Semiconductor luminescence equations

Last updated

The semiconductor luminescence equations (SLEs) [1] [2] describe luminescence of semiconductors resulting from spontaneous recombination of electronic excitations, producing a flux of spontaneously emitted light. This description established the first step toward semiconductor quantum optics because the SLEs simultaneously includes the quantized light–matter interaction and the Coulomb-interaction coupling among electronic excitations within a semiconductor. The SLEs are one of the most accurate methods to describe light emission in semiconductors and they are suited for a systematic modeling of semiconductor emission ranging from excitonic luminescence to lasers.

Contents

Due to randomness of the vacuum-field fluctuations, semiconductor luminescence is incoherent whereas the extensions of the SLEs include [2] the possibility to study resonance fluorescence resulting from optical pumping with coherent laser light. At this level, one is often interested to control and access higher-order photon-correlation effects, distinct many-body states, as well as light–semiconductor entanglement. Such investigations are the basis of realizing and developing the field of quantum-optical spectroscopy which is a branch of quantum optics.

Starting point

The derivation of the SLEs starts from a system Hamiltonian that fully includes many-body interactions, quantized light field, and quantized light–matter interaction. Like almost always in many-body physics, it is most convenient to apply the second-quantization formalism. For example, a light field corresponding to frequency is then described through Boson creation and annihilation operators and , respectively, where the "hat" over signifies the operator nature of the quantity. The operator-combination determines the photon-number operator.

When the photon coherences, here the expectation value , vanish and the system becomes quasistationary, semiconductors emit incoherent light spontaneously, commonly referred to as luminescence (L). (This is the underlying principle behind light-emitting diodes.) The corresponding luminescence flux is proportional to the temporal change in photon number, [2]

As a result, the luminescence becomes directly generated by a photon-assisted electron–hole recombination,

that describes a correlated emission of a photon when an electron with wave vector recombines with a hole, i.e., an electronic vacancy. Here, determines the corresponding electron–hole recombination operator defining also the microscopic polarization within semiconductor. Therefore, can also be viewed as photon-assisted polarization.

Many electron–hole pairs contribute to the photon emission at frequency ; the explicit notation within denotes that the correlated part of the expectation value is constructed using the cluster-expansion approach. The quantity contains the dipole-matrix element for interband transition, light-mode's mode function, and vacuum-field amplitude.

Principal structure of SLEs

In general, the SLEs includes all single- and two-particle correlations needed to compute the luminescence spectrum self-consistently. More specifically, a systematic derivation produces a set of equations involving photon-number-like correlations

Semiconductor luminescence equations(photon-number-like correlations)

whose diagonal form reduces to the luminescence formula above. The dynamics of photon-assisted correlations follows from

Semiconductor luminescence equations(photon-assisted correlations)

where the first contribution, , contains the Coulomb-renormalized single-particle energy that is determined by the bandstructure of the solid. The Coulomb renormalization are identical to those that appear in the semiconductor Bloch equations (SBEs), showing that all photon-assisted polarizations are coupled with each other via the unscreened Coulomb-interaction . The three-particle correlations that appear are indicated symbolically via the contributions – they introduce excitation-induced dephasing, screening of Coulomb interaction, and additional highly correlated contributions such as phonon-sideband emission. The explicit form of a spontaneous-emission source and a stimulated contribution are discussed below.

The excitation level of a semiconductor is characterized by electron and hole occupations, and , respectively. They modify the via the Coulomb renormalizations and the Pauli-blocking factor, . These occupations are changed by spontaneous recombination of electrons and holes, yielding

In its full form, the occupation dynamics also contains Coulomb-correlation terms. [2] It is straight forward to verify that the photon-assisted recombination [3] [4] [5] destroys as many electron–hole pairs as it creates photons because due to the general conservation law .

