Fact-checked by Grok 2 weeks ago

Electron hole

In semiconductor physics, an , often simply called a , is the absence of an from the valence band, which behaves as a positively charged capable of conducting . This vacancy arises when thermal energy, light, or other excitations promote an electron from the valence band to the conduction band, leaving behind a mobile positive charge that effectively moves as adjacent electrons shift to fill the void. Holes possess an effective mass, typically larger than that of electrons in the same material (for example, about 0.39 times the mass in ), and they contribute to current flow in the opposite direction to electrons due to their positive charge assignment of +e. Holes play a crucial role in the electrical properties of semiconductors, particularly in p-type materials where acceptor impurities, such as in , create an abundance of holes as majority carriers by accepting s from the valence band. In contrast, holes are minority carriers in n-type semiconductors doped with donors like , where s dominate conduction. The generation of electron-hole pairs is governed by the material's bandgap —for instance, approximately 1.1 eV in at —while recombination occurs when an fills a hole, often emitting a with matching the bandgap. This dynamic balance maintains , with intrinsic carrier concentrations around 10¹⁰ cm⁻³ in at 300 K. The concept of holes is fundamental to understanding and designing devices, including diodes, transistors, and solar cells, where controlled generation, movement, and recombination of holes enable functions like , , and photovoltaic conversion. In an applied , holes drift toward the negative electrode, enhancing conductivity and forming the basis for p-n junctions that underpin modern electronics. Their nature, distinct from true particles like positrons, arises from collective interactions in the crystal lattice, making holes indispensable for modeling charge transport in .

Basic Concepts

Definition

In , an electron hole (often simply called a hole) is defined as the absence of an from the valence band in a crystal lattice, effectively behaving as a positively charged carrier of . This absence allows adjacent electrons to move into the vacant site, resulting in the hole appearing to propagate in the opposite direction to the electron flow. The term "electron hole" originated in the 1930s, coined by solid-state physicists including and Yakov Frenkel during the development of band theory to describe charge transport in nearly filled electron bands. Peierls, in particular, introduced the idea in his 1930 analysis of electron transport through solids, recognizing that vacancies near the top of a filled band could be treated as positively charged entities with dynamics analogous to electrons. Frenkel further advanced the concept in the context of conduction mechanisms in ionic crystals around the same period. Unlike particles such as or positrons, holes are quasiparticles that emerge from collective many-body interactions within the crystal lattice, lacking independent existence outside the . They represent an effective description for simplifying the complex quantum behavior of electron-deficient states in periodic potentials. A key definitional relation in is the product of the electron concentration n and the hole concentration p, which equals the square of the intrinsic carrier concentration n_i: n \cdot p = n_i^2 This arises from in carrier statistics and holds for intrinsic and extrinsic semiconductors alike, providing a on carrier densities.

Simplified Analogy

In semiconductors, an electron hole can be intuitively understood through the of an empty seat in a crowded . Imagine a fully occupied row of seats where each seat represents an available state for an in the valence band. If one () leaves their seat to move to a higher , an empty seat () is left behind. As adjacent people shift to fill the vacancy—each moving into the empty spot—the apparent motion of the empty seat propagates in the opposite direction along the row. For example, if the original vacancy is in the middle and people shift rightward, the empty seat effectively "moves" leftward toward the end of the row, mimicking how a appears to carry positive charge as surrounding electrons collectively shift. A complementary analogy is that of a rising in a , such as in . The itself is not a solid entity but a void surrounded by moving molecules (); as the molecules flow around and past it, the displaces upward due to , opposite to the net flow. Similarly, in a , the 's "motion" results from the of filling the vacancy, giving the hole an effective positive charge and despite being an absence rather than a particle. This illustrates the emergent nature of hole propagation in a dense electron environment. These analogies provide a classical, intuitive to the concept of holes as absences in the valence band but have limitations: they simplify the complex many-body interactions among electrons and overlook quantum mechanical features, such as the hole's effective mass, which arises from band structure and determines its response to fields in ways not captured by macroscopic voids.

