Carrier generation and recombination

Last updated

In solid-state physics of semiconductors, carrier generation and carrier recombination are processes by which mobile charge carriers (electrons and electron holes) 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.

Contents

The electron–hole pair is the fundamental unit of generation and recombination in inorganic semiconductors, corresponding to an electron transitioning between the valence band and the conduction band where generation of an electron is a transition from the valence band to the conduction band and recombination leads to a reverse transition.

Overview

Electronic band structure of a semiconductor material. Electronic band diagram.svg
Electronic band structure of a semiconductor material.

Like other solids, semiconductor materials have an electronic band structure determined by the crystal properties of the material. Energy distribution among electrons is described by the Fermi level and the temperature of the electrons. At absolute zero temperature, all of the electrons have energy below the Fermi level; but at non-zero temperatures the energy levels are filled following a Fermi-Dirac distribution.

In undoped semiconductors the Fermi level lies in the middle of a forbidden band or band gap between two allowed bands called the valence band and the conduction band . The valence band, immediately below the forbidden band, is normally very nearly completely occupied. The conduction band, above the Fermi level, is normally nearly completely empty. Because the valence band is so nearly full, its electrons are not mobile, and cannot flow as electric current.

However, if an electron in the valence band acquires enough energy to reach the conduction band as a result of interaction with other electrons, holes, photons, or the vibrating crystal lattice itself, it can flow freely among the nearly empty conduction band energy states. Furthermore, it will also leave behind a hole that can flow like a physicaly charged particle.

Carrier generation describes processes by which electrons gain energy and move from the valence band to the conduction band, producing two mobile carriers; while recombination describes processes by which a conduction band electron loses energy and re-occupies the energy state of an electron hole in the valence band.

These processes must conserve quantized energy crystal momentum, and the vibrating lattice which plays a large role in conserving momentum as in collisions, photons can transfer very little momentum in relation to their energy.

Relation between generation and recombination

The following image shows change in excess carriers being generated (green:electrons and purple:holes) with increasing light intensity (generation rate /cm ) at the center of an intrinsic semiconductor bar. Electrons have higher diffusion constant than holes leading to fewer excess electrons at the center as compared to holes. Diffusion center.gif
The following image shows change in excess carriers being generated (green:electrons and purple:holes) with increasing light intensity (generation rate /cm ) at the center of an intrinsic semiconductor bar. Electrons have higher diffusion constant than holes leading to fewer excess electrons at the center as compared to holes.

Recombination and generation are always happening in semiconductors, both optically and thermally. As predicted by thermodynamics, a material at thermal equilibrium will have generation and recombination rates that are balanced so that the net charge carrier density remains constant. The resulting probability of occupation of energy states in each energy band is given by Fermi–Dirac statistics.

The product of the electron and hole densities ( and ) is a constant at equilibrium, maintained by recombination and generation occurring at equal rates. When there is a surplus of carriers (i.e., ), the rate of recombination becomes greater than the rate of generation, driving the system back towards equilibrium. Likewise, when there is a deficit of carriers (i.e., ), the generation rate becomes greater than the recombination rate, again driving the system back towards equilibrium. [1] As the electron moves from one energy band to another, the energy and momentum that it has lost or gained must go to or come from the other particles involved in the process (e.g. photons, electron, or the system of vibrating lattice atoms).

Carrier generation

When light interacts with a material, it can either be absorbed (generating a pair of free carriers or an exciton) or it can stimulate a recombination event. The generated photon has similar properties to the one responsible for the event. Absorption is the active process in photodiodes, solar cells and other semiconductor photodetectors, while stimulated emission is the principle of operation in laser diodes.

Besides light excitation, carriers in semiconductors can also be generated by an external electric field, for example in light-emitting diodes and transistors.

When light with sufficient energy hits a semiconductor, it can excite electrons across the band gap. This generates additional charge carriers, temporarily lowering the electrical resistance of materials. This higher conductivity in the presence of light is known as photoconductivity. This conversion of light into electricity is widely used in photodiodes.

