Fact-checked by Grok 2 weeks ago

Electron mobility

Electron mobility, denoted as \mu_n, is a fundamental parameter in that quantifies the ease with which electrons drift through a or under the influence of an applied , defined as the ratio of the drift v_d to the strength E, given by \mu_n = v_d / E. The typical units for electron mobility are cm²/V·s, reflecting its measurement of velocity per unit field. In semiconductors, electron mobility plays a critical role in determining electrical and device performance, as it directly influences the J_n = n q \mu_n E, where n is the concentration and q is the . For instance, in at , mobility is approximately 1350 cm²/V·s, significantly higher than hole mobility at 480 cm²/V·s, enabling faster electron transport in n-type materials. High electron mobility is essential for applications in high-speed electronics, such as field-effect transistors and solar cells, where materials like exhibit values exceeding 8000 cm²/V·s due to reduced . Electron mobility is primarily limited by scattering mechanisms, including interactions with lattice phonons, ionized impurities, and defects, which reduce the mean free time between collisions and thus lower \mu_n. dependence is notable, with mobility decreasing at higher temperatures due to increased , though this effect is less pronounced in heavily doped semiconductors. In two-dimensional materials like , mobilities can reach over 200,000 cm²/V·s at low temperatures, highlighting opportunities for engineering and doping to enhance performance in next-generation devices.

Fundamentals

Definition and Units

Electron mobility, denoted as \mu_e or simply \mu, quantifies the ease with which electrons respond to an applied in a solid material, such as a metal or . It serves as the proportionality constant between the average v_d of electrons and the electric field strength E, formally defined by the relation v_d = \mu_e E. This definition captures the directed motion of charge carriers superimposed on their random thermal motion, providing a key parameter for understanding electrical transport properties. The concept originated in early 20th-century as part of the , a classical theory of conduction in metals proposed by Paul Drude in 1900. In this model, s are treated as a free gas interacting with the ionic lattice, where mobility arises from the balance between accelerating electric forces and collisional scattering. The term "mobility" derives from its representation of the electrons' freedom of movement, distinguishing it from fixed lattice ions. A parallel quantity, hole mobility \mu_h, applies to positive charge carriers (holes) in semiconductors, which behave as if moving oppositely to s; in materials like , \mu_h is typically about one-third of \mu_e due to differences in effective mass and scattering. In the (), electron mobility has dimensions of area per volt-second, yielding m²/(V·s), which emerges dimensionally from (m/s) divided by (V/m). For practical applications in physics, the cgs-derived unit cm²/(V·s) is prevalent, where values for electrons in range up to approximately 1400 cm²/(V·s) at , and 1 m²/(V·s) equals 10⁴ cm²/(V·s). This unit choice facilitates comparison with experimental data on carrier transport.

Drift Velocity and Derivation

In the presence of an applied electric field \mathbf{E}, charge carriers such as electrons in a semiconductor or metal experience a net directed motion superimposed on their random thermal velocities. The thermal velocity arises from the random kinetic energy due to temperature, typically on the order of $10^7 cm/s for electrons at room temperature, resulting in no net current without a field. In contrast, the drift velocity \mathbf{v}_d is the average velocity acquired by the ensemble of electrons in response to the field, directed opposite to \mathbf{E} for electrons due to their negative charge, and is much smaller in magnitude under typical conditions. This steady-state drift arises from a balance between the accelerating electric force and the decelerating effect of events with vibrations, impurities, or other obstacles. The electric force on an is -[e](/page/E!) \mathbf{E} (where [e](/page/E!) > 0 is the magnitude), imparting an \mathbf{a} = - ([e](/page/E!) / m^*) \mathbf{E}, with m^* as the effective mass accounting for band structure effects in the solid. introduces a frictional drag, approximated in the relaxation time model as a momentum loss term proportional to the velocity, with relaxation time \tau representing the average time between collisions. In , this balance yields the drift velocity \mathbf{v}_d = - ([e](/page/E!) \tau / m^*) \mathbf{E}. To derive this formally, start with the semiclassical equation of motion for the average electron velocity \mathbf{v}: m^* \frac{d\mathbf{v}}{dt} = -e \mathbf{E} - \frac{m^* \mathbf{v}}{\tau}, where the second term on the right models the exponential decay of momentum toward zero after each scattering event, assuming carriers lose their drift momentum upon collision. For steady-state conditions, d\mathbf{v}/dt = 0, so -e \mathbf{E} = \frac{m^* \mathbf{v}_d}{\tau} \implies \mathbf{v}_d = -\frac{e \tau}{m^*} \mathbf{E}. The electron mobility \mu_n is defined as the proportionality constant between |\mathbf{v}_d| and E, giving \mu_n = e \tau / m^*, such that \mathbf{v}_d = -\mu_n \mathbf{E}. This derivation relies on the relaxation time approximation from the Boltzmann transport equation, where \tau is taken as constant. The model assumes low where the linear response v_d \propto E holds, an isotropic material with uniform , and neglect of or other external influences that could alter trajectories.