Physical Description

Quantum Mechanical View

In the quantum mechanical framework of band theory, electron holes emerge in semiconductors when an electron is thermally or optically excited from the valence band to the conduction band, creating a vacancy in the otherwise filled valence band at absolute zero temperature. This vacancy, or hole, acts as a whose properties arise from the of the remaining electrons in the valence band. The valence band consists of Bloch states occupied by electrons, and the excitation process leaves the system in a state equivalent to the removal of one electron from a specific Bloch state near the band maximum. Quantum mechanically, the hole is described as the absence of an electron in a particular Bloch wavefunction within the periodic crystal potential. The single-particle Bloch wavefunctions are of the form \psi_{n\mathbf{k}}(\mathbf{r}) = e^{i\mathbf{k}\cdot\mathbf{r}} u_{n\mathbf{k}}(\mathbf{r}), where n labels the band, \mathbf{k} is the crystal wavevector, and u_{n\mathbf{k}}(\mathbf{r}) is a periodic function with the lattice periodicity. For a missing electron in state \psi_{n\mathbf{k}} with spin s, the hole state is characterized by crystal momentum \mathbf{k}_h = -\mathbf{k} and spin s_h = -s, reflecting the time-reversal symmetry of the Schrödinger equation, which ensures \epsilon(\mathbf{k}, s) = \epsilon(-\mathbf{k}, -s) and group velocity \mathbf{v}(-\mathbf{k}, -s) = -\mathbf{v}(\mathbf{k}, s). This time-reversed description arises because the hole propagation corresponds to the reversed motion of the absent electron, transforming the many-body wavefunction of the filled band by removing the electron component. The detailed quantum picture reveals that the hole is not simply a negative-mass electron but the effective absence of one, leading to opposite charge and mass signs relative to valence electrons. Near the valence band maximum (typically at a high-symmetry point like \Gamma), valence electrons exhibit negative effective mass due to upward curvature in the dispersion (negative second derivative of energy with respect to \mathbf{k}). Removing such an electron inverts this behavior: the hole acquires a positive effective charge +e (opposite to the electron's -e) and positive effective mass, as the surrounding electrons respond collectively to mimic a positively charged particle moving with positive inertia. This reinterpretation avoids unphysical negative-mass dynamics while preserving the and Fermi-Dirac statistics for the . The dispersion relation for holes near the valence band maximum is derived using the \mathbf{k}\cdot\mathbf{p} perturbation theory, which expands the band structure around the extremum by treating the crystal momentum \mathbf{k} as a perturbation on the atomic-like states at \mathbf{k}=0. Start with the Schrödinger equation for the periodic potential: \left[ \frac{\mathbf{p}^2}{2m_0} + V(\mathbf{r}) \right] \psi_{n\mathbf{k}}(\mathbf{r}) = E_{n\mathbf{k}} \psi_{n\mathbf{k}}(\mathbf{r}), where \mathbf{p} = -i\hbar \nabla and m_0 is the free-electron mass. Substituting the Bloch form \psi_{n\mathbf{k}}(\mathbf{r}) = e^{i\mathbf{k}\cdot\mathbf{r}} u_{n\mathbf{k}}(\mathbf{r}) yields the equation for the periodic part: \left[ E - \frac{\hbar^2 k^2}{2m_0} + \frac{\hbar}{m_0} \mathbf{k} \cdot \mathbf{p} + V(\mathbf{r}) \right] u_{n\mathbf{k}}(\mathbf{r}) = 0. Here, H_0 = \mathbf{p}^2/(2m_0) + V(\mathbf{r}) gives the unperturbed bands at \mathbf{k}=0, with eigenstates |n\rangle and energies E_n(0). Treating H' = (\hbar/m_0) \mathbf{k} \cdot \mathbf{p} as the perturbation (first-order \mathbf{k}\cdot\mathbf{p} term), the energy correction for band v (valence) is E_v(\mathbf{k}) \approx E_v(0) + \frac{\hbar^2 k^2}{2m_0} + \sum_{n \neq v} \frac{|\langle n | (\hbar/m_0) \mathbf{k} \cdot \mathbf{p} | v \rangle|^2}{E_v(0) - E_n(0)}, where the sum is over other bands (e.g., conduction bands c). For small \mathbf{k}, the quadratic term dominates near the maximum, yielding a parabolic dispersion. For holes, energy is referenced from the valence maximum E_v(0), and the electron dispersion in the valence band has negative curvature; inverting for the hole gives E_h(\mathbf{k}) \approx \frac{\hbar^2 k^2}{2 m_h^*}, with positive hole effective mass m_h^* > 0 determined by the negative of the valence electron effective mass, where \frac{1}{m_h^*} = -\left( \frac{1}{m_0} + \frac{2}{m_0^2} \sum_c \frac{|\langle c | \mathbf{p} | v \rangle \cdot \hat{k}|^2}{E_v - E_c} \right), and the negative denominator (since E_v < E_c) ensures positive m_h^*. This approximation holds for low energies, capturing the quasiparticle's motion in the inverted band.

