Light-emitting diode physics

Last updated

Light-emitting diodes (LEDs) produce light (or infrared radiation) by the recombination of electrons and electron holes in a semiconductor, a process called "electroluminescence". The wavelength of the light produced depends on the energy band gap of the semiconductors used. Since these materials have a high index of refraction, [note 1] design features of the devices such as special optical coatings and die shape are required to efficiently emit light. A LED is a long-lived light source, but certain mechanisms can cause slow loss of efficiency of the device or sudden failure. The wavelength of the light emitted is a function of the band gap of the semiconductor material used; materials such as gallium arsenide, and others, with various trace doping elements, are used to produce different colors of light. Another type of LED uses a quantum dot which can have its properties and wavelength adjusted by its size. Light-emitting diodes are widely used in indicator and display functions, and white LEDs are displacing other technologies for general illumination purposes.

Contents

Electroluminescence

The inner workings of an LED, showing circuit (top) and band diagram (bottom) PnJunction-LED-E.svg
The inner workings of an LED, showing circuit (top) and band diagram (bottom)

The p–n junction in any direct band gap material emits light when electric current flows through it. This is electroluminescence. Electrons cross from the n-region and recombine with the holes existing in the p-region. Free electrons are in the conduction band of energy levels, while holes are in the valence energy band. Thus the energy level of the holes is lower than the energy levels of the electrons. Some portion of the energy must be dissipated to recombine the electrons and the holes. This energy is emitted in the form of heat and light.

As indirect band gap materials the electrons dissipate energy in the form of heat within the crystalline silicon and germanium diodes, but in gallium arsenide phosphide (GaAsP) and gallium phosphide (GaP) semiconductors, the electrons dissipate energy by emitting photons. If the semiconductor is translucent, the junction becomes the source of light, thus becoming a light-emitting diode.

I-V diagram for a diode. An LED begins to emit light when more than 2 or 3 volts is applied in the forward direction. The reverse bias region uses a different vertical scale from the forward bias region to show that the leakage current is nearly constant with voltage until breakdown occurs. In forward bias, the current starts small but increases exponentially with voltage. Diode-IV-Curve.svg
I-V diagram for a diode. An LED begins to emit light when more than 2 or 3 volts is applied in the forward direction. The reverse bias region uses a different vertical scale from the forward bias region to show that the leakage current is nearly constant with voltage until breakdown occurs. In forward bias, the current starts small but increases exponentially with voltage.

The wavelength of the light emitted, and thus its color, depends on the band gap energy of the materials forming the p-n junction. In silicon or germanium diodes, the electrons and holes usually recombine by a non-radiative transition, which produces no optical emission, because these are indirect band gap materials. The materials used for the LED have a direct band gap with energies corresponding to near-infrared, visible, or near-ultraviolet light.

LED development began with infrared and red devices made with gallium arsenide. Advances in materials science have enabled making devices with ever-shorter wavelengths, emitting light in a variety of colors.

LEDs are usually built on an n-type substrate, with an electrode attached to the p-type layer deposited on its surface. P-type substrates, while less common, occur as well. Many commercial LEDs, especially GaN/InGaN, also use sapphire substrates.

Refractive index

Idealized example of light emission cones in a simple square semiconductor, for a single point-source emission zone. The left illustration is for a translucent wafer, while the right illustration shows the half-cones formed when the bottom layer is opaque. The light is emitted equally in all directions from the point-source, but can only escape the semiconductor's surface within a few degrees of perpendicular, illustrated by the cone shapes. When the critical angle is exceeded, photons are reflected internally. The areas between the cones represent the trapped light energy wasted as heat. LED-chip-20-deg-crti-angle - both types - crop.png
Idealized example of light emission cones in a simple square semiconductor, for a single point-source emission zone. The left illustration is for a translucent wafer, while the right illustration shows the half-cones formed when the bottom layer is opaque. The light is emitted equally in all directions from the point-source, but can only escape the semiconductor's surface within a few degrees of perpendicular, illustrated by the cone shapes. When the critical angle is exceeded, photons are reflected internally. The areas between the cones represent the trapped light energy wasted as heat.