Besides the terms already described above, the photon-assisted polarization dynamics contains a spontaneous-emission source

Intuitively, describes the probability to find electron and hole with same when electrons and holes are uncorrelated, i.e., plasma. Such form is to be expected for a probability of two uncorrelated events to occur simultaneously at a desired value. The possibility to have truly correlated electron–hole pairs is defined by a two-particle correlation ; the corresponding probability is directly proportional to the correlation. In practice, becomes large when electron–hole pairs are bound as excitons via their mutual Coulomb attraction. Nevertheless, both the presence of electron–hole plasma and excitons can equivalently induce the spontaneous-emission source.

As the semiconductor emits light spontaneously, the luminescence is further altered by a stimulated contribution

that is particularly important when describing spontaneous emission in semiconductor microcavities and lasers because then spontaneously emitted light can return to the emitter (i.e., the semiconductor), either stimulating or inhibiting further spontaneous-emission processes. This term is also responsible for the Purcell effect.

To complete the SLEs, one must additionally solve the quantum dynamics of exciton correlations

The first line contains the Coulomb-renormalized kinetic energy of electron–hole pairs and the second line defines a source that results from a Boltzmann-type in- and out-scattering of two electrons and two holes due to the Coulomb interaction. The second line contains the main Coulomb sums that correlate electron–hole pairs into excitons whenever the excitation conditions are suitable. The remaining two- and three-particle correlations are presented symbolically by and , respectively. [2] [6]

Interpretation and consequences

Microscopically, the luminescence processes are initiated whenever the semiconductor is excited because at least the electron and hole distributions, that enter the spontaneous-emission source, are nonvanishing. As a result, is finite and it drives the photon-assisted processes for all those values that correspond to the excited states. This means that is simultaneously generated for many values. Since the Coulomb interaction couples with all values, the characteristic transition energy follows from the exciton energy, not the bare kinetic energy of an electron–hole pair. More mathematically, the homogeneous part of the dynamics has eigenenergies that are defined by the generalized Wannier equation not the free-carrier energies. For low electron–hole densities, the Wannier equation produces a set of bound eigenstates which define the exciton resonances.

Buildup of photon-assisted polarization (P correlation) that is initiated by the spontaneous-emission source. The buildup occurs equally for all momentum states. In a many-body system, a photon (wave arrow) is generated collectively through multiple coupled P-transition correlations. Photoluminescence schema.png
Buildup of photon-assisted polarization (Π correlation) that is initiated by the spontaneous-emission source. The buildup occurs equally for all momentum states. In a many-body system, a photon (wave arrow) is generated collectively through multiple coupled Π-transition correlations.

Therefore, shows a discrete set of exciton resonances regardless which many-body state initiated the emission through the spontaneous-emission source. These resonances are directly transferred to excitonic peaks in the luminescence itself. This yields an unexpected consequence; the excitonic resonance can equally well originate from an electron–hole plasma or the presence of excitons. [7] At first, this consequence of SLEs seems counterintuitive because in few-particle picture an unbound electron–hole pair cannot recombine and release energy corresponding to the exciton resonance because that energy is well below the energy an unbound electron–hole pair possesses.

However, the excitonic plasma luminescence is a genuine many-body effect where plasma emits collectively to the exciton resonance. Namely, when a high number of electronic states participate in the emission of a single photon, one can always distribute the energy of initial many-body state between the one photon at exciton energy and remaining many-body state (with one electron–hole pair removed) without violating the energy conservation. The Coulomb interaction mediates such energy rearrangements very efficiently. A thorough analysis of energy and many-body state rearrangement is given in Ref. [2]

In general, excitonic plasma luminescence explains many nonequilibrium emission properties observed in present-day semiconductor luminescence experiments. In fact, the dominance of excitonic plasma luminescence has been measured in both quantum-well [8] and quantum-dot systems. [9] Only when excitons are present abundantly, the role of excitonic plasma luminescence can be ignored.