Transport Relations

Relation to Current Density

The drift current density arises from the collective motion of electrons under an applied electric field, directly linking microscopic electron mobility to macroscopic charge transport in materials such as semiconductors and metals. When an electric field \mathbf{E} is applied, electrons experience a drift velocity \mathbf{v}_d = -\mu \mathbf{E}, where \mu is the electron mobility (positive by convention), leading to a net current due to the imbalance in electron flow. The resulting drift current density for electrons is expressed as \mathbf{J}_d = n e \mu \mathbf{E}, where n is the electron number density and e is the elementary charge ($1.602 \times 10^{-19} C). This formula quantifies how efficiently electrons contribute to current flow, with higher mobility yielding greater current for a given field and carrier density. From experimental measurements of current density under controlled conditions, electron mobility can be determined by rearranging the relation: \mu = \frac{J_d}{n e E}, assuming isotropic conditions and known n and E. This inversion is particularly useful in device characterization, where applied fields generate measurable currents to infer material properties. In the context of Ohm's law, \mathbf{J} = \sigma \mathbf{E} (with \sigma as electrical conductivity), the drift current forms the primary contribution in low-field regimes for extrinsic semiconductors, bridging microscopic carrier dynamics to the material's overall resistive behavior. For anisotropic materials, such as certain or layered semiconductors where electron transport varies by direction due to band structure asymmetry, the scalar mobility is replaced by a second-rank mobility tensor \mu_{ij}. The drift current density then takes the vector form J_{d,i} = n e \sum_j \mu_{ij} E_j, capturing direction-dependent responses; for example, in hexagonal like derivatives, \mu_{xx} may differ significantly from \mu_{zz}. This tensorial description is essential for modeling transport in non-cubic lattices, ensuring accurate predictions of current flow in devices exploiting .

Relation to Electrical Conductivity

In the classical Drude model of electrical conduction in metals, the electrical conductivity \sigma is directly related to electron mobility \mu through the formula \sigma = n e \mu, where n is the electron density and e is the elementary charge. This expression arises from the drift velocity of electrons under an applied electric field, where the average drift speed v_d = \mu E leads to a current density J = n e v_d = n e \mu E, and thus \sigma = J / E = n e \mu. The model, proposed by Paul Drude in 1900, treats conduction electrons as a classical gas subject to random scattering, with mobility \mu = e \tau / m incorporating the relaxation time \tau between collisions and the electron mass m. For semiconductors, the relation extends to account for both electrons and holes: the total conductivity is \sigma = e (n \mu_e + p \mu_h), where n and p are the electron and hole densities, respectively, and \mu_e and \mu_h are their respective mobilities. In intrinsic semiconductors, n = p = n_i, simplifying to \sigma = e n_i (\mu_e + \mu_h), while in extrinsic cases, one carrier type dominates, such as \sigma \approx e n \mu_e for n-type materials. This formulation highlights how mobility influences charge transport efficiency, with higher \mu values yielding greater \sigma and correspondingly lower resistivity \rho = 1/\sigma, enabling better electrical performance in devices. The dependence on effective mass m^* further explains variations across materials: in the quantum mechanical refinement, \mu \propto 1/m^*, as the effective mass accounts for lattice interactions in band structures, replacing the free electron mass in Drude's original expression./Solar_Basics/C._Semiconductors_and_Solar_Interactions/II._Conduction_in_Semiconductors/4%3A_Carrier_Drift_and_Mobility) Drude's classical approach was refined in 1928 by Arnold Sommerfeld using Fermi-Dirac statistics, which better describes the Pauli exclusion principle for degenerate electron gases in metals, while Felix Bloch's band theory incorporated periodic lattice potentials to validate the conductivity-mobility link quantum mechanically. These advancements preserved the core relation \sigma = n e \mu but provided a more accurate foundation for understanding material-specific behaviors.

Relation to Electron Diffusion

In semiconductors, the diffusion current arises from the random thermal motion of charge carriers, leading to a net flow from regions of higher concentration to lower concentration. For , this is expressed as J_{\text{diff}} = -e D \nabla n, where e is the , D is the diffusion coefficient, and n is the electron concentration. This component of current is distinct from but contributes to the total transport under non-uniform conditions. The Einstein relation connects the diffusion coefficient D to the electron mobility \mu through D = \frac{kT}{e} \mu, where k is Boltzmann's constant and T is the absolute temperature. This is derived by considering , where the net current is zero, balancing the drift current due to a built-in and the due to a concentration gradient. In , the carrier concentration follows n = n_0 \exp\left(-\frac{eU}{kT}\right), leading to \nabla n = -\frac{e}{kT} n \nabla U, and setting the total j = e \mu n (-\nabla U) - e D \nabla n = 0 yields the Einstein relation. Thermoelectric effects, such as the Seebeck effect, rely on this relation, as the S, which measures the voltage generated by a , incorporates the ratio D / \mu = kT / e to describe the -driven separation of carriers. In non-degenerate semiconductors, S for electrons is approximately S = -\frac{k}{e} \left( \frac{E_c - E_F}{kT} + A \right), where A accounts for , and the Einstein relation ensures consistency between thermal and in the underlying transport equations. The standard Einstein applies to non-degenerate semiconductors, where carrier statistics follow Maxwell-Boltzmann distributions. For degenerate cases, such as heavily doped materials where the enters the conduction , generalizations incorporate Fermi-Dirac integrals, modifying the to D = \frac{kT}{e} \mu \frac{\mathcal{F}_{1/2}(\eta)}{\mathcal{F}_{-1/2}(\eta)}, where \eta is the reduced and \mathcal{F}_j are Fermi-Dirac integrals.