Recombination mechanisms

Carrier recombination can happen through multiple relaxation channels. The main ones are band-to-band recombination, Shockley–Read–Hall (SRH) trap-assisted recombination, Auger recombination and surface recombination. These decay channels can be separated into radiative and non-radiative. The latter occurs when the excess energy is converted into heat by phonon emission after the mean lifetime , whereas in the former at least part of the energy is released by light emission or luminescence after a radiative lifetime . The carrier lifetime is then obtained from the rate of both type of events according to: [2]

From which we can also define the internal quantum efficiency or quantum yield, as:

Radiative recombination

Band-to-band radiative recombination

Band-to-band recombination is the name for the process of electrons jumping down from the conduction band to the valence band in a radiative manner. During band-to-band recombination, a form of spontaneous emission, the energy absorbed by a material is released in the form of photons. Generally these photons contain the same or less energy than those initially absorbed. This effect is how LEDs create light. Because the photon carries relatively little momentum, radiative recombination is significant only in direct bandgap materials. This process is also known as bimolecular recombination [3] .

This type of recombination depends on the density of electrons and holes in the excited state, denoted by and respectively. Let us represent the radiative recombination as and the carrier generation rate as G.

Total generation is the sum of thermal generation G0 and generation due to light shining on the semiconductor GL:

Here we will consider the case in which there is no illumination on the semiconductor. Therefore and , and we can express the change in carrier density as a function of time as

Because the rate of recombination is affected by both the concentration of free electrons and the concentration of holes that are available to them, we know that Rr should be proportional to np:

and we add a proportionality constant Br to eliminate the sign:

If the semiconductor is in thermal equilibrium, the rate at which electrons and holes recombine must be balanced by the rate at which they are generated by the spontaneous transition of an electron from the valence band to the conduction band. The recombination rate must be exactly balanced by the thermal generation rate . [4]

Therefore:

where and are the equilibrium carrier densities. Using the mass action law ,with being the intrinsic carrier density, we can rewrite it as

The non-equilibrium carrier densities are given by [5]

Then the new recombination rate becomes, [4] [5]

Because and , we can say that

In an n-type semiconductor,

and

thus

Net recombination is the rate at which excess holes disappear

Solve this differential equation to get a standard exponential decay

where pmax is the maximum excess hole concentration when t = 0. (It can be proved that , but here we will not discuss that).

When , all of the excess holes will have disappeared. Therefore, we can define the lifetime of the excess holes in the material

So the lifetime of the minority carrier is dependent upon the majority carrier concentration.

Stimulated emission

Stimulated emission is a process where an incident photon interacts with an excited electron causing it to recombine and emit a photon with the same properties as the incident photon , in terms of phase, frequency, polarization, and direction of travel. Stimulated emission together with the principle of population inversion are at the heart of operation of lasers and masers. It has been shown by Einstein at the beginning of the twentieth century that if the excited and the ground level are non degenerate then the absorption rate and the stimulated emission rate are the same. [6] Else if level 1 and level 2 are -fold and -fold degenerate respectively, the new relation is:

Trap emission

Trap emission is a multistep process wherein a carrier falls into defect-related wave states in the middle of the bandgap. A trap is a defect capable of holding a carrier. The trap emission process recombines electrons with holes and emits photons to conserve energy. Due to the multistep nature of trap emission, a phonon is also often emitted. Trap emission can proceed by use of bulk defects [7] or surface defects. [8]

Non-radiative recombination

Non-radiative recombination is a process in phosphors and semiconductors, whereby charge carriers recombine releasing phonons instead of photons. Non-radiative recombination in optoelectronics and phosphors is an unwanted process, lowering the light generation efficiency and increasing heat losses.

Non-radiative life time is the average time before an electron in the conduction band of a semiconductor recombines with a hole. It is an important parameter in optoelectronics where radiative recombination is required to produce a photon; if the non-radiative life time is shorter than the radiative, a carrier is more likely to recombine non-radiatively. This results in low internal quantum efficiency.

