiency Parameters

The efficiency of light-emitting diodes (LEDs) is quantified through several key parameters that assess the conversion of electrical energy into light, encompassing internal processes within the device and external factors affecting output. These metrics are essential for evaluating overall performance, guiding material and structural optimizations, and comparing devices across applications. Internal quantum efficiency (IQE) measures the fraction of injected electrons (or holes) that recombine radiatively to generate photons within the active region of the LED, relative to the total number of injected carriers. It is defined as the ratio of the internal optical power emitted from the active region to the electrical power associated with carrier injection, expressed as IQE = \frac{P_\mathrm{int}}{I \cdot (h\nu / q)}, where P_int is the internal optical power, I is the injection current, h\nu is the photon energy, and q is the elementary charge. This parameter fundamentally depends on the balance between radiative and non-radiative recombination processes in the semiconductor. External quantum efficiency (EQE) extends this by accounting for the fraction of internally generated photons that successfully escape the device as useful output light, incorporating carrier injection efficiency (the proportion of injected carriers reaching the active region) and light extraction efficiency (the portion of photons that avoid total internal reflection or absorption). It is given by EQE = IQE × injection efficiency × extraction efficiency, representing the number of emitted photons escaping the LED per injected electron. High EQE requires optimizing all these factors, with light extraction playing a critical role due to refractive index mismatches at interfaces. Wall-plug efficiency (WPE), also known as power efficiency, evaluates the overall energy conversion from electrical input to optical output, defined as WPE = P_opt / (I V) × 100%, where P_opt is the optical power output and V is the forward voltage. This metric relates to EQE through WPE = EQE × (hν / q V), where hν is the photon energy, highlighting the impact of operating voltage on practical efficiency. WPE is particularly important for high-power applications, as it directly indicates electrical-to-optical power utilization. For visible-spectrum LEDs, luminous efficacy provides a human-centric measure of efficiency, quantifying visible light output in lumens per watt (lm/W) while accounting for the eye's spectral sensitivity via the photopic luminosity function V(λ), which peaks at 555 nm with a maximum efficacy of 683 lm/W for monochromatic green light. It is calculated as luminous efficacy = (optical power × 683 × ∫ V(λ) S(λ) dλ / ∫ S(λ) dλ) / electrical power, where S(λ) is the LED's spectral power distribution; typical values for white LEDs range from 100 to 200 lm/W in modern devices. This parameter bridges radiometric efficiencies like EQE with perceptual brightness. Efficiency parameters exhibit strong dependence on operating conditions, with IQE typically peaking at low injection currents (around 1-10 mA/cm²) where non-radiative losses are minimized, and decreasing at higher currents due to increased carrier densities. Temperature also influences performance, as rising junction temperatures enhance non-radiative recombination rates, reducing IQE by 5-20% from 25°C to 100°C in depending on device quality, while extraction efficiency may vary due to thermal changes in refractive indices. Measurement of these efficiencies has evolved with standardized techniques, such as temperature-dependent electroluminescence for IQE assessment and integrating sphere photometry for EQE, enabling precise characterization. Advancements since the early 2000s have dramatically improved blue LED performance, with EQE rising from below 5% in 2000 to over 80% by the 2020s through refinements in epitaxial growth and device design.