Charge and Effective Mass

In semiconductors, an electron hole carries an effective charge of +e, where e is the magnitude of the elementary charge, opposite to the electron's charge of -e. This positive charge arises from the absence of an electron in the valence band, with the surrounding lattice ions and electrons providing a collective response that manifests the hole as a positively charged quasiparticle. The effective mass of a hole, denoted m_h^*, is a positive scalar or tensor quantity derived from the inverse curvature of the valence band energy dispersion E(\mathbf{k}) near the band maximum, typically given by m_h^* = \hbar^2 / \left( \frac{\partial^2 E}{\partial k^2} \right). Due to the generally flatter curvature of valence bands compared to conduction bands, hole effective masses are often heavier than those of electrons, influencing their inertial response in the crystal lattice. For example, in silicon, the density-of-states effective mass for holes is 0.81 m_e , where m_e is the free-electron mass. In direct-gap semiconductors such as gallium arsenide (GaAs), the valence band structure exhibits degeneracy at the \Gamma point, leading to anisotropic splitting into heavy-hole and light-hole subbands due to spin-orbit coupling and crystal symmetry. The heavy-hole effective mass m_{hh}^* is larger than the light-hole effective mass m_{lh}^*, with representative values of m_{hh}^* \approx 0.51 m_e and m_{lh}^* \approx 0.082 m_e in GaAs; these differences arise from the distinct band curvatures along different crystallographic directions. The dynamics of holes are described by the semiclassical equation of motion, which for a uniform electric field \mathbf{E} takes the form m_h^* \frac{d\mathbf{v}}{dt} = +e \mathbf{E}, contrasting with the electron equation m_e^* \frac{d\mathbf{v}}{dt} = -e \mathbf{E}; here, m_h^* is the effective mass tensor incorporating crystal anisotropy, ensuring the hole accelerates in the direction of the field as a positive charge carrier.