Shockley–Read–Hall (SRH)

In Shockley-Read-Hall recombination (SRH), also called trap-assisted recombination, the electron in transition between bands passes through a new energy state (localized state) created within the band gap by a dopant or a defect in the crystal lattice; such energy states are called traps. Non-radiative recombination occurs primarily at such sites. The energy is exchanged in the form of lattice vibration, a phonon exchanging thermal energy with the material.

Since traps can absorb differences in momentum between the carriers, SRH is the dominant recombination process in silicon and other indirect bandgap materials. However, trap-assisted recombination can also dominate in direct bandgap materials under conditions of very low carrier densities (very low level injection) or in materials with high density of traps such as perovskites. The process is named after William Shockley, William Thornton Read [9] and Robert N. Hall, [10] who published it in 1952.

Types of traps

Electron traps vs. hole traps

Even though all the recombination events can be described in terms of electron movements, it is common to visualize the different processes in terms of excited electron and the electron holes they leave behind. In this context, if trap levels are close to the conduction band, they can temporarily immobilize excited electrons or in other words, they are electron traps. On the other hand, if their energy lies close to the valence band they become hole traps.

Shallow traps vs. deep traps

The distinction between shallow and deep traps is commonly made depending on how close electron traps are to the conduction band and how close hole traps are to the valence band. If the difference between trap and band is smaller than the thermal energy kBT it is often said that it is a shallow trap. Alternatively, if the difference is larger than the thermal energy, it is called a deep trap. This difference is useful because shallow traps can be emptied more easily and thus are often not as detrimental to the performance of optoelectronic devices.

SRH model

Electron and hole trapping in the Shockley-Read-Hall model Shockley-Read-Hall model inorganic semiconductors.png
Electron and hole trapping in the Shockley-Read-Hall model

In the SRH model, four things can happen involving trap levels: [11]

  • An electron in the conduction band can be trapped in an intragap state.
  • An electron can be emitted into the conduction band from a trap level.
  • A hole in the valence band can be captured by a trap. This is analogous to a filled trap releasing an electron into the valence band.
  • A captured hole can be released into the valence band. Analogous to the capture of an electron from the valence band.

When carrier recombination occurs through traps, we can replace the valence density of states by that of the intragap state. [12] The term is replaced by the density of trapped electrons/holes .

Where is the density of trap states and is the probability of that occupied state. Considering a material containing both types of traps, we can define two trapping coefficients and two de-trapping coefficients . In equilibrium, both trapping and de-trapping should be balanced ( and ). Then, the four rates as a function of become:

Where and are the electron and hole densities when the quasi Fermi level matches the trap energy. In steady-state condition, the net recombination rate of electrons should match the net recombination rate for holes, in other words: . This eliminates the occupation probability and leads to the Shockley-Read-Hall expression for the trap-assisted recombination:

Where the average lifetime for electrons and holes are defined as: [12]

Auger recombination

In Auger recombination the energy is given to a third carrier which is excited to a higher energy level without moving to another energy band. After the interaction, the third carrier normally loses its excess energy to thermal vibrations. Since this process is a three-particle interaction, it is normally only significant in non-equilibrium conditions when the carrier density is very high. The Auger effect process is not easily produced, because the third particle would have to begin the process in the unstable high-energy state.

In thermal equilibrium the Auger recombination and thermal generation rate equal each other [13]

where are the Auger capture probabilities. The non-equilibrium Auger recombination rate and resulting net recombination rate under steady-state conditions are [13]

The Auger lifetime is given by [14]

The mechanism causing LED efficiency droop was identified in 2007 as Auger recombination, which met with a mixed reaction. [15] In 2013, an experimental study claimed to have identified Auger recombination as the cause of efficiency droop. [16] However, it remains disputed whether the amount of Auger loss found in this study is sufficient to explain the droop. Other frequently quoted evidence against Auger as the main droop-causing mechanism is the low-temperature dependence of this mechanism, which is the opposite of that found for the droop.