Examples

In Traditional Semiconductors and Metals

In traditional semiconductors, electron mobility values provide a for carrier transport in established materials used in . For intrinsic at 300 K, the electron mobility is approximately 1400 cm²/(V·s), dominated by in lightly doped conditions. exhibits higher mobility, with μ_e ≈ 3900 cm²/(V·s) in its intrinsic form at room temperature. , a III-V compound , shows even greater values, around 8500–8800 cm²/(V·s) for intrinsic material, enabling faster device speeds compared to . In metals, electron mobility is generally much lower due to the high of conduction electrons (n ≈ 10^{22}–10^{23} cm^{-3}), despite high electrical . For , a prototypical , μ_e ≈ 43 cm²/(V·s) at , limited primarily by electron-electron and electron-phonon interactions. These mobility values are sensitive to material purity, where impurities introduce that reduces μ_e, and to , with causing a decrease as rises.
MaterialElectron Mobility (cm²/(V·s))Conditions
Silicon (Si)1400Intrinsic, 300 K
Germanium (Ge)3900Intrinsic, room T
GaAs8500–8800Intrinsic
Copper (Cu)43Bulk, room T

In High-Mobility and Emerging Materials

High electron mobility in emerging materials arises primarily from minimized scattering events, such as reduced phonon interactions in two-dimensional (2D) structures and defect-free interfaces achieved through encapsulation or precise engineering techniques. These materials surpass traditional semiconductors by enabling faster charge transport, which is crucial for next-generation devices requiring high-speed performance and mechanical flexibility. In , hexagonal boron (hBN) encapsulation has enabled record room-temperature electron mobilities exceeding 150,000 cm²/(V·s) at carrier densities around 10¹¹ cm⁻², limited mainly by acoustic . This advancement, reported in 2025, stems from proximity screening that suppresses charge inhomogeneities and , preserving the intrinsic ballistic transport properties of graphene. Such high mobilities position graphene for applications in high-electron-mobility transistors (HEMTs) operating at frequencies. Transition metal dichalcogenides like MoS₂ exhibit electron mobilities of approximately 100–500 cm²/(V·s), significantly enhanced by engineering that modulates structure and reduces intervalley . Studies from 2020–2025 demonstrate that uniaxial tensile in flexible substrates can boost on-state currents and mobilities up to 185 cm²/(V·s) in field-effect transistors, leveraging the material's direct bandgap for optoelectronic . These properties support scalable 2D HEMTs for low-power, high-frequency . Organic semiconductors, particularly 2D conjugated coordination polymers such as Cu₃BHT formed via covalent bonding of benzenehexathiol linkers, achieve exceptional mobilities approaching 2,000 cm²/(V·s) for hot carriers at , as measured by time-resolved in 2025 developments. This ultrafast transport, enabled by crystalline lattices with minimal torsional defects, outperforms conventional organic polymers and enables efficient hot-electron harvesting. These materials are ideal for , including wearable sensors and bendable displays, due to their solution-processability and mechanical compliance. GeSn alloys in thin films have demonstrated electron mobilities over 1,500 cm²/(V·s) at (300 ) in strain-relaxed compositions like Ge₀.₈₈Sn₀.₁₂, grown by , with values reaching 6,200 cm²/(V·s) at 50 . The incorporation of tin shifts the conduction band minimum to the Γ-valley, reducing effective mass and , which is vital for photodetectors and lasers in mid-IR applications. These alloys bridge compatibility with direct-bandgap tunability for photonic integrated circuits.

High-Field Behavior

Electric Field Dependence

In the low-electric-field regime, typically below approximately 10 kV/cm in , electron mobility remains constant and independent of the applied strength E, as the average v_d is linearly proportional to E via v_d = \mu E, with scattering rates dominated by conditions. At higher fields, non-linear emerges, and the effective mobility \mu(E) decreases with increasing E, transitioning from a constant value to a field-dependent form. This behavior arises because the linear relation breaks down when carrier energies exceed , leading to enhanced . The primary mechanism for this mobility reduction is the hot electron effect, where electrons accelerate in the , gaining far above the (often reaching effective temperatures of several thousand ). These hot electrons interact more strongly with the , particularly through optical emission, which has an threshold of around 40–60 meV in common semiconductors; this increases the rate and limits net acceleration, causing \mu(E) to decline. In extreme cases, the dependence can manifest logarithmically due to the tail of the , though the 1/E proportionality often dominates in the saturation approach. Empirical models capture this transition effectively. A common form is \mu(E) = \frac{\mu_0}{1 + \left( \frac{\mu_0 E}{v_\mathrm{sat}} \right)^\beta}, where \mu_0 is the low-field , v_\mathrm{sat} is the saturation (typically \sim 10^7 cm/s), and \beta is an empirical exponent ranging from 1 to 2, reflecting the sharpness of the transition. This model, derived from time-of-flight measurements, fits experimental data across temperatures and fields. Material-specific variations highlight the dependence's sensitivity. In gallium arsenide (GaAs), with \mu_0 \approx 8500 cm²/V·s, the characteristic field for mobility reduction (E_c \approx v_\mathrm{sat}/\mu_0 \sim 1 kV/cm) is lower than in silicon (Si), where \mu_0 \approx 1400 cm²/V·s yields E_c \sim 7 kV/cm, despite comparable v_\mathrm{sat}; thus, GaAs exhibits a stronger field dependence, with mobility dropping more rapidly at moderate fields due to its higher baseline mobility amplifying hot electron scattering onset. In GaAs, \beta \approx 2 often provides a better fit to account for multivalley transfer effects enhancing the non-linearity.

Velocity Saturation

In high electric fields, the drift velocity of electrons in semiconductors approaches a limiting value called the saturation velocity, v_{\text{sat}}, beyond which further increases in field strength do not accelerate the carriers proportionally. This phenomenon arises because energetic electrons ("hot electrons") frequently emit optical phonons to dissipate excess energy, capping the average velocity at around $10^7 cm/s in typical semiconductors. At these high fields, the conventional low-field mobility \mu no longer applies directly, as the drift velocity v_d becomes field-independent and equal to v_{\text{sat}}. Instead, an effective mobility \mu_{\text{eff}} = v_{\text{sat}} / E is used, where E is the , reflecting the sublinear relationship between and field in this regime. This effective mobility decreases inversely with E, impacting device performance in high-speed . The Gunn effect, discovered in the early 1960s, exemplifies consequences of high-field behavior involving velocity overshoot, where electron velocity initially exceeds expectations before saturating, leading to negative differential resistance in materials like GaAs due to intervalley scattering. Saturation velocities vary by material: approximately $10^7 cm/s in silicon and GaAs, but higher at about $2 \times 10^7 cm/s in indium phosphide (InP), influencing choices for high-frequency applications.