Behavior in Materials

Generation and Recombination

Electron-hole pairs in semiconductors are generated through several primary mechanisms, each requiring sufficient energy to overcome the material's bandgap energy E_g, the minimum energy difference between the valence and conduction bands. Thermal generation, or intrinsic excitation, occurs when thermal energy at finite temperatures promotes electrons from the valence band to the conduction band, creating free electrons and holes; however, at room temperature (where kT \approx 0.026 eV), this process yields low carrier densities in wide-bandgap materials like silicon (E_g = 1.12 eV), with an intrinsic carrier concentration n_i \approx 10^{10} cm^{-3}. Optical generation involves the absorption of photons with energy h\nu \geq E_g, which excites electrons across the bandgap, producing electron-hole pairs; this is the basis for photoconductivity and requires photon wavelengths shorter than \lambda_c = hc / E_g (e.g., \lambda_c \approx 1.1 \mum for silicon). Doping-induced generation, or extrinsic carrier creation, arises from the ionization of dopant atoms: donor impurities (e.g., phosphorus in silicon) release conduction electrons, while acceptor impurities (e.g., boron) create holes in the valence band, significantly increasing carrier concentrations (up to $10^{17} cm^{-3}) without direct pair formation but effectively altering the free carrier populations. Recombination is the annihilation of electron-hole pairs, releasing energy equal to E_g or less, and occurs via radiative or non-radiative pathways. Radiative recombination involves direct band-to-band transitions where an electron recombines with a hole, emitting a photon with energy approximately E_g; this process is efficient in direct-bandgap semiconductors like gallium arsenide but negligible in indirect-bandgap materials due to momentum conservation requirements. Non-radiative recombination dissipates energy through lattice vibrations (phonons) or defect-assisted mechanisms, such as Shockley-Read-Hall (SRH) processes where traps capture carriers before they recombine; Auger recombination, another non-radiative type, transfers energy to a third carrier rather than emitting light. The carrier recombination lifetime \tau, which quantifies the average time before recombination, is often expressed as \tau = \frac{1}{A + B n} for minority carriers in n-type material, where A is the non-radiative recombination coefficient (related to defect density) and B n accounts for the radiative bimolecular rate proportional to electron density n. In indirect semiconductors like silicon, non-radiative recombination dominates because the momentum mismatch between band-edge electrons and holes prohibits efficient direct transitions, necessitating phonon or defect mediation, resulting in lifetimes typically around 1 \mus. At thermal equilibrium, generation and recombination rates balance according to detailed balance principles from statistical mechanics, leading to the mass-action law n p = n_i^2, where n and p are the electron and hole concentrations, and n_i is the intrinsic carrier concentration. This law derives from the Fermi-Dirac distribution of carriers: the electron density n = N_c \exp\left(\frac{E_F - E_c}{kT}\right) and hole density p = N_v \exp\left(\frac{E_v - E_F}{kT}\right), where N_c and N_v are effective densities of states, E_F is the Fermi level, E_c and E_v are band edges, and multiplying yields n p = N_c N_v \exp\left(-\frac{E_g}{kT}\right) = n_i^2, independent of doping as long as equilibrium holds. For silicon at 300 K, n_i^2 \approx 10^{20} cm^{-6}, ensuring that excess carriers generated out of equilibrium (e.g., by light) will recombine until this product is restored.

Mobility and Transport

The mobility of an electron hole, denoted as \mu_h, quantifies the average drift velocity of holes per unit electric field strength and is given by \mu_h = \frac{e \tau}{m_h^*}, where e is the elementary charge, \tau is the momentum relaxation time, and m_h^* is the effective mass of the hole. This relation arises from the balance between the accelerating force from the electric field and the decelerating effect of scattering events that randomize hole momentum over the relaxation time \tau. In silicon at room temperature (300 K), the hole mobility is approximately 470 cm²/V·s for low doping concentrations, significantly lower than electron mobility due to the heavier effective mass of holes. Hole transport is limited primarily by three scattering mechanisms: phonon scattering from lattice vibrations, ionized impurity scattering from charged dopants, and Coulomb scattering from interactions with other charged carriers or defects. Phonon scattering dominates in high-purity semiconductors at elevated temperatures, while impurity and Coulomb scattering prevail in doped materials at lower temperatures. The temperature dependence of hole mobility reflects these mechanisms; for phonon-limited transport, \mu_h \propto T^{-3/2}, as increased thermal vibrations enhance scattering rates. In contrast, ionized impurity scattering yields \mu_h \propto T^{3/2}, since higher temperatures increase the average hole velocity and screening, reducing the scattering cross-section. The effective mass m_h^* influences mobility inversely, as heavier holes respond less readily to fields, though detailed values are addressed in prior discussions of hole properties. Holes also exhibit diffusive transport driven by concentration gradients, with the diffusion coefficient D_h related to mobility via the Einstein relation D_h = \frac{kT}{e} \mu_h, where k is Boltzmann's constant and T is temperature. The total hole current density J_h combines drift and diffusion components in the drift-diffusion equation: \mathbf{J}_h = e \mu_h p \mathbf{E} - e D_h \nabla p where p is the hole concentration and \mathbf{E} is the electric field. In steady state, \nabla \cdot \mathbf{J}_h = 0, leading to solutions where drift balances diffusion, such as uniform current in a constant field or exponential decay of concentration gradients. The Hall effect provides a means to characterize hole transport; in p-type semiconductors, the Hall voltage exhibits a positive sign due to the positive effective charge of holes, enabling measurement of carrier type, density, and mobility via the Hall coefficient R_H = \frac{1}{e p}. This transverse voltage arises from the Lorentz force deflecting holes oppositely to electrons under a perpendicular magnetic field.