Efficiency Droop

Efficiency droop is a critical limitation in light-emitting diodes (LEDs), characterized by a decrease in external quantum efficiency (EQE) as the injection current density increases beyond a certain threshold. In typical blue LEDs, the EQE peaks at low current densities of approximately 20–100 mA/cm² and can decline by more than 50% at 1 A/cm², hindering the scaling of light output for high-power operation. This phenomenon primarily affects the internal quantum efficiency (IQE), which is governed by the competition between radiative and non-radiative recombination processes under high carrier injection. The main mechanisms contributing to efficiency droop include Auger recombination, a non-radiative process where the energy from electron-hole recombination excites another carrier, leading to losses proportional to the cube of the carrier density (∝ n³, where n is the carrier concentration). Direct measurement of hot electrons emitted from operating LEDs has confirmed Auger's dominance, with coefficients C on the order of 10⁻³⁰ cm⁶/s. Additionally, carrier delocalization in InGaN quantum wells (QWs) at high densities reduces localization in potential fluctuations, promoting spillover and non-radiative pathways, while thermal heating causes a rise in junction temperature that exacerbates recombination losses. These effects are particularly pronounced in c-plane InGaN/GaN structures due to the quantum-confined Stark effect (QCSE), which separates electron and hole wavefunctions. The ABC model provides a framework for quantifying these recombination dynamics, expressing the total recombination rate R as: R = A n + B n^2 + C n^3 where A represents Shockley-Read-Hall (SRH) non-radiative recombination (defect-related, linear in carrier density), B is the radiative bimolecular coefficient, and C captures the Auger term; at high currents, the C n³ term dominates, driving the IQE downturn. Experimental fits to this model across various LED designs reveal A values around 10⁸ s⁻¹ and B near 10⁻¹¹ cm³/s, underscoring Auger's role in droop. Mitigation strategies focus on suppressing these mechanisms, such as employing thinner QWs to enhance carrier confinement and wavefunction overlap against , alongside polarization engineering using nonpolar or semipolar orientations to eliminate piezoelectric fields. Recent advancements in the 2020s, including ultra-low-defect-density homoepitaxial substrates with threading dislocation densities below 10⁵ cm⁻², have reduced droop to less than 20% at 500 A/cm² in blue micro-LEDs, primarily by minimizing losses. These improvements enable higher flux in lighting applications, where droop otherwise limits output compared to low-current display uses that operate near peak efficiency.

Reliability and Advanced Topics

Lifetime and Failure Modes

The lifetime of light-emitting diodes (LEDs) is typically quantified using the L70 metric, which represents the operational time in hours until the luminous flux depreciates to 70% of its initial value under specified conditions. This metric accounts for gradual degradation rather than abrupt failure, as LEDs rarely cease functioning suddenly but instead exhibit diminishing output over extended periods. Lifetime predictions often employ the for extrapolation, incorporating an activation energy E_a (typically 0.1–0.5 eV depending on the LED type) to model thermal acceleration of degradation processes from accelerated test data. Key failure modes in LEDs, particularly those based on GaN, involve structural and chemical changes that increase non-radiative recombination. Dislocation climb in GaN layers, driven by recombination-enhanced mechanisms, generates additional defects that serve as non-radiative centers, reducing radiative efficiency over time. In p-type GaN doped with magnesium (Mg), diffusion of Mg atoms into the active region during operation or high-temperature processing can introduce impurities that form deep-level traps, accelerating carrier leakage and flux decay. Contact degradation, often at n- or p-type interfaces, arises from electromigration or oxidation under high current densities, leading to increased series resistance and uneven current distribution. Thermal effects play a central role in degradation, as elevated junction temperatures exacerbate failure modes. The junction temperature \theta_j rises due to power dissipation, approximated by \theta_j = P_\text{elec} \times R_\text{th}, where P_\text{elec} is electrical power input and R_\text{th} is thermal resistance (typically 10–50 K/W for high-power LEDs). This heating shortens the Shockley-Read-Hall (SRH) recombination lifetime exponentially, promoting defect generation and non-radiative losses that contribute to both gradual degradation and transient efficiency droop under heat. Acceleration factors amplify these processes in real-world operation. Current crowding, where current density is non-uniform across the active area, creates localized hot spots that intensify thermal and defect-related degradation, particularly at high drive currents above 100 mA/mm². In packaged LEDs, -induced corrosion at metal interfaces or encapsulants, often involving sulfur or chloride contaminants, erodes electrical contacts and phosphor layers, hastening lumen depreciation in moist environments (e.g., >85% relative humidity). Standardized testing under IES LM-80 protocols measures maintenance over at least 6,000 hours at controlled currents (e.g., 350–700 ) and case temperatures (55°C, 85°C, or 105°C), enabling reliable extrapolation to longer . Post-2010 advancements in , thermal management, and packaging have elevated L70 for modern white LEDs beyond 50,000 hours at nominal operating conditions, reflecting reduced defect densities and improved heat dissipation.