Bare uncoated semiconductors such as silicon exhibit a very high refractive index relative to air. Photons that approach the surface at too great an angle to the perpendicular experience total internal reflection. This property affects both the light-emission efficiency of LEDs as well as the light-absorption efficiency of photovoltaic cells. The refractive index of silicon is 3.96 (at 590 nm), [2] while air's refractive index is 1.0002926. [3]

In general, a flat-surface uncoated LED semiconductor chip emits only light that arrives nearly perpendicular to the semiconductor's surface, in a cone shape referred to as the light cone, cone of light, [4] or the escape cone. [1] Photons arriving at the surface more obliquely, with incidence angle exceeding the critical angle, undergo total internal reflection, and return inside the semiconductor crystal as if its surface were a mirror. [1]

Internal reflections can escape through other crystalline faces if the incidence angle is low enough and the crystal is sufficiently transparent to not re-absorb the photon emission. But for a simple square LED with 90-degree angled surfaces on all sides, the faces all act as equal angle mirrors. In this case, most of the light can not escape and is lost as waste heat in the crystal. [1]

A convoluted chip surface with angled facets similar to a jewel or fresnel lens can increase light output by distributing light perpendicular to the chip surface and far to the sides of the photon emission point. [5]

The ideal shape of a semiconductor with maximum light output would be a microsphere with the photon emission occurring at the exact center, with electrodes penetrating to the center to contact at the emission point. All light rays emanating from the center would be perpendicular to the entire surface of the sphere, resulting in no internal reflections. A hemispherical semiconductor would also work, with the flat back-surface serving as a mirror to back-scattered photons. [6]

Transition coatings

After the doping of the wafer, it is usually cut apart into individual dies. Each die is commonly called a chip.

Many LED semiconductor chips are encapsulated or potted in clear or colored molded solid plastic. The plastic encapsulation has three purposes:

  1. Mounting the semiconductor chip in devices is easier to accomplish.
  2. The tiny fragile electrical wiring is physically supported and protected from damage.
  3. The plastic acts as a refractive intermediary between the relatively high-index semiconductor and low-index open air. [7]

The third feature helps to boost the light emission from the semiconductor by reducing Fresnel reflections of photons within the light cone. A flat coating does not directly increase the size of the light cone in the semiconductor; it provides an intermediate wider cone angle in the coating, but the critical angle between rays in the semiconductor and in the air beyond the coating does not change. With a curved coating or encapsulation, however, efficiency can be further increased.

Efficiency and operational parameters

Typical indicator LEDs are designed to operate with no more than 30–60 milliwatts (mW) of electrical power. Around 1999, Philips Lumileds introduced power LEDs capable of continuous use at one watt. These LEDs used much larger semiconductor die sizes to handle the large power inputs. Also, the semiconductor dies were mounted onto metal slugs to allow for greater heat dissipation from the LED die.

One of the key advantages of LED-based lighting sources is high luminous efficacy. White LEDs quickly matched and overtook the efficacy of standard incandescent lighting systems. In 2002, Lumileds made five-watt LEDs available with luminous efficacy of 18–22 lumens per watt (lm/W). For comparison, a conventional incandescent light bulb of 60–100 watts emits around 15 lm/W, and standard fluorescent lights emit up to 100 lm/W.

As of 2012, Philips had achieved the following efficacies for each color. [8] The efficiency values show the physics – light power out per electrical power in. The lumen-per-watt efficacy value includes characteristics of the human eye and is derived using the luminosity function.

ColorWavelength range (nm)Typical efficiency coefficientTypical efficacy (lm/W)
Red 620 < λ < 6450.3972
Red-orange 610 < λ < 6200.2998
Green 520 < λ < 5500.1593
Cyan 490 < λ < 5200.2675
Blue 460 < λ < 4900.3537

In September 2003, a new type of blue LED was demonstrated by Cree. This produced a commercially packaged white light giving 65 lm/W at 20 mA, becoming the brightest white LED commercially available at the time, and more than four times as efficient as standard incandescents. In 2006, they demonstrated a prototype with a record white LED luminous efficacy of 131 lm/W at 20 mA. Nichia Corporation has developed a white LED with luminous efficacy of 150 lm/W at a forward current of 20 mA. [9] Cree's XLamp XM-L LEDs, commercially available in 2011, produce 100 lm/W at their full power of 10 W, and up to 160 lm/W at around 2 W input power. In 2012, Cree announced a white LED giving 254 lm/W, [10] and 303 lm/W in March 2014. [11] Practical general lighting needs high-power LEDs, of one watt or more. Typical operating currents for such devices begin at 350 mA.