Connections and generalizations

Structurally, the SLEs resemble the semiconductor Bloch equations (SBEs) if the are compared with the microscopic polarization within the SBEs. As the main difference, also has a photon index , its dynamics is driven spontaneously, and it is directly coupled to three-particle correlations. Technically, the SLEs are more difficult to solve numerically than the SBEs due to the additional degree of freedom. However, the SLEs often are the only (at low carrier densities) or more convenient (lasing regime) to compute luminescence accurately. Furthermore, the SLEs not only yield a full predictability without the need for phenomenological approximations but they also can be used as a systematic starting point for more general investigations such as laser design [10] [11] and disorder studies. [12]

The presented SLEs discussion does not specify the dimensionality or the band structure of the system studied. As one analyses a specified system, one often has to explicitly include the electronic bands involved, the dimensionality of wave vectors, photon, and exciton center-of-mass momentum. Many explicit examples are given in Refs. [6] [13] for quantum-well and quantum-wire systems, and in Refs. [4] [14] [15] for quantum-dot systems.

Semiconductors also can show several resonances well below the fundamental exciton resonance when phonon-assisted electron–hole recombination takes place. These processes are describable by three-particle correlations (or higher) where photon, electron–hole pair, and a lattice vibration, i.e., a phonon, become correlated. The dynamics of phonon-assisted correlations are similar to the phonon-free SLEs. Like for the excitonic luminescence, also excitonic phonon sidebands can equally well be initiated by either electron–hole plasma or excitons. [16]

The SLEs can also be used as a systematic starting point for semiconductor quantum optics. [2] [17] [18] As a first step, one also includes two-photon absorption correlations, , and then continues toward higher-order photon-correlation effects. This approach can be applied to analyze the resonance fluorescence effects and to realize and understand the quantum-optical spectroscopy.

See also

Related Research Articles

Spontaneous emission is the process in which a quantum mechanical system transits from an excited energy state to a lower energy state and emits a quantized amount of energy in the form of a photon. Spontaneous emission is ultimately responsible for most of the light we see all around us; it is so ubiquitous that there are many names given to what is essentially the same process. If atoms are excited by some means other than heating, the spontaneous emission is called luminescence. For example, fireflies are luminescent. And there are different forms of luminescence depending on how excited atoms are produced. If the excitation is effected by the absorption of radiation the spontaneous emission is called fluorescence. Sometimes molecules have a metastable level and continue to fluoresce long after the exciting radiation is turned off; this is called phosphorescence. Figurines that glow in the dark are phosphorescent. Lasers start via spontaneous emission, then during continuous operation work by stimulated emission.

<span class="mw-page-title-main">Exciton</span> Quasiparticle which is a bound state of an electron and an electron hole

An electron and an electron hole that are attracted to each other by the Coulomb force can form a bound state called an exciton. It is an electrically neutral quasiparticle that exists mainly in condensed matter, including insulators, semiconductors, some metals, but also in certain atoms, molecules and liquids. The exciton is regarded as an elementary excitation that can transport energy without transporting net electric charge.

<span class="mw-page-title-main">Bremsstrahlung</span> Electromagnetic radiation due to deceleration of charged particles

In particle physics, bremsstrahlung is electromagnetic radiation produced by the deceleration of a charged particle when deflected by another charged particle, typically an electron by an atomic nucleus. The moving particle loses kinetic energy, which is converted into radiation, thus satisfying the law of conservation of energy. The term is also used to refer to the process of producing the radiation. Bremsstrahlung has a continuous spectrum, which becomes more intense and whose peak intensity shifts toward higher frequencies as the change of the energy of the decelerated particles increases.

<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.

In physics, screening is the damping of electric fields caused by the presence of mobile charge carriers. It is an important part of the behavior of charge-carrying fluids, such as ionized gases, electrolytes, and charge carriers in electronic conductors . In a fluid, with a given permittivity ε, composed of electrically charged constituent particles, each pair of particles interact through the Coulomb force as

<span class="mw-page-title-main">Density of states</span> Number of available physical states per energy unit

In solid-state physics and condensed matter physics, the density of states (DOS) of a system describes the number of modes per unit frequency range. The density of states is defined as , where is the number of states in the system of volume whose energies lie in the range from to . It is mathematically represented as a distribution by a probability density function, and it is generally an average over the space and time domains of the various states occupied by the system. The density of states is directly related to the dispersion relations of the properties of the system. High DOS at a specific energy level means that many states are available for occupation.