Scattering Mechanisms

Phonon and Lattice Scattering

Phonon scattering arises from the interaction between electrons and vibrations in , serving as a primary intrinsic limiter of electron mobility in semiconductors. These vibrations manifest as phonons—quantized modes of oscillation—that couple to electrons through electron-phonon interactions, disrupting carrier motion and reducing the . At elevated temperatures, the population of phonons increases according to the Bose-Einstein distribution, intensifying scattering events and thereby decreasing mobility. This mechanism is particularly relevant in high-purity materials where extrinsic scattering is minimized, and it establishes the fundamental temperature-dependent baseline for transport properties. Acoustic scattering, involving low-energy longitudinal and transverse modes, predominates at lower s and is mediated by the deformation potential, which quantifies the band-edge shift due to dilation or . This interaction is an elastic process, conserving energy while changing momentum, as the energy is much smaller than typical carrier energies. Within the Boltzmann transport equation framework, the limited by acoustic s scales as \mu_{ac} \propto T^{-3/2}, stemming from the linear increase in occupation number with and the energy dependence of the scattering rate. For instance, in materials like or , this leads to a characteristic decrease in as rises from cryogenic levels. Optical phonon scattering becomes increasingly dominant at higher temperatures, where the higher-energy optical modes (typically 10-50 meV) facilitate inelastic processes that absorb or emit discrete quanta. This occurs in both polar and non-polar forms: polar optical , prevalent in ionic semiconductors like GaAs, arises from the long-range Coulombic Fröhlich due to relative ionic displacements, while non-polar optical relies on short-range deformation potential similar to acoustic modes but involving optical frequencies. The dependence varies; polar optical often yields an exponential form for at low temperatures due to the activation over the energy barrier, transitioning to power-law behaviors such as \mu_{op} \propto T^{-1/2} or T^{-3/2} in the high-temperature , whereas non-polar optical more consistently follows \mu_{op} \propto T^{-1/2}. These processes limit more severely above , as the occupation factor n_q \approx kT / \hbar \omega_{op} grows, enhancing rates. The relaxation time \tau_{ph} for phonon scattering, derived from solutions to the Boltzmann transport equation, encapsulates these interactions; for acoustic phonons, \tau_{ph} \propto 1/(kT) in the equipartition regime, reflecting the proportionality of the scattering rate to the thermal phonon density. This temperature scaling directly underlies the observed mobility dependences, as mobility \mu \propto e \langle \tau \rangle / m^*, where \langle \tau \rangle averages over the carrier distribution. At 300 K, a typical reference temperature for room-temperature device operation, phonon scattering sets intrinsic mobilities on the order of $10^3 to $10^4 cm²/V·s in conventional semiconductors like Si or GaN, though values vary with material-specific coupling strengths and phonon spectra.

Ionized Impurity Scattering

Ionized impurity scattering arises from the electrostatic interaction between conduction s and charged dopant ions in semiconductors, becoming the dominant mobility-limiting mechanism in moderately to heavily doped materials at low s. Unlike , which increases with , ionized impurity scattering decreases as rises due to enhanced electron velocities and screening effects. This process is particularly relevant in n-type or p-type semiconductors where dopants are ionized, creating fixed positive or negative charges that deflect electrons via long-range forces. The is a Rutherford-like , modified by the semiconductor's constant ε, which reduces the potential strength compared to . The bare potential between an and an ionized of charge Ze is V(r) = \frac{Z e^2}{4 \pi \epsilon_0 \epsilon r}, where ε_0 is the , but this is screened by surrounding free carriers and lattice polarization. The Brooks-Herring model provides a widely used quantum-mechanical treatment of this , assuming a Debye-Hückel screened potential of the form V(r) = \frac{Z e^2}{4 \pi \epsilon_0 \epsilon r} \exp\left(-\frac{r}{\lambda_D}\right), where λ_D is the Debye screening length. This model derives the scattering rate using Born approximation, leading to an ionized impurity mobility μ_{ii} that scales as μ_{ii} \propto \frac{T^{3/2}}{N_i \ln(1 + \gamma)}, where N_i is the ionized impurity density, T is temperature, and γ involves the ratio of electron energy to screening energy; the logarithmic term accounts for small-angle scattering dominance. The T^{3/2} dependence reflects faster averaging over the potential at higher thermal velocities. This approach is applicable for low to moderate doping where screening is effective. For higher impurity densities, where the Debye screening assumption breaks down due to overlapping impurity potentials, the Conwell-Weisskopf approximation offers an alternative. It employs a hard-sphere cutoff at a distance related to the average impurity spacing, rather than exponential screening, yielding a mobility μ_{ii} \propto \frac{T^{3/2}}{N_i \ln(1 + \beta^2) - \beta^2}, where β incorporates screening parameters. This formulation better captures scattering in dense, degenerate cases but overestimates rates at low densities compared to Brooks-Herring. Screening effects are central to both models, governed by the Debye length λ_D = \sqrt{\frac{\epsilon_0 \epsilon k_B T}{e^2 n}}, where n is the free carrier density and k_B is Boltzmann's constant. This length scale represents the distance over which the electric field of an impurity is shielded by mobile charges; λ_D increases with temperature (∝ √T) and decreases with carrier density (∝ 1/√n), leading to weaker screening and potentially stronger long-range scattering at low T or low n. In doped semiconductors, typical λ_D values range from 1-10 nm at room temperature for n ~ 10^{17}-10^{19} cm^{-3}, significantly influencing mobility in devices like transistors.