These efficiencies are for the light-emitting diode only, held at low temperature in a lab. Since LEDs installed in real fixtures operate at higher temperature and with driver losses, real-world efficiencies are much lower. United States Department of Energy (DOE) testing of commercial LED lamps designed to replace incandescent lamps or CFLs showed that average efficacy was still about 46 lm/W in 2009 (tested performance ranged from 17 lm/W to 79 lm/W). [12]

Efficiency droop

Efficiency droop is the decrease in luminous efficacy of LEDs as the electric current increases.

This effect was initially thought related to elevated temperatures. Scientists proved the opposite is true: though the life of an LED is shortened, the efficiency droop is less severe at elevated temperatures. [13] The mechanism causing efficiency droop was identified in 2007 as Auger recombination. [14] [15]

In addition to being less efficient, operating LEDs at higher electric currents creates more heat, which can compromise LED lifetime. High-brightness LEDs often operate at 350 mA, which is a compromise between light output, efficiency, and longevity. [14]

Instead of increasing current levels, luminance is usually increased by combining multiple LEDs in one bulb. Solving the problem of efficiency droop would mean that household LED light bulbs would need fewer LEDs, which would significantly reduce costs.

Researchers at the U.S. Naval Research Laboratory have found a way to lessen the efficiency droop. They found that the droop arises from non-radiative Auger recombination of the injected carriers. They created quantum wells with a soft confinement potential to lessen the non-radiative Auger processes. [16]

Researchers at Taiwan National Central University and Epistar Corp are developing a way to lessen the efficiency droop by using ceramic aluminium nitride (AlN) substrates, which are more thermally conductive than the commercially used sapphire. The higher thermal conductivity reduces self-heating effects. [17]

Lifetime and failure

Solid-state devices such as LEDs are subject to very limited wear and tear if operated at low currents and at low temperatures. Typical lifetimes quoted are 25,000 to 100,000 hours, but heat and current settings can extend or shorten this time significantly. [18] It is important to note that these projections are based on a standard test that may not accelerate all the potential mechanisms that can induce failures in LEDs. [19]

The most common symptom of LED failure is the gradual lowering of light output. Sudden failures, although rare, can also occur. Early red LEDs were notable for their short service life. With the development of high-power LEDs, the devices are subjected to higher junction temperatures and higher current densities than traditional devices. This causes stress on the material and may cause early light-output degradation. Lifetime of a LED may be given as the running time to 70% or 50% of initial output. [20]

Unlike combustion or incandescent lamps, LEDs only operate if they are kept cool enough. The manufacturer commonly specifies a maximum junction temperature of 125 or 150 °C, and lower temperatures are advisable in the interests of long life. At these temperatures, relatively little heat is lost by radiation, which means that the light beam generated by an LED is cool.

The waste heat in a high-power LED is conducted through the device to a heat sink, which dissipates heat to the surrounding air. Since the maximum operating temperature of the LED is limited, the thermal resistances of the package, the heat sink and the interface must be calculated. Medium-power LEDs are often designed to solder directly to a printed circuit board that contains a thermally conductive metal layer. High-power LEDs are packaged in large-area ceramic packages that attach to a metal heat sink using thermal grease or other material to conduct heat.

If an LED lamp does not have free air circulation, the LED is likely to overheat, resulting in reduced life or early failure. The thermal design of the system must allow for the ambient temperature surrounding the lamp; a lamp in a freezer experiences a lower ambient than a lamp in a billboard in a sunny climate. [21]

Materials

LEDs are made from a variety of inorganic semiconductor materials. The following table shows the available colors with wavelength range, voltage drop, and material:

Color Wavelength [nm] Voltage drop [ΔV]Semiconductor material
Infrared λ > 760 Δ V < 1.63 Gallium arsenide (GaAs)
Aluminium gallium arsenide (AlGaAs)
Red 610 < λ < 7601.63 < ΔV < 2.03 Aluminium gallium arsenide (AlGaAs)
Gallium arsenide phosphide (GaAsP)
Aluminium gallium indium phosphide (AlGaInP)
Gallium(III) phosphide (GaP)
Orange 590 < λ < 6102.03 < ΔV < 2.10 Gallium arsenide phosphide (GaAsP)
Aluminium gallium indium phosphide (AlGaInP)
Gallium(III) phosphide (GaP)
Yellow 570 < λ < 5902.10 < ΔV < 2.18 Gallium arsenide phosphide (GaAsP)
Aluminium gallium indium phosphide (AlGaInP)
Gallium(III) phosphide (GaP)
Green 500 < λ < 5701.9 [22] < ΔV < 4.0Traditional green:
Gallium(III) phosphide (GaP)
Aluminium gallium indium phosphide (AlGaInP)
Aluminium gallium phosphide (AlGaP)
Pure green:
Indium gallium nitride (InGaN) / Gallium(III) nitride (GaN)
Blue 450 < λ < 5002.48 < ΔV < 3.7 Zinc selenide (ZnSe)
Indium gallium nitride (InGaN)
Synthetic sapphire, Silicon carbide (SiC) as substrate with or without epitaxy,
Silicon (Si) as substrate—under development (epitaxy on silicon is hard to control)
Violet 400 < λ < 4502.76 < ΔV < 4.0 Indium gallium nitride (InGaN)
Ultraviolet λ < 4003 < ΔV < 4.1 Indium gallium nitride (InGaN) (385-400 nm)

Diamond (235 nm) [23]
Boron nitride (215 nm) [24] [25]
Aluminium nitride (AlN) (210 nm) [26]
Aluminium gallium nitride (AlGaN)
Aluminium gallium indium nitride (AlGaInN)—down to 210 nm [27]

Pink Multiple typesΔV ≈3.3 [28] Blue with one or two phosphor layers,
yellow with red, orange or pink phosphor added afterwards,

white with pink plastic,
or white phosphors with pink pigment or dye over top. [29]

Purple Multiple types2.48 < ΔV < 3.7Dual blue/red LEDs,
blue with red phosphor,
or white with purple plastic
WhiteBroad spectrum2.8 < ΔV < 4.2Cool / Pure White: Blue/UV diode with yellow phosphor
Warm White: Blue diode with orange phosphor

Quantum-dot LEDs

Quantum dots (QD) are semiconductor nanocrystals with optical properties that let their emission color be tuned from the visible into the infrared spectrum. [30] [31] This allows quantum dot LEDs to create almost any color on the CIE diagram. This provides more color options and better color rendering than white LEDs since the emission spectrum is much narrower, characteristic of quantum confined states.

There are two types of schemes for QD excitation. One uses photo excitation with a primary light source LED (typically blue or UV LEDs are used). The other is direct electrical excitation first demonstrated by Alivisatos et al. [32]

One example of the photo-excitation scheme is a method developed by Michael Bowers, at Vanderbilt University in Nashville, involving coating a blue LED with quantum dots that glow white in response to the blue light from the LED. This method emits a warm, yellowish-white light similar to that made by incandescent light bulbs. [33] Quantum dots are also being considered for use in white light-emitting diodes in liquid crystal display (LCD) televisions. [34]

In February 2011 scientists at PlasmaChem GmbH were able to synthesize quantum dots for LED applications and build a light converter on their basis, which was able to efficiently convert light from blue to any other color for many hundred hours. [35] Such QDs can be used to emit visible or near infrared light of any wavelength being excited by light with a shorter wavelength.

The structure of QD-LEDs used for the electrical-excitation scheme is similar to basic design of OLEDs. A layer of quantum dots is sandwiched between layers of electron-transporting and hole-transporting materials. An applied electric field causes electrons and holes to move into the quantum dot layer and recombine forming an exciton that excites a QD. This scheme is commonly studied for quantum dot display. The tunability of emission wavelengths and narrow bandwidth is also beneficial as excitation sources for fluorescence imaging. Fluorescence near-field scanning optical microscopy (NSOM) utilizing an integrated QD-LED has been demonstrated. [36]

In February 2008, a luminous efficacy of 300 lumens of visible light per watt of radiation (not per electrical watt) and warm-light emission was achieved by using nanocrystals. [37]

Notes

  1. thereby prone to internal reflection, i.e. not actually emitting the light to the outside

Related Research Articles

<span class="mw-page-title-main">Laser</span> Device which emits light via optical amplification

A laser is a device that emits light through a process of optical amplification based on the stimulated emission of electromagnetic radiation. The word laser is an anacronym that originated as an acronym for light amplification by stimulated emission of radiation. The first laser was built in 1960 by Theodore Maiman at Hughes Research Laboratories, based on theoretical work by Charles H. Townes and Arthur Leonard Schawlow.