In quantum physics, Fermi's golden rule is a formula that describes the transition rate from one energy eigenstate of a quantum system to a group of energy eigenstates in a continuum, as a result of a weak perturbation. This transition rate is effectively independent of time and is proportional to the strength of the coupling between the initial and final states of the system as well as the density of states. It is also applicable when the final state is discrete, i.e. it is not part of a continuum, if there is some decoherence in the process, like relaxation or collision of the atoms, or like noise in the perturbation, in which case the density of states is replaced by the reciprocal of the decoherence bandwidth.

The Franz–Keldysh effect is a change in optical absorption by a semiconductor when an electric field is applied. The effect is named after the German physicist Walter Franz and Russian physicist Leonid Keldysh.

Resonance fluorescence is the process in which a two-level atom system interacts with the quantum electromagnetic field if the field is driven at a frequency near to the natural frequency of the atom.

The quantum-confined Stark effect (QCSE) describes the effect of an external electric field upon the light absorption spectrum or emission spectrum of a quantum well (QW). In the absence of an external electric field, electrons and holes within the quantum well may only occupy states within a discrete set of energy subbands. Only a discrete set of frequencies of light may be absorbed or emitted by the system. When an external electric field is applied, the electron states shift to lower energies, while the hole states shift to higher energies. This reduces the permitted light absorption or emission frequencies. Additionally, the external electric field shifts electrons and holes to opposite sides of the well, decreasing the overlap integral, which in turn reduces the recombination efficiency of the system. The spatial separation between the electrons and holes is limited by the presence of the potential barriers around the quantum well, meaning that excitons are able to exist in the system even under the influence of an electric field. The quantum-confined Stark effect is used in QCSE optical modulators, which allow optical communications signals to be switched on and off rapidly.

<span class="mw-page-title-main">Förster coupling</span> Resonant energy transfer between excitons within adjacent QDs (quantum dots)

Förster coupling is the resonant energy transfer between excitons within adjacent QD's. The first studies of Forster were performed in the context of the sensitized luminescence of solids. Here, an excited sensitizer atom can transfer its excitation to a neighbouring acceptor atom, via an intermediate virtual photon. This same mechanism has also been shown to be responsible for exciton transfer between QD's and within molecular systems and biosystems, all of which may be treated in a similar formulation.

Heat transfer physics describes the kinetics of energy storage, transport, and energy transformation by principal energy carriers: phonons, electrons, fluid particles, and photons. Heat is energy stored in temperature-dependent motion of particles including electrons, atomic nuclei, individual atoms, and molecules. Heat is transferred to and from matter by the principal energy carriers. The state of energy stored within matter, or transported by the carriers, is described by a combination of classical and quantum statistical mechanics. The energy is different made (converted) among various carriers. The heat transfer processes are governed by the rates at which various related physical phenomena occur, such as the rate of particle collisions in classical mechanics. These various states and kinetics determine the heat transfer, i.e., the net rate of energy storage or transport. Governing these process from the atomic level to macroscale are the laws of thermodynamics, including conservation of energy.

The semiconductor Bloch equations describe the optical response of semiconductors excited by coherent classical light sources, such as lasers. They are based on a full quantum theory, and form a closed set of integro-differential equations for the quantum dynamics of microscopic polarization and charge carrier distribution. The SBEs are named after the structural analogy to the optical Bloch equations that describe the excitation dynamics in a two-level atom interacting with a classical electromagnetic field. As the major complication beyond the atomic approach, the SBEs must address the many-body interactions resulting from Coulomb force among charges and the coupling among lattice vibrations and electrons.

The Elliott formula describes analytically, or with few adjustable parameters such as the dephasing constant, the light absorption or emission spectra of solids. It was originally derived by Roger James Elliott to describe linear absorption based on properties of a single electron–hole pair. The analysis can be extended to a many-body investigation with full predictive powers when all parameters are computed microscopically using, e.g., the semiconductor Bloch equations or the semiconductor luminescence equations.