Interface, Alloy, and Other Scattering

Interface scattering, particularly scattering, significantly limits electron in devices like MOSFETs where electrons are confined near material interfaces. In such structures, fluctuations in the interface plane perturb the electron wavefunction, leading to relaxation. The limited by scattering, μ_sr, is inversely proportional to the product of the square of the root-mean-square roughness height Δ and the correlation length Λ of the roughness, expressed as μ_sr ∝ 1/(Δ² Λ). This dependence arises because larger Δ increases the perturbation strength, while longer Λ affects the spatial spectrum of events, both enhancing the scattering rate. For typical MOSFETs, Δ values around 0.2-0.5 nm and Λ ≈ 1-2 nm can reduce effective mobility by orders of magnitude compared to bulk values, emphasizing the need for atomically smooth interfaces in advanced devices. Alloy scattering becomes prominent in compound semiconductors with mixed compositions, such as ternary alloys like Al_x Ga_{1-x} As, where random fluctuations in atomic arrangement create potential barriers for electrons. These compositional potentials scatter electrons via short-range interactions, with the scattering potential often modeled using the virtual crystal approximation. The alloy-limited mobility μ_alloy scales inversely with the x(1-x), which peaks at x=0.5 for maximum alloying effect, as μ_alloy ∝ 1/[x(1-x)]. In p-type Al_x Ga_{1-x} As with doping around 2×10^{17} cm^{-3}, this mechanism can reduce hole mobility from approximately 150 cm²/V·s at x=0 to below 90 cm²/V·s at x=0.5, highlighting its role in limiting transport in heterostructures like HEMTs. The strength of alloy scattering is characterized by a potential on the order of 0.5 , influencing device performance in and high-speed transistors. Piezoelectric scattering occurs in non-centrosymmetric crystals like GaAs, where lattice strain from acoustic s generates electric fields that couple to electrons via the piezoelectric effect. This mechanism is particularly relevant at low s and moderate carrier densities, where it competes with deformation potential . The piezoelectric-limited mobility μ_pz follows a dependence of μ_pz ∝ T^{-1/2}, as the rate increases with thermal population while the matrix element scales with wavevector. In GaAs, this contributes to overall mobility reductions at cryogenic s, with theoretical models showing continued decline unlike saturation in other processes; experimental validations confirm this behavior in high-purity samples. For two-dimensional electron gases in GaAs heterostructures, piezoelectric can dominate acoustic limited transport, affecting high-mobility applications. Other scattering mechanisms, such as electron-electron interactions, further influence through inelastic processes like and oscillations. In semiconductors, electron-electron scattering redistributes momentum within the carrier ensemble but relaxes it via phonon-assisted events or collective excitations, particularly in degenerate systems. The resulting μ_ee scales as μ_ee ∝ T^{3/2} / n, where n is carrier density, reflecting increased collision rates with and screening effects at higher densities. In GaAs at carrier concentrations around 10^{16} cm^{-3}, this can reduce total by up to 10% at 80 K, especially when combined with polar optical or impurity scattering, though it is less dominant than phonon mechanisms at . These rates are computed using , which provides the transition probability 1/τ = (2π/ℏ) |M|^2 g(E_f), where |M| is the matrix element, g(E_f) the density of final states at the , linking microscopic interactions to macroscopic relaxation times τ that determine via μ = e τ / m^*.

Temperature and Material Dependencies

Temperature Dependence

Electron mobility in semiconductors exhibits a pronounced dependence governed by the interplay of mechanisms. In pure or lightly doped materials, where acoustic dominates, decreases with increasing according to the power-law relation \mu \propto T^{-m} with m typically between 1.5 and 2.5, reflecting enhanced lattice vibrations that impede motion. In contrast, heavily doped semiconductors under impurity-limited conditions show increasing with , approximately \mu \propto T^{1.5}, as improves screening of ionized impurities. At low temperatures (below ~100 ), ionized prevails, causing to rise sharply with due to higher velocities and screening effects that reduce the effective rate, following \mu \propto T^{3/2} in the Brooks-Herring model. As rises (above ~200 ), overtakes, leading to a decline in as s interact more frequently with s. This transition manifests in experimental log-log plots of , where samples display an initial upward slope ( regime) followed by a downward trend ( regime), with peak mobilities around 500-1500 cm²/V·s at intermediate temperatures (e.g., 100-200 ) for doping levels of 10^{16}-10^{18} cm^{-3}. Similar behavior is observed in , where log-log plots reveal phonon-limited mobility scaling as \mu \propto T^{-1.5} at higher temperatures, yielding values from ~35,000 cm²/V·s at 77 to ~8500 cm²/V·s at 300 in lightly doped samples. In emerging two-dimensional materials like and MoS₂, recent studies (2020-2025) indicate a weaker dependence, attributed to reduced and suppressed in low dimensions, with mobilities often remaining above 1000 cm²/V·s up to 400 under phonon-limited conditions.