Quantum-Dot LEDs

Quantum-dot light-emitting diodes (QD-LEDs), also known as QLEDs, incorporate colloidal nanocrystals known as quantum dots (QDs) as the emissive layer to achieve enhanced color purity and efficiency in electroluminescent devices. These QDs exhibit zero-dimensional (0D) quantum confinement, where electrons and holes are restricted in all three spatial dimensions within nanocrystals typically 2-10 nm in diameter, such as CdSe or InP. This confinement leads to discrete energy levels, enabling size-tunable bandgaps that span the ; the confinement energy shift follows the particle-in-a-box model, with \Delta E \propto 1/r^2, where r is the particle radius. In QD-LED structures, the QDs form the active emissive layer in organic-inorganic devices, sandwiched between charge transport layers—typically a hole transport layer (e.g., organic polymers like poly(9,9-dioctylfluorene)) and an electron transport layer (e.g., inorganic ZnO nanoparticles)—to facilitate balanced injection and recombination. The solution-processable nature of colloidal QDs allows for low-cost fabrication via spin-coating or , distinguishing these devices from epitaxial quantum well-based LEDs. Key advantages of QD-LEDs stem from the physics of confined excitons in QDs, including narrow spectra with (FWHM) below 30 nm, which provides superior color purity compared to LEDs. Internal quantum efficiencies (IQE) exceeding 90% are achievable due to suppressed non-radiative Auger recombination, as the 0D confinement reduces multi-exciton interactions that dominate in bulk materials or higher-dimensional structures. Stability is further enhanced by core-shell architectures, such as CdSe cores overcoated with ZnS shells, which passivate surface traps, boost quantum yields to 70–90%, and protect against environmental degradation. Despite these benefits, challenges persist from charging effects and , where intermittent fluorescence arises from recombination in charged QDs or of carriers at surface defects, leading to non-radiative decay and reduced device reliability. Toxicity concerns with cadmium-based QDs, regulated under hazardous substance directives, have driven development of lead-free alternatives like InP QDs, which maintain comparable performance while offering environmental safety for display applications. Progress in QD-LEDs traces back to the , with seminal colloidal synthesis of monodisperse CdSe QDs via hot-injection methods enabling controlled size distributions. The first electroluminescent QD-LED was demonstrated in using CdSe nanocrystals as the emissive layer. By the , advancements in shell engineering and charge balance have yielded commercial QLED displays with external quantum efficiencies (EQE) over 20%, with records reaching 31% for red devices as of November 2025, approaching the theoretical limits for lighting and next-generation televisions.

References

  1. [1]
    Physics of Light and Color - Introduction to Light Emitting Diodes
    Nov 13, 2015 · The basic structure of a light emitting diode consists of the semiconductor material (commonly referred to as a die), a lead frame on which the ...
  2. [2]
    Light Emitting Diodes - Engineering LibreTexts
    Sep 7, 2021 · LEDs are pn junction devices made from extrinsic semiconductors. An n-type and a p-type semiconductor are put in contact with each other to form a pn junction ...
  3. [3]
    Light Emitting Diodes - HyperPhysics
    In light emitting diodes (LEDs), light is produced by a solid state process called electroluminescence. Under specific conditions, solid state light sources can ...Missing: fundamentals | Show results with:fundamentals
  4. [4]
    [PDF] Chapter 7 Semiconductor Light Emitting Diodes and Solid State ...
    Semiconductor light emitting diodes are forward biased pn junction diodes in which electron-hole recombination due to spontaneous emission in the junction ...
  5. [5]
    LED Basics | Department of Energy
    LEDs emit nearly monochromatic light, not white. White light is achieved via phosphor conversion, color mixing, or a hybrid method. LEDs have high efficiency ...
  6. [6]
    The P-N Junction - HyperPhysics
    The solid circles on the right of the junction represent the available electrons from the n-type dopant. Near the junction, electrons diffuse across to ...
  7. [7]
    [PDF] Lecture 5 PN Junctions in Thermal Equilibrium - Cornell University
    Both the N-doped and P-doped materials were charge neutral before the junction was ... A PN Junction in Equilibrium: Depletion Region Widths. 0 x. P-doped. N ...
  8. [8]
    [PDF] PN and Metal–Semiconductor Junctions
    The depletion layer penetrates primarily into the lighter doping side, and the width of the depletion layer in the heavily doped material can often be neglected ...