<span class="mw-page-title-main">Light-emitting diode</span> Semiconductor and solid-state light source

A light-emitting diode (LED) is a semiconductor device that emits light when current flows through it. Electrons in the semiconductor recombine with electron holes, releasing energy in the form of photons. The color of the light is determined by the energy required for electrons to cross the band gap of the semiconductor. White light is obtained by using multiple semiconductors or a layer of light-emitting phosphor on the semiconductor device.

<span class="mw-page-title-main">Cathodoluminescence</span> Photon emission under the impact of an electron beam

Cathodoluminescence is an optical and electromagnetic phenomenon in which electrons impacting on a luminescent material such as a phosphor, cause the emission of photons which may have wavelengths in the visible spectrum. A familiar example is the generation of light by an electron beam scanning the phosphor-coated inner surface of the screen of a television that uses a cathode ray tube. Cathodoluminescence is the inverse of the photoelectric effect, in which electron emission is induced by irradiation with photons.

<span class="mw-page-title-main">Band gap</span> Energy range in a solid where no electron states exist

In solid-state physics and solid-state chemistry, a band gap, also called a bandgap or energy gap, is an energy range in a solid where no electronic states exist. In graphs of the electronic band structure of solids, the band gap refers to the energy difference between the top of the valence band and the bottom of the conduction band in insulators and semiconductors. It is the energy required to promote an electron from the valence band to the conduction band. The resulting conduction-band electron are free to move within the crystal lattice and serve as charge carriers to conduct electric current. It is closely related to the HOMO/LUMO gap in chemistry. If the valence band is completely full and the conduction band is completely empty, then electrons cannot move within the solid because there are no available states. If the electrons are not free to move within the crystal lattice, then there is no generated current due to no net charge carrier mobility. However, if some electrons transfer from the valence band to the conduction band, then current can flow. Therefore, the band gap is a major factor determining the electrical conductivity of a solid. Substances having large band gaps are generally insulators, those with small band gaps are semiconductor, and conductors either have very small band gaps or none, because the valence and conduction bands overlap to form a continuous band.

<span class="mw-page-title-main">Laser diode</span> Semiconductor laser

A laser diode is a semiconductor device similar to a light-emitting diode in which a diode pumped directly with electrical current can create lasing conditions at the diode's junction.

<span class="mw-page-title-main">Emission spectrum</span> Frequencies of light emitted by atoms or chemical compounds

The emission spectrum of a chemical element or chemical compound is the spectrum of frequencies of electromagnetic radiation emitted due to electrons making a transition from a high energy state to a lower energy state. The photon energy of the emitted photons is equal to the energy difference between the two states. There are many possible electron transitions for each atom, and each transition has a specific energy difference. This collection of different transitions, leading to different radiated wavelengths, make up an emission spectrum. Each element's emission spectrum is unique. Therefore, spectroscopy can be used to identify elements in matter of unknown composition. Similarly, the emission spectra of molecules can be used in chemical analysis of substances.

<span class="mw-page-title-main">Vertical-cavity surface-emitting laser</span> Type of semiconductor laser diode

The vertical-cavity surface-emitting laser is a type of semiconductor laser diode with laser beam emission perpendicular from the top surface, contrary to conventional edge-emitting semiconductor lasers which emit from surfaces formed by cleaving the individual chip out of a wafer. VCSELs are used in various laser products, including computer mice, fiber optic communications, laser printers, Face ID, and smartglasses.

Wide-bandgap semiconductors are semiconductor materials which have a larger band gap than conventional semiconductors. Conventional semiconductors like silicon have a bandgap in the range of 0.6 – 1.5 electronvolt (eV), whereas wide-bandgap materials have bandgaps in the range above 2 eV. Generally, wide-bandgap semiconductors have electronic properties which fall in between those of conventional semiconductors and insulators.

<span class="mw-page-title-main">Quantum dot</span> Zero-dimensional, nano-scale semiconductor particles with novel optical and electronic properties

Quantum dots (QDs) or semiconductor nanocrystals are semiconductor particles a few nanometres in size with optical and electronic properties that differ from those of larger particles via quantum mechanical effects. They are a central topic in nanotechnology and materials science. When a quantum dot is illuminated by UV light, an electron in the quantum dot can be excited to a state of higher energy. In the case of a semiconducting quantum dot, this process corresponds to the transition of an electron from the valence band to the conductance band. The excited electron can drop back into the valence band releasing its energy as light. This light emission (photoluminescence) is illustrated in the figure on the right. The color of that light depends on the energy difference between the conductance band and the valence band, or the transition between discrete energy states when the band structure is no longer well-defined in QDs.