Matthiessen's Rule

Matthiessen's rule originated in the 1860s from the work of Augustus Matthiessen, who studied the electrical conductivities of pure metals and alloys to understand how impurities and temperature affect resistivity. In its original form for metals, the rule posits that the total electrical resistivity \rho_{\text{total}} is the sum of a temperature-independent component due to impurities \rho_{\text{residual}} and a temperature-dependent ideal component \rho_{\text{ideal}}(T), such that \rho_{\text{total}}(T) = \rho_{\text{residual}} + \rho_{\text{ideal}}(T). This empirical relation arose from Matthiessen's extensive measurements on over 200 alloys, aiming to establish standards for electrical resistance amid contemporary rivalries in and . The was later extended to semiconductors in the context of , where electron mobility \mu inversely relates to resistivity via \rho = m / (n e^2 \tau) and \mu = e \tau / m^* (with \tau the relaxation time, m^* the effective mass, n the carrier density, and e the charge). For constant carrier density, the additivity of resistivities translates to the reciprocal additivity of mobilities from independent mechanisms. In semiconductors, Matthiessen's rule is expressed as \frac{1}{\mu_{\text{total}}} = \sum_i \frac{1}{\mu_i}, where \mu_i is the mobility limited by the i-th scattering mechanism, such as phonons or impurities. Equivalently, in terms of relaxation times for uncorrelated processes, \frac{1}{\tau_{\text{total}}} = \sum_i \frac{1}{\tau_i}. The rule holds under the assumption that scattering mechanisms are independent and do not interfere, allowing simple superposition of their rates. However, it breaks down when scatterers exhibit strong correlations or interactions, such as in or where competing mechanisms alter individual rates. For instance, in Ga_{0.47}In_{0.53}As , deviations up to 20% arise due to the failure of reciprocal additivity in the presence of . This framework finds practical application in analyzing experimental electron mobility data, particularly by decomposing total into contributions from distinct mechanisms like and ionized . In Ga_{1-x}Al_xAs alloys, for example, the rule enables accurate fits to Hall mobility measurements across compositions at , isolating and effects with errors below 10%. Such separations aid in material optimization for devices like transistors, though corrections are needed at low temperatures where deviations increase.

Doping Concentration Effects

In lightly doped semiconductors, electron mobility remains relatively high and is dominated by lattice scattering, but as doping concentration (N_d) increases into the moderate range (typically 10^{16} to 10^{18} cm^{-3}), ionized impurity scattering from donor ions causes mobility to decrease approximately as 1/N_d, reflecting the enhanced interactions between carriers and fixed charges. This dependence arises from the quantum-mechanical treatment of scattering probabilities in the Brooks-Herring approximation, where the relaxation time scales inversely with the impurity density under screened conditions. In heavily doped , electron mobility (μ_e) drops significantly to values below 100 cm²/(V·s) for N_d exceeding 10^{19} cm^{-3}, primarily due to degeneracy effects where the penetrates into the conduction band and increased from clustering. Degeneracy alters distribution, confining electrons to higher energies near the (E_F), leading to enhanced rates; in this regime, mobility scales approximately with degeneracy factors and density, as derived in dopant-specific models. Empirical models, such as that developed by et al., provide fits for these behaviors in across concentrations up to 10^{20} cm^{-3} and temperatures from 250 to 500 , capturing the transition from non-degenerate to degenerate transport. Recent advancements in nanoscale devices, such as FinFETs, reveal further mobility reductions due to high local doping concentrations in channels or extensions, exacerbating and quantum confinement effects beyond bulk models; for instance, in ultra-scaled nanosheet FETs with N_d ~10^{19} cm^{-3}, effective μ_e can fall 20-30% lower than bulk predictions owing to these integrated mechanisms. Updated empirical fits incorporating nanoscale , building on Arora's , have been proposed in the 2020s to account for these deviations in device simulations.

Mobility in Disordered Systems

Multiple Trapping and Release

In amorphous semiconductors, electron mobility is frequently governed by the multiple trapping and release () model, in which carriers excited into extended states above the mobility edge experience repeated trapping in localized tail states and subsequent thermal release back into the extended states. This mechanism dominates transport in materials lacking long-range order, such as hydrogenated (a-Si:H), where the localized states arise from structural disorder in the band tails. The MTR framework, originally formulated to explain dispersive time-of-flight experiments, assumes rapid trapping compared to transport in extended states, with release rates determined by thermal activation over energy barriers associated with trap depths. The effective mobility \mu_\mathrm{eff} under this model is reduced from the untrapped extended-state mobility \mu_0 by the fraction of time carriers spend in delocalized states, expressed as \mu_\mathrm{eff} = \mu_0 f_c, where f_c represents the proportion of carriers in the conduction band. For a distribution of trap depths, f_c is influenced by the and profile of localized states, typically leading to a thermally activated behavior where \mu_\mathrm{eff} \propto \exp(-E_t / kT), with E_t denoting the characteristic trap depth and kT the . This activation arises because deeper traps require higher temperatures for efficient release, limiting the average carrier velocity under an applied field. In practical applications, such as a-Si:H-based solar cells, the MTR model accounts for the observed low effective electron mobilities, typically ranging from 1 to 10 cm²/(V·s) at 300 K, which constrain device efficiency but enable cost-effective thin-film fabrication. These values reflect the balance between extended-state and in tail states with depths around 0.1–0.2 , as verified through drift-mobility measurements.