Applications in Technology

Semiconductor Devices

In p-type semiconductors, doping with acceptor impurities such as introduces holes as the majority charge carriers by creating energy levels near the valence band edge, allowing electrons to be thermally excited from the valence band, leaving behind positively charged holes. The Fermi level in p-type materials shifts closer to the valence band compared to intrinsic semiconductors, resulting in a high hole concentration (p ≈ N_a) and low electron concentration (n ≈ n_i^2 / N_a), where N_a is the acceptor density. The p-n junction forms at the interface between p-type and n-type semiconductors, where diffusion of majority carriers creates a depletion region depleted of free charge carriers due to the exposure of fixed ionized dopants. This region establishes a built-in potential barrier given by V_{bi} = \frac{kT}{e} \ln\left(\frac{N_a N_d}{n_i^2}\right), where k is Boltzmann's constant, T is temperature, e is the elementary charge, N_d is the donor density, and n_i is the intrinsic carrier concentration; this potential prevents further net carrier diffusion in equilibrium. In diode operation, applying a forward bias reduces the built-in potential, allowing majority carriers to overcome the barrier and inject minority carriers across the junction—specifically, holes from the p-side into the n-side. The resulting current-voltage characteristic follows the I = I_s \left( e^{eV / kT} - 1 \right), where I_s is the reverse saturation current dependent on material parameters and doping, and V is the applied voltage; this exponential increase in current arises primarily from the diffusion of these injected minority carriers. Holes play a central role in transistor operation, particularly in p-channel metal-oxide-semiconductor field-effect transistors (MOSFETs), where an applied gate voltage induces an inversion layer of holes in an n-type substrate, enabling hole conduction from source to drain under a drain-source bias. In bipolar junction transistors (BJTs), such as pnp types, holes are injected from the p-type emitter into the n-type base and collected at the p-type collector, amplifying the base current to produce a larger collector current. Complementary metal-oxide-semiconductor (CMOS) technology fundamentally relies on this complementary electron-hole transport, pairing n-channel MOSFETs (electron conduction) with p-channel MOSFETs (hole conduction) to achieve low-power switching with minimal static current. Hole mobility in silicon, typically about one-third that of electrons, influences the performance balance in such devices.

Optoelectronics and Photovoltaics

In optoelectronics, electron holes play a central role in light-emitting devices such as and , where radiative recombination with electrons in the conduction band generates photons. This process is particularly efficient in like , which allow momentum-conserving transitions without phonon involvement, enabling high internal quantum efficiencies approaching 100% under optimal conditions. In LEDs, injected electrons and holes recombine across the p-n junction, with the emitted photon's energy corresponding to the bandgap; GaAs-based red LEDs, for instance, were among the first to demonstrate practical visible emission in the 1960s due to this direct recombination pathway. Semiconductor lasers extend this principle by achieving population inversion, where stimulated emission dominates, amplifying light output; materials like enhance confinement of holes and electrons, supporting lasing thresholds as low as a few amps per centimeter. In photovoltaics, electron holes are essential for charge separation and collection in solar cells, particularly in p-type layers where photogenerated holes drift toward the contact under the built-in field of the p-n junction. Absorption of photons above the bandgap creates electron-hole pairs, but recombination—via radiative, non-radiative, or Auger processes—limits efficiency by reducing collectible carriers. The , derived from detailed balance principles, establishes a theoretical maximum efficiency of approximately 33% for single-junction cells under AM1.5 illumination, primarily due to unavoidable recombination losses and thermalization of high-energy photons. This limit highlights the role of holes in the voltage-dependent recombination current, where minimizing hole trapping in defects is key to approaching the ideal. Photodetectors rely on hole transport or trapping to convert absorbed light into measurable photocurrent, with generated holes contributing to the external signal in p-i-n or Schottky structures. The quantum efficiency (QE), a key performance metric, quantifies this by measuring the ratio of collected charge carriers to absorbed photons; specifically for hole-dominated response, \eta = \frac{\text{number of holes collected}}{\text{number of photons absorbed}} High hole mobility in materials like organic semiconductors or perovskites enhances QE by reducing transit time losses, enabling external QE values exceeding 80% in optimized devices across visible wavelengths. A notable advancement in photovoltaics involves perovskite solar cells, which emerged prominently in the 2010s and utilize hole transport layers (HTLs) such as spiro-OMeTAD or PTAA to selectively extract holes from the absorber, minimizing recombination at interfaces. These HTLs facilitate efficient hole collection while blocking electrons, contributing to certified power conversion efficiencies surpassing 25% by 2025, with recent polymer-based HTLs achieving 26.31% in inverted architectures through improved energy alignment and stability. This progress underscores the tunability of hole dynamics in hybrid perovskites, enabling tandem configurations that exceed the single-junction Shockley-Queisser limit.