  9. [9]
    [PDF] ECE606: Solid State Devices Lecture 14 Electrostatics of p-n junctions
    Oct 5, 2012 · Equilibrium: Many electrons N- side. Few electrons P-side. Detailed balance of drift and diffusion forces. → No net current. Forward Bias: ...
  10. [10]
    [PDF] Lecture 11: pn junctions under bias - An-Najah Staff
    Thus, typical diffusion lengths in a pn junction, where the injected minority carriers recombine, is of the order of µm.
  11. [11]
    [PDF] Lecture 14 - MIT
    Complete physical picture for pn diode under bias: • In forward bias, injected minority carriers diffuse through QNR and recombine at semiconductor surface.
  12. [12]
    [PDF] ( )x ( )
    Carrier Injection in a Forward Biased PN Junction Diode. In forward bias, the minority carrier concentrations increase exponentially at the edges of the ...
  13. [13]
    Diode Equation | PVEducation
    The diode equation gives an expression for the current through a diode as a function of voltage. The Ideal Diode Law, expressed as:Missing: I_s | Show results with:I_s
  14. [14]
    Direct and Indirect Band Gap Semiconductors - DoITPoMS
    In a direct band gap semiconductor, the top of the valence band and the bottom of the conduction band occur at the same value of momentum.
  15. [15]
    22.5: Direct and Indirect Band Gap Semiconductors
    Feb 4, 2021 · In direct band gap semiconductors, the valence and conduction bands occur at the same momentum. In indirect, they occur at different momentum ...
  16. [16]
    COHERENT (VISIBLE) LIGHT EMISSION FROM Ga(As1−xPx ...
    Nick Holonyak, Jr., S. F. Bevacqua; COHERENT (VISIBLE) LIGHT EMISSION FROM Ga(As1−xPx) JUNCTIONS, Applied Physics Letters, Volume 1, Issue 4, 1 December ...
  17. [17]
    Statistics of the Recombinations of Holes and Electrons | Phys. Rev.
    W. Shockley, Electrons and Holes in Semiconductors (D. van Nostrand Company, Inc., New York, 1950), p. 347 R. N. Hall, Phys. Rev. 83, 228 (1951) ibid.87, 387 ( ...
  18. [18]
    Auger effect in semiconductors | Proceedings of the Royal Society of ...
    This paper presents a calculation of the lifetimes (τ) of excess electrons and holes in a semiconductor aaanming the Auger effect between bands.
  19. [19]
    Simple analysis method for determining internal quantum efficiency ...
    May 30, 2013 · The EQE of an LED is defined as the product of the internal quantum efficiency (IQE) and the light extraction efficiency (LEE).
  20. [20]
    III-V Semiconductor - an overview | ScienceDirect Topics
    GaAs is a direct bandgap material with energy bandgap equal to 1.424 eV. They exhibit higher thermal conductivity, chemical resistance, mechanical strength, and ...
  21. [21]
    3.3.9: Light Emitting Diode - Engineering LibreTexts
    Jan 4, 2023 · How light-emitting diodes work. Discussion of direct and indirect band-gap semiconductors, and recombination centers.
  22. [22]
    [PDF] Refractive index of InGaN/GaN quantum well - HKU Scholars Hub
    GaN has a direct band gap of 3.4 eV, and the AlGaN and InGaN alloys demonstrate a band gap rang- ing from 1.95 to 6.2 eV, depending on the alloy composition.
  23. [23]
    Aluminium Gallium Arsenide - an overview | ScienceDirect Topics
    The lattice mismatch between these two materials is just 0.2% at normal room temperature. As a result, very high-quality films of AlGaAs materials can be grown ...
  24. [24]
    [PDF] Red to green emitters from InGaP/InAlGaP laser structure by strain ...
    Apr 28, 2016 · Reducing Al leads to the increases in wavelength, and the resultant alloys still maintain the lattice matching condition. As can be seen from ...
  25. [25]
    Progress and Challenges of p-type GaN Power Devices
    Sep 2, 2020 · Although magnesium (Mg) is the only shallow acceptor for GaN, it is difficult to achieve a high free hole concentration owing to two major ...
  26. [26]
    Study of the heavily p-type doping of cubic GaN with Mg - Nature
    Oct 8, 2020 · However, for Mg-doped h-GaN there are major issues, namely: (a) a large acceptor activation energy ≃ 200–265 meV, (b) Mg solubility limit around ...