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.

<span class="mw-page-title-main">Semiconductor laser theory</span> Theory of laser diodes

Semiconductor lasers or laser diodes play an important part in our everyday lives by providing cheap and compact-size lasers. They consist of complex multi-layer structures requiring nanometer scale accuracy and an elaborate design. Their theoretical description is important not only from a fundamental point of view, but also in order to generate new and improved designs. It is common to all systems that the laser is an inverted carrier density system. The carrier inversion results in an electromagnetic polarization which drives an electric field . In most cases, the electric field is confined in a resonator, the properties of which are also important factors for laser performance.

Quantum-optical spectroscopy is a quantum-optical generalization of laser spectroscopy where matter is excited and probed with a sequence of laser pulses.

Optical gain is the most important requirement for the realization of a semiconductor laser because it describes the optical amplification in the semiconductor material. This optical gain is due to stimulated emission associated with light emission created by recombination of electrons and holes. While in other laser materials like in gas lasers or solid state lasers, the processes associated with optical gain are rather simple, in semiconductors this is a complex many-body problem of interacting photons, electrons, and holes. Accordingly, understanding these processes is a major objective as being a basic requirement for device optimization. This task can be solved by development of appropriate theoretical models to describe the semiconductor optical gain and by comparison of the predictions of these models with experimental results found.

The Wannier equation describes a quantum mechanical eigenvalue problem in solids where an electron in a conduction band and an electronic vacancy within a valence band attract each other via the Coulomb interaction. For one electron and one hole, this problem is analogous to the Schrödinger equation of the hydrogen atom; and the bound-state solutions are called excitons. When an exciton's radius extends over several unit cells, it is referred to as a Wannier exciton in contrast to Frenkel excitons whose size is comparable with the unit cell. An excited solid typically contains many electrons and holes; this modifies the Wannier equation considerably. The resulting generalized Wannier equation can be determined from the homogeneous part of the semiconductor Bloch equations or the semiconductor luminescence equations.

<span class="mw-page-title-main">Superradiant phase transition</span> Process in quantum optics

In quantum optics, a superradiant phase transition is a phase transition that occurs in a collection of fluorescent emitters, between a state containing few electromagnetic excitations and a superradiant state with many electromagnetic excitations trapped inside the emitters. The superradiant state is made thermodynamically favorable by having strong, coherent interactions between the emitters.