Variable Range

In highly disordered systems, such as amorphous semiconductors or doped organics, electron transport at low temperatures occurs primarily through phonon-assisted hopping between localized states near the , rather than extended-state conduction. This (VRH) mechanism arises because electrons are strongly localized due to disorder, forming a tail of states in the bandgap. To minimize the required for hopping, electrons do not jump to nearest-neighbor sites but instead select optimal paths that balance spatial distance and energy mismatch, effectively trading off (more available distant states) against energy barriers for thermally activated transitions. Nevill F. Mott developed the foundational model for VRH in , assuming a near the and phonon-assisted transitions governed by Miller-Abrahams rates. In this framework, the follows \sigma = \sigma_0 \exp\left[ -\left( \frac{T_0}{T} \right)^{1/4} \right], where T_0 = \frac{18}{k_B N(E_F) \xi^3} incorporates the N(E_F) at the E_F, the localization radius \xi, and Boltzmann's k_B. Since carrier density n remains roughly at low temperatures in these insulating regimes, the electron mobility \mu scales as \mu \propto \frac{\sigma}{n e} \propto \frac{1}{T} \exp\left[ -\left( \frac{T_0}{T} \right)^{1/4} \right], reflecting the temperature-dependent hopping probability and factors. This T^{-1/4} exponent is characteristic of three-dimensional systems and has been widely observed in experiments on amorphous materials. Subsequent refinements by Efros and Shklovskii in 1975 accounted for electron-electron interactions, which deplete states near the due to the Coulomb gap—a soft gap of width \Delta \approx e^2 / \kappa \xi in the , where \kappa is the . In interacting systems, this leads to a modified VRH conductivity \sigma = \sigma_0 \exp\left[ -\left( \frac{T_{ES}}{T} \right)^{1/2} \right], with T_{ES} = \frac{e^2}{4\pi \epsilon_0 \kappa k_B \xi}, and a corresponding \mu \propto \exp\left[ -\left( \frac{T_{ES}}{T} \right)^{1/2} \right] / T^{1/2}. The T^{-1/2} form dominates at sufficiently low temperatures where the Coulomb interaction energy exceeds the Mott hopping energy scale, often observed as a crossover from Mott to Efros-Shklovskii behavior in doped samples. In practical examples, VRH governs transport in doped , where disorder from molecular packing localizes carriers, yielding electron mobilities below $10^{-3} cm²/(V·s) at cryogenic temperatures (e.g., below 100 ). Similarly, in chalcogenide glasses like Ge-Sb-S alloys, low-temperature conduction follows Mott VRH across wide ranges (10–300 ), with mobilities in the $10^{-4}–$10^{-3} cm²/(V·s) regime due to extended hopping networks in the localized state tail. These systems highlight VRH's role in enabling finite, albeit low, conductivity in deeply insulating materials without delocalization.

Measurement Techniques

Hall Effect Mobility

The Hall effect provides a direct method to measure electron mobility in semiconductors by exploiting the on charge carriers in a . When a I flows through a sample of thickness t in the presence of a B, a transverse Hall voltage V_H develops across the sample, given by V_H = \frac{I B}{n e t} for electrons, where n is the carrier density and e is the . The Hall coefficient R_H, defined as R_H = \frac{V_H t}{I B}, equals -\frac{1}{n e} for electrons, allowing independent determination of n from the measurement. Hall mobility \mu_H is then calculated as \mu_H = |R_H| \sigma, where \sigma is the electrical conductivity obtained from separate resistivity measurements. For practical implementation, especially in thin-film or arbitrary-shaped samples, the Van der Pauw geometry is widely used, employing four ohmic contacts placed at the periphery of the sample. The procedure involves measuring the longitudinal voltage V_x across one pair of contacts under current I (with B = 0) to determine , followed by applying B and measuring the transverse Hall voltage V_H across the orthogonal pair, often with field and current reversals to eliminate offsets like thermoelectric voltages. The Hall mobility is then approximated as \mu_H = \frac{V_H}{V_x B}, with corrections applied for non-ideal geometries using Van der Pauw equations to ensure accuracy. This method requires no precise knowledge of sample dimensions beyond thickness for bulk interpretation. A key advantage of Hall effect measurements is the ability to separately extract carrier density n and \mu_H, providing insights into both doping and mechanisms without assumptions about material uniformity. However, limitations arise in inhomogeneous samples, where spatial variations in carrier density or thickness can distort the Hall voltage, leading to erroneous \mu_H values that do not represent average bulk properties. In uniform semiconductors with high-quality ohmic contacts and controlled conditions, the technique achieves accuracy of approximately \pm 5\% for determination.

Field-Effect Mobility

Field-effect mobility (\mu_{FE}) is a key parameter extracted from the drain current-gate voltage (I_D-V_G) characteristics of field-effect transistors (FETs), reflecting carrier transport in the channel under gate modulation. This method is widely used in MOSFETs and thin-film transistors to evaluate device performance, as it directly relates to transconductance and output conductance in different operating regimes. The extraction relies on the gradual channel approximation, which assumes a uniform potential drop along the channel for long-channel devices. In the linear region, applicable at low drain-source voltages (V_D \ll V_G - V_T) where the exhibits ohmic , \mu_{FE} is derived from the : \mu_{FE} = \frac{L}{W C_{ox} V_D} \frac{\partial I_D}{\partial V_G} Here, L and W are the length and width, C_{ox} is the per unit area, and \partial I_D / \partial V_G is obtained from the slope of the I_D-V_G curve at constant low V_D. This formula assumes constant and neglects short- effects, providing a low-field measure of transport. In the saturation regime, at higher V_D > V_G - V_T where the channel pinches off near the drain, \mu_{FE} is extracted using the saturation current: \mu_{FE} = \frac{2 L}{W C_{ox}} \frac{I_{D,sat}}{(V_G - V_T)^2} This expression stems from the square-law model for long-channel devices but accounts for velocity saturation in shorter channels, where carriers reach a maximum velocity (v_{sat}) under high lateral fields, reducing the apparent \mu_{FE} compared to low-field values. The linear regime probes uniform low-field conditions, while saturation reveals high-field limitations, with velocity saturation becoming prominent in sub-micron channels. Accurate extraction in saturation requires corrections for series contact resistance, which can overestimate \mu_{FE} by up to a factor of 2 if unaddressed; methods such as the transfer length model or analysis of channel-length dependence are employed to isolate intrinsic channel mobility from contact contributions. In MOSFETs, \mu_{FE} for electrons typically ranges from 200 to 600 cm²/V·s in the linear regime, but values are often 40-50% lower than Hall mobilities due to Coulomb scattering from interface traps at the Si/SiO₂ boundary, which overestimate inversion charge in capacitance-based corrections. Similarly, in 2D material FETs like MoS₂ or WSe₂, \mu_{FE} (e.g., 10-100 cm²/V·s) is interface-limited by substrate and dielectric scattering, frequently much lower than bulk or Hall-measured mobilities in exfoliated flakes. These discrepancies highlight \mu_{FE}'s sensitivity to surface effects in gated structures.