<span class="mw-page-title-main">Photodetector</span> Sensors of light or other electromagnetic energy

Photodetectors, also called photosensors, are sensors of light or other electromagnetic radiation. There are a wide variety of photodetectors which may be classified by mechanism of detection, such as photoelectric or photochemical effects, or by various performance metrics, such as spectral response. Semiconductor-based photodetectors typically use a p–n junction that converts photons into charge. The absorbed photons make electron–hole pairs in the depletion region. Photodiodes and photo transistors are a few examples of photo detectors. Solar cells convert some of the light energy absorbed into electrical energy.

Luminous efficacy is a measure of how well a light source produces visible light. It is the ratio of luminous flux to power, measured in lumens per watt in the International System of Units (SI). Depending on context, the power can be either the radiant flux of the source's output, or it can be the total power consumed by the source. Which sense of the term is intended must usually be inferred from the context, and is sometimes unclear. The former sense is sometimes called luminous efficacy of radiation, and the latter luminous efficacy of a light source or overall luminous efficacy.

In solid-state physics of semiconductors, carrier generation and carrier recombination are processes by which mobile charge carriers are created and eliminated. Carrier generation and recombination processes are fundamental to the operation of many optoelectronic semiconductor devices, such as photodiodes, light-emitting diodes and laser diodes. They are also critical to a full analysis of p-n junction devices such as bipolar junction transistors and p-n junction diodes.

Quantum-cascade lasers (QCLs) are semiconductor lasers that emit in the mid- to far-infrared portion of the electromagnetic spectrum and were first demonstrated by Jérôme Faist, Federico Capasso, Deborah Sivco, Carlo Sirtori, Albert Hutchinson, and Alfred Cho at Bell Laboratories in 1994.

Thermophotovoltaic (TPV) energy conversion is a direct conversion process from heat to electricity via photons. A basic thermophotovoltaic system consists of a hot object emitting thermal radiation and a photovoltaic cell similar to a solar cell but tuned to the spectrum being emitted from the hot object.

<span class="mw-page-title-main">Energy conversion efficiency</span> Ratio between the useful output and the input of a machine

Energy conversion efficiency (η) is the ratio between the useful output of an energy conversion machine and the input, in energy terms. The input, as well as the useful output may be chemical, electric power, mechanical work, light (radiation), or heat. The resulting value, η (eta), ranges between 0 and 1.

<span class="mw-page-title-main">LED lamp</span> Electric light that produces light using LEDs

An LED lamp or LED light is an electric light that produces light using light-emitting diodes (LEDs). LED lamps are significantly more energy-efficient than equivalent incandescent lamps and fluorescent lamps. The most efficient commercially available LED lamps have efficiencies exceeding 200 lumens per watt (lm/W) and convert more than half the input power into light. Commercial LED lamps have a lifespan several times longer than both incandescent and fluorescent lamps.

<span class="mw-page-title-main">Shockley–Queisser limit</span> Maximum theoretical efficiency of a solar cell

In physics, the radiative efficiency limit is the maximum theoretical efficiency of a solar cell using a single p-n junction to collect power from the cell where the only loss mechanism is radiative recombination in the solar cell. It was first calculated by William Shockley and Hans-Joachim Queisser at Shockley Semiconductor in 1961, giving a maximum efficiency of 30% at 1.1 eV. The limit is one of the most fundamental to solar energy production with photovoltaic cells, and is one of the field's most important contributions.

<span class="mw-page-title-main">Quantum dot display</span> Type of display device

A quantum dot display is a display device that uses quantum dots (QD), semiconductor nanocrystals which can produce pure monochromatic red, green, and blue light. Photo-emissive quantum dot particles are used in LCD backlights or display color filters. Quantum dots are excited by the blue light from the display panel to emit pure basic colors, which reduces light losses and color crosstalk in color filters, improving display brightness and color gamut. Light travels through QD layer film and traditional RGB filters made from color pigments, or through QD filters with red/green QD color converters and blue passthrough. Although the QD color filter technology is primarily used in LED-backlit LCDs, it is applicable to other display technologies which use color filters, such as blue/UV active-matrix organic light-emitting diode (AMOLED) or QNED/MicroLED display panels. LED-backlit LCDs are the main application of photo-emissive quantum dots, though blue organic light-emitting diode (OLED) panels with QD color filters are being researched.