Surface recombination

Trap-assisted recombination at the surface of a semiconductor is referred to as surface recombination. This occurs when traps at or near the surface or interface of the semiconductor form due to dangling bonds caused by the sudden discontinuation of the semiconductor crystal. Surface recombination is characterized by surface recombination velocity which depends on the density of surface defects. [17] In applications such as solar cells, surface recombination may be the dominant mechanism of recombination due to the collection and extraction of free carriers at the surface. In some applications of solar cells, a layer of transparent material with a large band gap, also known as a window layer, is used to minimize surface recombination. Passivation techniques are also employed to minimize surface recombination. [18]

Langevin recombination

For free carriers in low-mobility systems, the recombination rate is often described with the Langevin recombination rate. [19] The model is often used for disordered systems such as organic materials (and is hence relevant for organic solar cells [20] ) and other such systems. The Langevin recombination strength is defined as .

See also

Related Research Articles

<span class="mw-page-title-main">Photoluminescence</span> Light emission from substances after they absorb photons

Photoluminescence is light emission from any form of matter after the absorption of photons. It is one of many forms of luminescence and is initiated by photoexcitation, hence the prefix photo-. Following excitation, various relaxation processes typically occur in which other photons are re-radiated. Time periods between absorption and emission may vary: ranging from short femtosecond-regime for emission involving free-carrier plasma in inorganic semiconductors up to milliseconds for phosphoresence processes in molecular systems; and under special circumstances delay of emission may even span to minutes or hours.

<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.

The laser diode rate equations model the electrical and optical performance of a laser diode. This system of ordinary differential equations relates the number or density of photons and charge carriers (electrons) in the device to the injection current and to device and material parameters such as carrier lifetime, photon lifetime, and the optical gain.

<span class="mw-page-title-main">Quantum well</span> Concept in quantum mechanics

A quantum well is a potential well with only discrete energy values.

In solid-state physics, the electron mobility characterises how quickly an electron can move through a metal or semiconductor when pulled by an electric field. There is an analogous quantity for holes, called hole mobility. The term carrier mobility refers in general to both electron and hole mobility.

In semiconductor physics, the depletion region, also called depletion layer, depletion zone, junction region, space charge region, or space charge layer, is an insulating region within a conductive, doped semiconductor material where the mobile charge carriers have diffused away, or forced away by an electric field. The only elements left in the depletion region are ionized donor or acceptor impurities. This region of uncovered positive and negative ions is called the depletion region due to the depletion of carriers in this region, leaving none to carry a current. Understanding the depletion region is key to explaining modern semiconductor electronics: diodes, bipolar junction transistors, field-effect transistors, and variable capacitance diodes all rely on depletion region phenomena.

<span class="mw-page-title-main">Mercury cadmium telluride</span> Alloy

Hg1−xCdxTe or mercury cadmium telluride is a chemical compound of cadmium telluride (CdTe) and mercury telluride (HgTe) with a tunable bandgap spanning the shortwave infrared to the very long wave infrared regions. The amount of cadmium (Cd) in the alloy can be chosen so as to tune the optical absorption of the material to the desired infrared wavelength. CdTe is a semiconductor with a bandgap of approximately 1.5 electronvolts (eV) at room temperature. HgTe is a semimetal, which means that its bandgap energy is zero. Mixing these two substances allows one to obtain any bandgap between 0 and 1.5 eV.

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.

In semiconductor physics, the Haynes–Shockley experiment was an experiment that demonstrated that diffusion of minority carriers in a semiconductor could result in a current. The experiment was reported in a short paper by Haynes and Shockley in 1948, with a more detailed version published by Shockley, Pearson, and Haynes in 1949. The experiment can be used to measure carrier mobility, carrier lifetime, and diffusion coefficient.

A definition in semiconductor physics, carrier lifetime is defined as the average time it takes for a minority carrier to recombine. The process through which this is done is typically known as minority carrier recombination.

A quasi Fermi level is a term used in quantum mechanics and especially in solid state physics for the Fermi level that describes the population of electrons separately in the conduction band and valence band, when their populations are displaced from equilibrium. This displacement could be caused by the application of an external voltage, or by exposure to light of energy , which alter the populations of electrons in the conduction band and valence band. Since recombination rate tends to be much slower than the energy relaxation rate within each band, the conduction band and valence band can each have an individual population that is internally in equilibrium, even though the bands are not in equilibrium with respect to exchange of electrons. The displacement from equilibrium is such that the carrier populations can no longer be described by a single Fermi level, however it is possible to describe using concept of separate quasi-Fermi levels for each band.