  27. [27]
    [PDF] EFFICIENT BLUE LIGHT-EMITTING DIODES LEADING TO BRIGHT ...
    Oct 7, 2014 · The invention of efficient blue LEDs has led to white light sources for illumination. When exciting a phosphor material with a blue LED ...
  28. [28]
    Quantum dot LEDs go cadmium-free - Physics World
    Dec 4, 2019 · To date, the chief drawback of these quantum-dot LEDs has been their toxicity, since most contain cadmium or other heavy metals. Now ...
  29. [29]
    GaN on GaN LED breaks the "curse" of efficiency and cost - EEWorld
    Nov 17, 2013 · GaN on GaN LED breaks the "curse" of efficiency and cost · GaN on GaN efficiency is stable without being affected by lattice constant mismatch.Missing: limitations | Show results with:limitations<|separator|>
  30. [30]
    Defects and Reliability of GaN‐Based LEDs: Review and Perspectives
    Feb 11, 2022 · This article provides an overview of the main factors and mechanisms that limit the short- and long-term reliability of modern GaN-based ...
  31. [31]
    A proposed class of hetero-junction injection lasers - IEEE Xplore
    Published in: Proceedings of the IEEE ( Volume: 51 , Issue: 12 , December 1963 ). Article #:. Page(s): 1782 - 1783. Date of Publication: 31 December 1963.
  32. [32]
    Comparison of single heterostructure and double ... - NASA ADS
    This paper presents a theoretical and experimental comparison of the performance of single heterostructure (sh) and double heterostructure (dh) GaAs-GaAIAs ...Missing: AlGaAs 1970s
  33. [33]
    [PDF] THE BELL SYSTEM - World Radio History
    Double Heterostructure GaAs-GaAIAs LEDs ... 1582 THE BELL SYSTEM TECHNICAL JOURNAL, SEPTEMBER 1979 ... (AlGa) As Double-Heterostructure Junction Lasers," IEEE J.
  34. [34]
    High-Brightness InGaN Blue, Green and Yellow Light-Emitting ...
    Copyright (c) 1995 The Japan Society of Applied Physics Japanese Journal of Applied Physics, Volume 34, Number 7ACitation Shuji Nakamura et al 1995 Jpn. J. Appl ...
  35. [35]
    Quantum-Confined Stark Effect due to Piezoelectric Fields in GaInN ...
    We have studied the influence of piezoelectric fields on luminescence properties of GaInN strained quantum wells. Our calculation suggests that an electric ...
  36. [36]
    P-GaN/N-InGaN/N-GaN Double-Heterostructure Blue-Light-Emitting ...
    P-GaN/n-InGaN/n-GaN double-heterostructure (DH) blue-light-emitting diodes (LEDs) were fabricated successfully for the first time. The output power was 125 ...
  37. [37]
    Novel patterned sapphire substrates for enhancing the efficiency of ...
    Apr 24, 2020 · When light penetrates from GaN into the air, the refractive index of GaN (n = 2.4) causes a critical angle of only 23.6 degrees and make ...
  38. [38]
    Broadband measurements of the refractive indices of bulk gallium ...
    The observed dispersion fits well to a two-pole Sellmeier equation with an estimated overall accuracy of ± 0.002. Our results are found to be in good agreement ...
  39. [39]
    Direct band gap white light emission from charge carrier diffusion ...
    Dec 15, 2024 · As an example, GaN-based LEDs have a critical angle as small as 24.6°, resulting in a light extraction efficiency (LEE) from an unprocessed ...
  40. [40]
    Anisotropic nanocrystal superlattices overcoming intrinsic light ...
    Apr 19, 2022 · The classical ray optics theory gives a first approximation, ηout = 1/(2n2), where n is the refractive index of emissive layer. For ...<|control11|><|separator|>
  41. [41]
    Optically functional surface composed of patterned graded-refractive ...
    Sep 9, 2011 · Only a small percentage of light (≈1/4n2 per surface, where n is the high refractive index of the semiconductor) is extracted from the LED chip ...
  42. [42]
    In-depth insights into polarization-dependent light extraction ...
    Apr 25, 2023 · The LED chip model was exposed to air to simulate the on-wafer ... multiple reflections, as schematically illustrated by rays 5 and 6.