Comparisons and Extensions

To Positron

The electron hole and the positron share superficial similarities as positively charged entities in their respective domains, both carrying a charge of +e and capable of interacting with electrons in processes that effectively neutralize them. The positron serves as the antiparticle to the electron in particle physics, predicted by Paul Dirac's relativistic quantum equation in 1928, which implied the existence of particles with the same mass but opposite charge to balance the quantum mechanical description of the electron. In contrast, the electron hole emerges from solid-state physics as a conceptual tool to describe the absence of an electron in a filled valence band, behaving as if it were a positively charged carrier with properties derived from the surrounding lattice. Key differences highlight their distinct natures: the positron is a fundamental lepton, an elementary particle with a rest mass m_e \approx 9.11 \times 10^{-31} kg, existing independently in vacuum and observable in cosmic rays or particle accelerators. The electron hole, however, is a composite quasiparticle, lacking independent existence outside a solid material; its effective mass m_h^* is typically a fraction of m_e and varies with the semiconductor (e.g., m_h^* \approx 0.49 m_e for heavy holes in silicon), reflecting the collective response of the crystal lattice rather than intrinsic particle properties. Unlike the positron, which can be isolated and propagated in free space, the hole's behavior is confined to the band structure of solids and ceases to exist if the material is dismantled. Their annihilation processes further underscore these contrasts. When a positron collides with an electron, they annihilate into two gamma-ray photons, each with energy 511 keV corresponding to $2m_e c^2, conserving energy, momentum, and charge in a fundamental interaction. In semiconductors, electron-hole recombination instead releases energy as phonons (lattice vibrations) in non-radiative processes or photons with energy approximately equal to the material's bandgap E_g (e.g., \sim1.1 eV in silicon), without producing high-energy gamma rays, as the process involves band-to-band transitions rather than particle-antiparticle destruction. Historically, the positron's theoretical foundation in Dirac's 1928 work preceded the hole's conceptualization, which arose in the 1930s amid the development of band theory by researchers like Alan Wilson, who formalized holes as effective carriers to explain conductivity in semiconductors. This temporal contrast reflects the shift from relativistic quantum field theory to quantum mechanics applied to many-body systems in condensed matter.

To Other Quasiparticles

Excitons represent a composite quasiparticle formed by the Coulomb attraction between an electron and a hole in semiconductors and insulators, resulting in a bound state with a characteristic binding energy. The binding energy for , which are delocalized over many lattice sites, is given by the hydrogenic formula E_b = \frac{\mu e^4}{2 \hbar^2 \epsilon^2}, where \mu is the reduced mass of the electron-hole pair, e is the elementary charge, \hbar is the reduced Planck's constant, and \epsilon is the dielectric constant of the material. In contrast, , which are more tightly bound and localized on single molecules, exhibit higher binding energies and are particularly stable in insulators due to weaker screening of the Coulomb interaction. Polarons extend the concept of a bare by incorporating the influence of the surrounding , where the hole becomes "dressed" with a cloud of phonons arising from electron-phonon coupling. This coupling distorts the ionic around the hole, leading to an enhanced effective mass m_p^* > m_h^*, where m_h^* is the bare hole effective mass, as the quasiparticle's motion is impeded by the accompanying . The formation was first theoretically described in the strong-coupling limit, highlighting how such quasiparticles dominate charge transport in polar materials. Unlike magnons, which are bosonic quasiparticles representing collective excitations, or phonons, which are bosonic vibrational modes of the , electron holes behave as fermionic charge carriers obeying the and contributing to electrical conduction. This fermionic distinguishes holes from these bosonic excitations, as holes maintain particle-like statistics in the many-body description of the . In modern contexts, the hole concept has been extended to topological quasiparticles in two-dimensional materials such as , where post-2010 discoveries revealed hole states exhibiting Dirac-like linear near the neutrality point, enabling unique topological properties like valley-dependent transport. These topological holes arise from the massless Dirac fermion description in 's , with experimental observations confirming their role in electron-hole puddles and dynamics.