<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.

Defect types include atom vacancies, adatoms, steps, and kinks that occur most frequently at surfaces due to the finite material size causing crystal discontinuity. What all types of defects have in common, whether surface or bulk defects, is that they produce dangling bonds that have specific electron energy levels different from those of the bulk. This difference occurs because these states cannot be described with periodic Bloch waves due to the change in electron potential energy caused by the missing ion cores just outside the surface. Hence, these are localized states that require separate solutions to the Schrödinger equation so that electron energies can be properly described. The break in periodicity results in a decrease in conductivity due to defect scattering.

<span class="mw-page-title-main">Direct and indirect band gaps</span> Types of energy range in a solid where no electron states can exist

In semiconductor physics, the band gap of a semiconductor can be of two basic types, a direct band gap or an indirect band gap. The minimal-energy state in the conduction band and the maximal-energy state in the valence band are each characterized by a certain crystal momentum (k-vector) in the Brillouin zone. If the k-vectors are different, the material has an "indirect gap". The band gap is called "direct" if the crystal momentum of electrons and holes is the same in both the conduction band and the valence band; an electron can directly emit a photon. In an "indirect" gap, a photon cannot be emitted because the electron must pass through an intermediate state and transfer momentum to the crystal lattice.

Laser linewidth is the spectral linewidth of a laser beam.

This article provides a more detailed explanation of p–n diode behavior than is found in the articles p–n junction or diode.

Two-photon photovoltaic effect is an energy collection method based on two-photon absorption (TPA). The TPP effect can be thought of as the nonlinear equivalent of the traditional photovoltaic effect involving high optical intensities. This effect occurs when two photons are absorbed at the same time resulting in an electron-hole pair.

The interaction of matter with light, i.e., electromagnetic fields, is able to generate a coherent superposition of excited quantum states in the material. Coherent denotes the fact that the material excitations have a well defined phase relation which originates from the phase of the incident electromagnetic wave. Macroscopically, the superposition state of the material results in an optical polarization, i.e., a rapidly oscillating dipole density. The optical polarization is a genuine non-equilibrium quantity that decays to zero when the excited system relaxes to its equilibrium state after the electromagnetic pulse is switched off. Due to this decay which is called dephasing, coherent effects are observable only for a certain temporal duration after pulsed photoexcitation. Various materials such as atoms, molecules, metals, insulators, semiconductors are studied using coherent optical spectroscopy and such experiments and their theoretical analysis has revealed a wealth of insights on the involved matter states and their dynamical evolution.

Optoelectronic reciprocity relations relate properties of a diode under illumination to the photon emission of the same diode under applied voltage. The relations are useful for interpretation of luminescence based measurements of solar cells and modules and for the analysis of recombination losses in solar cells.

Photoconductance decay or Photoconductivity decay, is a non-destructive analytical technique used to measure the lifetime of minority charge carriers in a semiconductor, especially in silicon wafers. The technique studies the transient photoconductivity of a semiconductor sample during or after it is illuminated by a light pulse. Electron–hole pairs are first generated by the light pulse, and the photoconductivity of the sample declines as the carriers recombine.