  43. [43]
    Fitting the refractive indices of GaN at different conditions ... - Nature
    Mar 4, 2025 · According to Sellmeier formula, the ordinary (\:{n}_{o}) and extraordinary (\:{n}_{e}) RIs of GaN can be simulated (λ in µm) as follows. ...
  44. [44]
  45. [45]
    Plug Efficiency - an overview | ScienceDirect Topics
    Plug efficiency is defined as the ratio of the watts of light produced by a p-n junction LED to the watts spent in injecting current into the device.
  46. [46]
    Direct measurement of internal quantum efficiency in light emitting ...
    Apr 15, 2011 · The WPE is related to the EQE by the voltage drop (V) at the device due to the diode forward ... The ηEQE can be written as η EQE = η IQE ...Missing: luminous efficacy
  47. [47]
    Enhancement of wall-plug efficiency in AlGaN deep-ultraviolet light ...
    Apr 8, 2025 · We report an enhancement of the wall-plug efficiency (WPE) of AlGaN-based deep-ultraviolet light-emitting diodes (DUV-LEDs).
  48. [48]
    3 Assessment of LED and OLED Technologies
    The concepts of internal quantum efficiency (IQE) and external quantum efficiency (EQE) are introduced in Annex 3.A, “An LED Primer,” and Figure 3.4 ...
  49. [49]
    Understanding and evaluating the mean photon energy and the ...
    Nov 8, 2022 · The power efficiency (PE), sometimes referred to as the wall-plug efficiency, is the efficiency of the utmost importance that characterises an ...Missing: metrics | Show results with:metrics
  50. [50]
    Luminous Efficacy - an overview | ScienceDirect Topics
    Luminous efficacy is defined as a measure of how bright radiation is perceived by the average human eye, expressed in lumens per watt (lm/W).
  51. [51]
    From physics to fixtures to food: current and potential LED efficacy
    Mar 30, 2020 · Blue LEDs can be 93% efficient, phosphor-converted “whites” 76% efficient, and red LEDs 81% efficient. These improvements open new opportunities ...
  52. [52]
    Internal quantum efficiency of high-brightness AlGaInP light-emitting ...
    Oct 18, 2005 · The evaluation is based on measurements of the external quantum efficiency of the LEDs as a function of the operating current and temperature.
  53. [53]
  54. [54]
    [PDF] Temperature-Dependent Internal Quantum Efficiency of Blue High ...
    Internal quantum efficiency (IQE) of a blue high- brightness InGaN/GaN LED was evaluated from the external quantum efficiency measured as a function ...
  55. [55]
    Efficiency studies for LEDs - NPL - National Physical Laboratory
    Temperature dependent electro luminescence (TDEL) is widely considered the most direct and standard method to study LED quantum efficiency, and is particularly ...<|control11|><|separator|>
  56. [56]
    Recent Progress in Long‐Wavelength InGaN Light‐Emitting Diodes ...
    May 26, 2023 · To date, state-of-the-art InGaN blue LEDs achieved a considerably high external quantum efficiency (EQE) beyond 80%, and when used as excitation ...<|control11|><|separator|>
  57. [57]
  58. [58]
  59. [59]
    Direct Measurement of Auger Electrons Emitted from a ...
    Apr 25, 2013 · In GaN light-emitting diodes (LEDs), Auger recombination is also invoked as a possible origin of the so-called efficiency droop phenomenon ...
  60. [60]
  61. [61]
    [PDF] LED Luminaire Lifetime: Recommendations for Testing and Reporting
    In earlier editions of this guide, “standard” or “default” lifetime of an LED luminaire (or lamp) was defined only in terms of lumen output and specified as the ...
  62. [62]
    Lifetime Modelling Issues of Power Light Emitting Diodes - MDPI
    Determine the pre-exponential factor A and the activation energy E a of the Arrhenius-equation from the continuously increasing junction temperature values of ...
  63. [63]
    Overview of high-power LED life prediction algorithms - Frontiers
    Mar 3, 2024 · Arrhenius derives its life acceleration model from activation energy, focusing solely on thermal acceleration factors as the degradation stress.