<span class="mw-page-title-main">Interband cascade laser</span>

Interband cascade lasers (ICLs) are a type of laser diode that can produce coherent radiation over a large part of the mid-infrared region of the electromagnetic spectrum. They are fabricated from epitaxially-grown semiconductor heterostructures composed of layers of indium arsenide (InAs), gallium antimonide (GaSb), aluminum antimonide (AlSb), and related alloys. These lasers are similar to quantum cascade lasers (QCLs) in several ways. Like QCLs, ICLs employ the concept of bandstructure engineering to achieve an optimized laser design and reuse injected electrons to emit multiple photons. However, in ICLs, photons are generated with interband transitions, rather than the intersubband transitions used in QCLs. Consequently, the rate at which the carriers injected into the upper laser subband thermally relax to the lower subband is determined by interband Auger, radiative, and Shockley-Read carrier recombination. These processes typically occur on a much slower time scale than the longitudinal optical phonon interactions that mediates the intersubband relaxation of injected electrons in mid-IR QCLs. The use of interband transitions allows laser action in ICLs to be achieved at lower electrical input powers than is possible with QCLs.

References

  1. 1 2 3 4 Mueller, Gerd (2000) Electroluminescence I, Academic Press, ISBN   0-12-752173-9, p. 67, "escape cone of light" from semiconductor, illustrations of light cones on p. 69
  2. "Optical Properties of Silicon". PVCDROM.PVEducation.org. Archived from the original on 2009-06-05.
  3. Refraction — Snell's Law. Interactagram.com. Retrieved on March 16, 2012.
  4. Lipták, Bela G. (2005) Instrument Engineers' Handbook: Process control and optimization, CRC Press, ISBN   0-8493-1081-4 p. 537, "cone of light" in context of optical fibers
  5. Capper, Peter; Mauk, Michael (2007). Liquid phase epitaxy of electronic, optical, and optoelectronic materials. Wiley. p. 389. ISBN   978-0-470-85290-3. faceted structures are of interest for solar cells, LEDs, thermophotovoltaic devices, and detectors in that nonplanar surfaces and facets can enhance optical coupling and light-trapping effects, [with example microphotograph of a faceted crystal substrate].
  6. Dakin, John and Brown, Robert G. W. (eds.) Handbook of optoelectronics, Volume 2, Taylor & Francis, 2006 ISBN   0-7503-0646-7 p. 356, "Die shaping is a step towards the ideal solution, that of a point light source at the center of a spherical semiconductor die."
  7. Schubert, E. Fred (2006) Light-emitting diodes, Cambridge University Press, ISBN   0-521-86538-7 p. 97, "Epoxy Encapsulants", "The light extraction efficiency can be enhanced by using dome-shaped encapsulants with a large refractive index."
  8. "All in 1 LED Lighting Solutions Guide". PhilipsLumileds.com. Philips. 2012-10-04. p. 15. Archived from the original (PDF) on March 14, 2013. Retrieved 2015-11-18.
  9. "Nichia Unveils White LED with 150 lm/W Luminous Efficiency". Tech-On!. December 21, 2006. Retrieved August 13, 2007.
  10. "Cree Sets New Record for White LED Efficiency", Tech-On, April 23, 2012.
  11. "Cree First to Break 300 Lumens-Per-Watt Barrier", Cree news
  12. DOE Solid-State Lighting CALiPER Program Summary of Results: Round 9 of Product Testing (PDF). U.S. Department of Energy. October 2009.
  13. Identifying the Causes of LED Efficiency Droop Archived December 13, 2013, at the Wayback Machine , By Steven Keeping, Digi-Key Corporation Tech Zone
  14. 1 2 Stevenson, Richard (August 2009) The LED’s Dark Secret: Solid-state lighting won't supplant the lightbulb until it can overcome the mysterious malady known as droop Archived 2009-08-05 at the Wayback Machine . IEEE Spectrum
  15. Iveland, Justin; Martinelli, Lucio; Peretti, Jacques; Speck, James S.; Weisbuch, Claude. "Cause of LED Efficiency Droop Finally Revealed". Physical Review Letters, 2013. Science Daily. Retrieved 23 April 2013.