References

  1. [1]
    Basics of Semiconductors - BYU Physics & Astronomy
    Hole: A missing electron in the valence band of a solid. Holes move in the opposite direction of electrons since they are formed as an electron moves to fill a ...Missing: definition | Show results with:definition
  2. [2]
    [PDF] Electrons and Holes in Semiconductors
    Of particular importance are the concepts of energy band, the two kinds of electrical charge carriers called electrons and holes, and how the carrier ...
  3. [3]
    The Feynman Lectures on Physics Vol. III Ch. 14: Semiconductors
    Electron hole pairs can also be produced by high-energy particles. When a fast-moving charged particle—for instance, a proton or a pion with an energy of tens ...
  4. [4]
    [PDF] Physics of the XXth century
    May 22, 2019 · 1928-1930 Bloch, Rudolf Peierls - quantum theory of electrical conductivity of metals. 1928. Peierls, Yakov Frenkel – 'hole' conduction. 1931.Missing: concept | Show results with:concept
  5. [5]
  6. [6]
    [PDF] SEMICONDUCTORS AND P-N JUNCTIONS
    The concept of a moving hole is analogous to that of a bubble in a liquid. ... break bonds and create electron-hole pairs, which can then wander off independently ...
  7. [7]
    Quasiparticle | Coherent Excitations, Electron-Hole Pairs & Interactions
    Quasiparticle, in physics, a disturbance, in a medium, that behaves as a particle and that may conveniently be regarded as one.Missing: authoritative | Show results with:authoritative
  8. [8]
    [PDF] Optical Physics of Quantum Wells - Stanford Electrical Engineering
    Both electrons and holes see lower energy in the "well" layer, hence the name (by analogy with a "potential well"). ... A free electron-hole pair is analogous to ...
  9. [9]
    [PDF] Electrons and Holes in Semiconductors - Pearl HiFi
    The hole, or deficit produced by removing an electron from the valence bond structure of a crystal, is the chief reason for existence of this book. Although the ...
  10. [10]
    [PDF] Solid State Theory I - TKM (KIT)
    Thus far we characterized holes by the crystal momentum and the spin of the absent electron k, s. It is more logical to say that the hole has a crystal ...
  11. [11]
    [PDF] PH575 Spring 2019 Lecture #10 Electrons, Holes; Effective mass ...
    Thus hole has momentum -k. • If electron with negative charge and negative mass in state k is missing, the remaining electrons can be described as a hole.
  12. [12]
    NSM Archive - Band structure and carrier concentration of Silicon (Si)
    Effective mass of density of states, mc = 0.36m ; There are 6 equivalent valleys in the conduction band. ; mcc= 0.26m ; Holes: ; Heavy, mh = 0.49m ...Missing: m0 | Show results with:m0
  13. [13]
    [PDF] 6.730 “Real” Semiconductor Project GaAs Part 3
    Apr 18, 2001 · Table 5: Valence band effective mass ... Light Hole Band ml = .082 m0 ml = .33m0. Split Off Band ms0=0.154 m0 ms0=0.15 m0. Heavy Hole Band mh = ...
  14. [14]
    [PDF] Solid State Devices Lecture 9 Recombination Processes and Rates
    In a typical semiconductor device there are 1e17 to 1e20 electrons in the conduction band. ... one electron-hole pair. No change in nT and pT. Klimeck – ECE606 ...
  15. [15]
    [PDF] Motion and Recombination of Electrons and Holes
    The electron–hole recombination rate is proportional to. (= ): τ is the recombination lifetime and ranges from nanoseconds to milliseconds for Si, depending on ...Missing: limitations | Show results with:limitations
  16. [16]
    [PDF] Lecture 14: Semiconductors
    A hole contributes a positive amount to the energy of the electron gas, since the electron that should have been in the hole is missing. If a electron in.
  17. [17]
    [PDF] Hall Effect Measurement Handbook - Quantum Design
    Carrier type: electrons if RH is negative; holes if RH is positive. Measurement of. Hall voltage in Hall bar samples. The Hall voltage is typically measured ...
  18. [18]
    The Hall Effect | NIST - National Institute of Standards and Technology