  64. [64]
    [PDF] UC Santa Barbara - eScholarship
    ... dislocation climb (REDC), a failure mechanism ... dislocations, discussing dislocations as non-radiative centers and recombination enhanced dislocation motion.<|separator|>
  65. [65]
    [PDF] Role of p-GaN layer thickness in the degradation of InGaN-GaN ...
    The impurities involved in the diffusion processes possibly originate from the p-side of the devices, therefore a thicker p-GaN layer reduces the amount of ...
  66. [66]
    N-contacts degradation analysis of white flip chip LEDs during ...
    Abstract: Current and temperature aging have been conducted on flip chip high power light emitting diodes (LEDs) with Al-Ni-Ti-Au n-contacts.
  67. [67]
    Understanding LED Internal Thermal Resistance | DigiKey
    Sep 8, 2015 · This article takes a look at the definition of an LED's internal thermal resistance and its impact on the junction temperature of commercial products.Missing: SRH | Show results with:SRH
  68. [68]
    [PDF] Thermally enhanced blue light-emitting diode - Bose Fellows
    Sep 24, 2015 · (5), we can see that a higher temperature LED junction results in an exponentially shorter SRH lifetime and thus a lower IQE. An empirical ...
  69. [69]
    Current crowding as a major cause for InGaN LED degradation at ...
    Current crowding as a major cause for InGaN LED degradation at extreme high current density. Abstract: To increase the optical output of a lighting system, ...
  70. [70]
    [PDF] Delft University of Technology Corrosion Sensitivity of LED Packages
    Finally the package integrity for corrosion of LED packages is influenced by long-term storage or exposure to thermal shock or high humidity in which even the.Missing: induced | Show results with:induced
  71. [71]
    (PDF) Effects of Moist Environments on LED Module Reliability
    The results demonstrate that the light output of LEDs decreases as the environmental moisture changes. Moisture diffuses into the interfaces of the packaging ...
  72. [72]
    LED Lumen Maintenance and Reliability | LM-80 - Lumileds
    The LM-80 standard calls for operating a set of LED devices for a minimum of 6,000 hours under well controlled operating conditions (e.g. a constant DC drive ...
  73. [73]
    [PDF] Lifetime of White LEDs - Stanford
    For example, the latest high power white LED packages claim a lifetime of 50,000 hours to the 70% lumen maintenance level, when operated at. 700 milliamps ...Missing: modern 2010
  74. [74]
  75. [75]
    [PDF] QUANTUM DOTS – SEEDS OF NANOSCIENCE - Nobel Prize
    Oct 4, 2023 · Evidence of quantum size effects in colloidal nanoparticles was found in 1983 when Louis Brus and co-workers investigated CdS crystallites ...
  76. [76]
    Recent Progress of Quantum Dots Light‐Emitting Diodes: Materials ...
    May 30, 2024 · QLED device structures can be categorized into three major types based on the CTL used, that is, all-organic CTLs, organic-inorganic hybrid CTLs ...Design and Synthesis of... · Characteristics of QLEDs · The Development Trend of...
  77. [77]
    Controlling the influence of Auger recombination on the ... - NIH
    Oct 25, 2013 · Here the authors demonstrate that designs such as heterostructure quantum dots can considerably suppress Auger decay in light-emitting diodes.Missing: FWHM IQE
  78. [78]
    (CdSe)ZnS Core−Shell Quantum Dots: Synthesis and ...
    This paper describes the synthesis and characterization of a series of room-temperature high quantum yield (30%−50%) core−shell (CdSe)ZnS nanocrystallites with ...
  79. [79]
    Blinking effect in quantum dots, its suppression mechanism, and ...
    May 14, 2025 · The phenomenon of blinking is the consequence of charge carrier trapping that obstructs charge transport as well as radiative recombination in ...II. BLINKING MECHANISMS · Surface treatment with ligands... · Shell engineering
  80. [80]
    Highly efficient green InP-based quantum dot light-emitting diodes ...
    May 30, 2022 · InP-based quantum dot light-emitting diodes (QLEDs), as less toxic than Cd-free and Pb-free optoelectronic devices, have become the most ...
  81. [81]
    A review on the electroluminescence properties of quantum-dot light ...
    Therefore, this paper gives review on the representative electroluminescence (EL) properties, including working mechanisms proposed for the QLEDs, QD structure ...