References

  1. Kira, M.; Jahnke, F.; Koch, S.; Berger, J.; Wick, D.; Nelson, T.; Khitrova, G.; Gibbs, H. (1997). "Quantum Theory of Nonlinear Semiconductor Microcavity Luminescence Explaining "Boser" Experiments". Physical Review Letters79 (25): 5170–5173. doi:10.1103/PhysRevLett.79.5170
  2. 1 2 3 4 5 6 7 Kira, M.; Koch, S. W. (2011). Semiconductor Quantum Optics. Cambridge University Press. ISBN   978-0521875097.
  3. Li, Jianzhong (2007). "Laser cooling of semiconductor quantum wells: Theoretical framework and strategy for deep optical refrigeration by luminescence upconversion". Physical Review B75 (15). doi:10.1103/PhysRevB.75.155315
  4. 1 2 Berstermann, T.; Auer, T.; Kurtze, H.; Schwab, M.; Yakovlev, D.; Bayer, M.; Wiersig, J.; Gies, C.; Jahnke, F.; Reuter, D.; Wieck, A. (2007). "Systematic study of carrier correlations in the electron–hole recombination dynamics of quantum dots". Physical Review B76 (16). doi:10.1103/PhysRevB.76.165318
  5. Shuvayev, V.; Kuskovsky, I.; Deych, L.; Gu, Y.; Gong, Y.; Neumark, G.; Tamargo, M.; Lisyansky, A. (2009). "Dynamics of the radiative recombination in cylindrical nanostructures with type-II band alignment". Physical Review B79 (11). doi:10.1103/PhysRevB.79.115307
  6. 1 2 Kira, M.; Koch, S.W. (2006). "Many-body correlations and excitonic effects in semiconductor spectroscopy". Progress in Quantum Electronics30 (5): 155–296. doi:10.1016/j.pquantelec.2006.12.002
  7. Kira, M.; Jahnke, F.; Koch, S. (1998). "Microscopic Theory of Excitonic Signatures in Semiconductor Photoluminescence". Physical Review Letters81 (15): 3263–3266. doi:10.1103/PhysRevLett.81.3263
  8. Chatterjee, S.; Ell, C.; Mosor, S.; Khitrova, G.; Gibbs, H.; Hoyer, W.; Kira, M.; Koch, S.; Prineas, J.; Stolz, H. (2004). "Excitonic Photoluminescence in Semiconductor Quantum Wells: Plasma versus Excitons". Physical Review Letters92 (6). doi:10.1103/PhysRevLett.92.067402
  9. Schwab, M.; Kurtze, H.; Auer, T.; Berstermann, T.; Bayer, M.; Wiersig, J.; Baer, N.; Gies, C.; Jahnke, F.; Reithmaier, J.; Forchel, A.; Benyoucef, M.; Michler, P. (2006). "Radiative emission dynamics of quantum dots in a single cavity micropillar". Physical Review B74 (4). doi:10.1103/PhysRevB.74.045323
  10. Hader, J.; Moloney, J. V.; Koch, S. W. (2006). "Influence of internal fields on gain and spontaneous emission in InGaN quantum wells". Applied Physics Letters89 (17): 171120. doi:10.1063/1.2372443
  11. Hader, J.; Hardesty, G.; Wang, T.; Yarborough, M. J.; Kaneda, Y.; Moloney, J. V.; Kunert, B.; Stolz, W. et al. (2010). "Predictive Microscopic Modeling of VECSELs". IEEE J. Quantum Electron.46: 810. doi:10.1109/JQE.2009.2035714
  12. Rubel, O.; Baranovskii, S. D.; Hantke, K.; Heber, J. D.; Koch, J.; Thomas, P. V.; Marshall, J. M.; Stolz, W. et al. (2005). "On the theoretical description of luminescence in disordered quantum structures". J. Optoelectron. Adv. M.7 (1): 115.
  13. Imhof, S.; Bückers, C.; Thränhardt, A.; Hader, J.; Moloney, J. V.; Koch, S. W. (2008). "Microscopic theory of the optical properties of Ga(AsBi)/GaAs quantum wells". Semicond. Sci. Technol.23 (12): 125009.
  14. Feldtmann, T.; Schneebeli, L.; Kira, M.; Koch, S. (2006). "Quantum theory of light emission from a semiconductor quantum dot". Physical Review B73 (15). doi:10.1103/PhysRevB.73.155319
  15. Baer, N.; Gies, C.; Wiersig, J.; Jahnke, F. (2006). "Luminescence of a semiconductor quantum dot system". The European Physical Journal B50 (3): 411–418. doi:10.1140/epjb/e2006-00164-3
  16. Böttge, C. N.; Kira, M.; Koch, S. W. (2012). "Enhancement of the phonon-sideband luminescence in semiconductor microcavities". Physical Review B85 (9). doi:10.1103/PhysRevB.85.094301
  17. Gies, Christopher; Wiersig, Jan; Jahnke, Frank (2008). "Output Characteristics of Pulsed and Continuous-Wave-Excited Quantum-Dot Microcavity Lasers". Physical Review Letters101 (6). doi:10.1103/PhysRevLett.101.067401
  18. Aßmann, M.; Veit, F.; Bayer, M.; Gies, C.; Jahnke, F.; Reitzenstein, S.; Höfling, S.; Worschech, L. et al. (2010). "Ultrafast tracking of second-order photon correlations in the emission of quantum-dot microresonator lasers". Physical Review B81 (16). doi:10.1103/PhysRevB.81.165314

Further reading