References

  1. Elhami Khorasani, Arash; Schroder, Dieter K.; Alford, T. L. (2014). "Optically Excited MOS-Capacitor for Recombination Lifetime Measurement". IEEE Electron Device Letters. 35 (10): 986–988. Bibcode:2014IEDL...35..986K. doi:10.1109/LED.2014.2345058. S2CID   19785166.
  2. Pelant, Ivan; Valenta, Jan (2012-02-09), "Luminescence of disordered semiconductors", Luminescence Spectroscopy of Semiconductors, Oxford University Press, pp. 242–262, doi:10.1093/acprof:oso/9780199588336.003.0009, ISBN   9780199588336
  3. Stranks, Samuel D.; Burlakov, Victor M.; Leijtens, Tomas; Ball, James M.; Goriely, Alain; Snaith, Henry J. (2014-09-11). "Recombination Kinetics in Organic-Inorganic Perovskites: Excitons, Free Charge, and Subgap States". Physical Review Applied. 2 (3): 034007. Bibcode:2014PhRvP...2c4007S. doi:10.1103/PhysRevApplied.2.034007.
  4. 1 2 Li, Sheng S., ed. (2006). Semiconductor Physical Electronics. p. 140. doi:10.1007/0-387-37766-2. ISBN   978-0-387-28893-2.
  5. 1 2 NISOLI, MAURO. (2016). SEMICONDUCTOR PHOTONICS. SOCIETA EDITRICE ESCULAPIO. ISBN   978-8893850025. OCLC   964380194.
  6. Svelto. (1989). Principles of Lasers ... p. 3. OCLC   249201544.
  7. Blumenau (2001). "Dislocation Related Photoluminescence in Silicon". Physical Review Letters. 87 (18): 187404. Bibcode:2001PhRvL..87r7404B. doi:10.1103/PhysRevLett.87.187404.
  8. van Dijken, Addy; Meulenkamp, Eric A.; Vanmaekelbergh, Daniël; Meijerink, Andries (2000-03-01). "The Kinetics of the Radiative and Nonradiative Processes in Nanocrystalline ZnO Particles upon Photoexcitation". The Journal of Physical Chemistry B. 104 (8): 1715–1723. doi:10.1021/jp993327z. ISSN   1520-6106.
  9. Shockley, W.; Read, W. T. (1 September 1952). "Statistics of the Recombinations of Holes and Electrons". Physical Review. 87 (5): 835–842. Bibcode:1952PhRv...87..835S. doi:10.1103/PhysRev.87.835.
  10. Hall, R.N. (1951). "Germanium rectifier characteristics". Physical Review. 83 (1): 228.
  11. NISOLI, MAURO. (2016). SEMICONDUCTOR PHOTONICS. SOCIETA EDITRICE ESCULAPIO. ISBN   978-8893850025. OCLC   964380194.
  12. 1 2 Kandada, Ajay Ram Srimath; D'Innocenzo, Valerio; Lanzani, Guglielmo; Petrozza, Annamaria (2016), Da Como, Enrico; De Angelis, Filippo; Snaith, Henry; Walker, Alison (eds.), "Chapter 4. Photophysics of Hybrid Perovskites", Unconventional Thin Film Photovoltaics, Royal Society of Chemistry, pp. 107–140, doi:10.1039/9781782624066-00107, ISBN   9781782622932
  13. 1 2 Li, Sheng S., ed. (2006). Semiconductor Physical Electronics. p. 143. doi:10.1007/0-387-37766-2. ISBN   978-0-387-28893-2.
  14. Li, Sheng S., ed. (2006). Semiconductor Physical Electronics. p. 144. doi:10.1007/0-387-37766-2. ISBN   978-0-387-28893-2.
  15. 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". IEEE Spectrum
  16. Justin Iveland; Lucio Martinelli; Jacques Peretti; James S. Speck; Claude Weisbuch. "Cause of LED Efficiency Droop Finally Revealed". Physical Review Letters, 2013. Science Daily. Retrieved 23 April 2013.
  17. Nelson, Jenny (2003). The Physics of Solar Cells. London: Imperial College Press. p. 116. ISBN   978-1-86094-340-9.
  18. Eades, W.D.; Swanson, R.M. (1985). "Calculation of surface generation and recombination velocities at the Si-SiO2 interface". Journal of Applied Physics. 58 (11): 4267–4276. Bibcode:1985JAP....58.4267E. doi: 10.1063/1.335562 . ISSN   0021-8979.
  19. "Recombination in low mobility semiconductors: Langevin theory". 4 April 2008.
  20. Lakhwani, Girish; Rao, Akshay; Friend, Richard H. (2014). "Bimolecular Recombination in Organic Photovoltaics". Annual Review of Physical Chemistry. 65 (1): 557–581. Bibcode:2014ARPC...65..557L. doi: 10.1146/annurev-physchem-040513-103615 . ISSN   0066-426X. PMID   24423376.

Further reading