    Apr 15, 2010 · ... Hall voltage is negative for n-type semiconductors and positive for p-type semiconductors. The sheet resistance RS of the semiconductor can ...
  19. [19]
    [PDF] PN and Metal–Semiconductor Junctions
    Let's divide the PN junction into three regions—the neutral regions at x > xP and x < –xN, and the depletion layer or depletion region in between, where p = n = ...
  20. [20]
    [PDF] Part 3 Solid State Electronic Devices - EE3310 Class notes Transistors
    Nov 27, 2002 · NOTE THAT MOSFET IS A SPECIAL VERSION OF MISFET. CMOS – Complimentary MOS FETs. This device consists of two MOSFETs – one p-channel and one n- ...Missing: complementary | Show results with:complementary
  21. [21]
    Direct and Indirect Band Gap Semiconductors - DoITPoMS
    The recombination process is much more efficient for a direct band gap semiconductor than for an indirect band gap semiconductor, where the process must be ...
  22. [22]
    Laser Diode and LED Physics - Newport
    When the electron returns from the conduction band to this empty state in the valence band, a process known as electron-hole recombination occurs. This can ...
  23. [23]
    Detailed Balance Limit of Efficiency of p‐n Junction Solar Cells
    A preliminary report of the analysis of this paper was presented at the Detroit meeting of the American Physical Society: H. J.. Queisser. and. W. Shockley. ,.
  24. [24]
    An Introduction to Quantum Efficiency | External and Internal - Ossila
    In solar cells, the quantum efficiency is the ratio of the number of charge carriers (electron-hole pair) collected to the number of incident photons. This is ...
  25. [25]
    External quantum efficiency versus charge carriers mobility in ...
    Jul 18, 2007 · The paper studies the role of electrons and holes mobility in determining the external quantum efficiency (EQE) in photodetectors based on a ...
  26. [26]
    Green-Solvent-Processable Polymer Hole Transport Material for ...
    Green-Solvent-Processable Polymer Hole Transport Material for Achieving 26.31% Efficiency in Inverted Perovskite Solar Cells.
  27. [27]
    Positron - an overview | ScienceDirect Topics
    A positron is the antiparticle of an electron. It has all the properties of an electron except for the polarity of the electrical charge, which is positive.
  28. [28]
    Conduction in Semiconductors - PVEducation
    This empty space is commonly called a "hole", and is similar to an electron, but with a positive charge.
  29. [29]
    Antimatter | CERN
    Dirac interpreted the equation to mean that for every particle there exists a corresponding antiparticle, exactly matching the particle but with opposite charge ...
  30. [30]
    Electron band semiconductor devices - Book chapter - IOPscience
    In the early 1930s Alan Wilson [21] used the evolving quantum theory of solids to propose the idea of electronic semiconductors, and associated their many ...
  31. [31]
    What is Positron - Properties - Nuclear Power
    The positrons are positively charged (+1e), almost massless particles. Their rest mass equals 9.109 × 10−31 kg (510.998 keV/c2) (approximately 1/1836 that of ...
  32. [32]
    Possible Electron-positron Annihilation Products? - Physics Van
    Jun 23, 2010 · The usual products of electron-positron annihilation are photons; sometimes two and sometimes three.
  33. [33]
    Types of Recombination - PVEducation
    In radiative recombination, an electron from the conduction band directly combines with a hole in the valence band and releases a photon; and. The emitted ...
  34. [34]
    [PDF] Solid State Physics 1925-33: Opportunities Missed and ...
    Feb 9, 2005 · The mechanism by which free electrons could be produced in an otherwise empty energy band, or free holes in an otherwise full energy band, could ...
  35. [35]
    Spin dynamics and relaxation in graphene dictated by electron-hole ...
    Feb 15, 2016 · Electron-hole puddles are real-space fluctuations of the chemical potential, induced by the underlying substrate, which locally shift the Dirac ...