  16. McKinney, Donna (19 February 2014) A Roadmap to Efficient Green-Blue-Ultraviolet Light-Emitting Diodes, U.S. Naval Research Laboratory
  17. Cooke, Mike (11 February 2014) Enabling high-voltage InGaN LED operation with ceramic substrate, Semiconductor Today
  18. "Lifetime of White LEDs". Archived from the original on April 10, 2009. Retrieved 2009-04-10., US Department of Energy
  19. Arnold, J. When the Lights Go Out: LED Failure Modes and Mechanisms. DfR Solutions
  20. Narendran, N.; Y. Gu (2005). "Life of LED-based white light sources". Journal of Display Technology. 1 (1): 167–171. Bibcode:2005JDisT...1..167N. doi:10.1109/JDT.2005.852510. S2CID   32847174.
  21. Conway, K. M. and J. D. Bullough. 1999. Will LEDs transform traffic signals as they did exit signs? Proceedings of the Illuminating Engineering Society of North America Annual Conference (pp. 1–9), New Orleans, Louisiana, August 9–11. New York, NY: Illuminating Engineering Society of North America.
  22. OSRAM: green LED Archived July 21, 2011, at the Wayback Machine . osram-os.com. Retrieved on March 16, 2012.
  23. Koizumi, S.; Watanabe, K.; Hasegawa, M.; Kanda, H. (2001). "Ultraviolet Emission from a Diamond pn Junction". Science. 292 (5523): 1899–1901. Bibcode:2001Sci...292.1899K. doi:10.1126/science.1060258. PMID   11397942. S2CID   10675358.
  24. Kubota, Y.; Watanabe, K.; Tsuda, O.; Taniguchi, T. (2007). "Deep Ultraviolet Light-Emitting Hexagonal Boron Nitride Synthesized at Atmospheric Pressure". Science. 317 (5840): 932–934. Bibcode:2007Sci...317..932K. doi:10.1126/science.1144216. PMID   17702939. S2CID   196949298.
  25. Watanabe, K.; Taniguchi, T.; Kanda, H. (2004). "Direct-bandgap properties and evidence for ultraviolet lasing of hexagonal boron nitride single crystal". Nature Materials. 3 (6): 404–409. Bibcode:2004NatMa...3..404W. doi:10.1038/nmat1134. PMID   15156198. S2CID   23563849.
  26. Taniyasu, Y.; Kasu, M.; Makimoto, T. (2006). "An aluminium nitride light-emitting diode with a wavelength of 210 nanometres". Nature. 441 (7091): 325–328. Bibcode:2006Natur.441..325T. doi:10.1038/nature04760. PMID   16710416. S2CID   4373542.
  27. "LEDs move into the ultraviolet". physicsworld.com. May 17, 2006. Retrieved August 13, 2007.
  28. How to Wire/Connect LEDs Archived March 2, 2012, at the Wayback Machine . Llamma.com. Retrieved on March 16, 2012.
  29. LED types by Color, Brightness, and Chemistry. Donklipstein.com. Retrieved on March 16, 2012.
  30. Quantum-dot LED may be screen of choice for future electronics Massachusetts Institute of Technology News Office, December 18, 2002
  31. Neidhardt, H.; Wilhelm, L.; Zagrebnov, V. A. (February 2015). "A New Model for Quantum Dot Light Emitting-Absorbing Bevices: Proofs and Supplements". Nanosystems: Physics, Chemistry, Mathematics. 6 (1): 6–45. doi: 10.17586/2220-8054-2015-6-1-6-45 .
  32. Colvin, V. L.; Schlamp, M. C.; Alivisatos, A. P. (1994). "Light-emitting diodes made from cadmium selenide nanocrystals and a semiconducting polymer". Nature. 370 (6488): 354–357. Bibcode:1994Natur.370..354C. doi:10.1038/370354a0. S2CID   4324973.
  33. "Accidental Invention Points to End of Light Bulbs". LiveScience.com. October 21, 2005. Retrieved January 24, 2007.
  34. Nanoco Signs Agreement with Major Japanese Electronics Company, nanocogroup.com (September 23, 2009)
  35. Nanotechnologie Aktuell, pp. 98–99, v. 4, 2011, ISSN   1866-4997
  36. Hoshino, K.; Gopal, A.; Glaz, M. S.; Vanden Bout, D. A.; Zhang, X. (2012). "Nanoscale fluorescence imaging with quantum dot near-field electroluminescence". Applied Physics Letters. 101 (4): 043118. Bibcode:2012ApPhL.101d3118H. doi:10.1063/1.4739235.
  37. Inman, Mason (February 1, 2008). "Crystal Coat Warms up LED Light". newscientist.com. Retrieved January 30, 2012.