Advanced Optical and Time-Resolved Methods

Advanced optical and time-resolved methods provide non-contact approaches to probe electron mobility by exciting carriers with light pulses and monitoring their transient response, offering insights into intrinsic transport properties without interference. These techniques are particularly valuable for studying ultrafast dynamics in materials where traditional electrical methods are challenging due to contacts or thin geometries. The time-of-flight (TOF) method involves photoexcitation of charge carriers near one in a sample under an applied , followed by measurement of the transit time t_{\trans} for carriers to reach the opposite , yielding mobility via \mu = \frac{d^2}{V t_{\trans}}, where d is the sample thickness and V is the applied voltage. This technique, pioneered for amorphous semiconductors, reveals drift mobilities by analyzing decay signals, distinguishing between dispersive and non-dispersive transport regimes. TOF is widely used for , where it quantifies field-dependent mobilities on the order of 1–10 cm²/(V·s). Terahertz (THz) employs broadband THz pulses to probe carrier dynamics, fitting or spectra to the to extract mobility as \mu = \frac{e \tau}{m^*}, with e the charge, \tau the relaxation time, and m^* the effective . Time-resolved variants, such as optical-pump THz-probe, track photoinduced changes in on timescales, enabling separation of intraband processes. This approach has been instrumental in characterizing semiconductors and nanostructures, providing frequency-dependent mobility values up to thousands of cm²/(V·s) in high-quality samples. Time-resolved microwave conductivity (TRMC) measures the change in reflectivity or following a photoexcitation , quantifying transient \Delta \sigma to derive an effective mobility \Sigma \mu_{\TRMC} = \frac{\Delta \sigma}{e \Delta n}, where e is the and \Delta n the photoinduced carrier density. Developed for studying charge generation and recombination in photoconductive materials, TRMC is sensitive to one-dimensional or anisotropic , yielding sum mobilities \Sigma \mu that reflect the product of and . It excels in assessing bulk properties of powdered or thin-film samples without geometric constraints. These methods offer key advantages for investigating thin films, , and two-dimensional () materials, where contact fabrication can introduce artifacts or limit accessibility. In organics and perovskites, they enable mobility assessments in non-ideal morphologies, often revealing higher intrinsic values than electrical techniques due to avoidance of trapping at interfaces. Recent applications (2020–2025) in materials, such as and transition metal dichalcogenides, have demonstrated contactless mobilities exceeding 10^5 cm²/(V·s) via THz , highlighting ballistic and enabling optimization for high-speed electronics.

References

  1. [1]
    4: Carrier Drift and Mobility - Engineering LibreTexts
    Jul 5, 2021 · Carrier mobility is useful as it is the ratio of drift velocity to the electric field strength. Below we will give the mathematical definition ...
  2. [2]
    High electron mobility in strained GaAs nanowires - Nature
    Nov 17, 2021 · Owing to the high mobility of electrons in III–V semiconductors, this class of materials is the ideal candidate for high-speed low-power logic ...
  3. [3]
    Improving carrier mobility in two-dimensional semiconductors with ...
    Jun 9, 2022 · Here we show that lattice distortions can reduce electron–phonon scattering in 2D materials and thus improve the charge carrier mobility.
  4. [4]
    Metals: the Drude model of electrical conduction - DoITPoMS
    The Drude model is a simplistic model for conduction. It uses classical mechanics, and treats the solid as a fixed array of ions, with electrons not bound.
  5. [5]
    The Drude Model - SciELO
    Oct 7, 2024 · In 1900, Paul Drude made a groundbreaking advancement with a model to describe the electrical conductivity of metals. Over 120 years later, ...Missing: paper | Show results with:paper
  6. [6]
    [PDF] Motion and Recombination of Electrons and Holes
    When there is more than one scattering mechanism, the total scattering rate. (1/τ) and therefore the total mobility are determined by the sum of the inverses.
  7. [7]
    [PDF] Lecture1- Drude Model - IISc Physics
    Ion cores of atoms immobile (core electrons remain tightly bound to nucleus); valence electrons are loosely bound and in metal can wander around the entire ...
  8. [8]
    Electrical properties of Silicon (Si)
    Breakdown field, ≈3·105V/cm. Mobility electrons, ≤1400 cm2 V-1s-1. Mobility holes, ≤450 cm2 V-1s-1. Diffusion coefficient electrons, ≤36 cm2/s.
  9. [9]
    Drift | PVEducation
    Drift Equation.​​ To derive the drift equation let's consider the bulk of semiconductor. If the electric field Ex is applied in the x-direction each electron ...
  10. [10]
    [PDF] Lecture 3 Electron and Hole Transport in Semiconductors Review
    • The constant μn is called the electron mobility. It has units: • In pure ... Units of conductivity are: Ohm-1-cm-1 or Ω-1-cm-1 or S-cm-1. Units of ...
  11. [11]
    [PDF] ( ) ( )r
    μ = mobility tensor. Electrons in the conduction band are to be thought of as negatively charged particles. In case of multiple electron pockets, current ...
  12. [12]
    [PDF] Drude, P., 1900, Annalen der Physik 1, 566
    Zur Elektronentheorie der Metalle; von P. Drude. I. Teil. Dass die Elektricitätsleitung der Metalle ihrem Wesen nach nicht allzu verschieden von der der ...
  13. [13]
    [PDF] Book 2 Basic Semiconductor Devices for Electrical Engineers
    𝜎 = J. ℇ. Therefore. 𝜎 = q (n 𝜇n + p 𝜇p). (2.74). The resistivity ρ is the reciprocal of σ. 𝜌 = 1. 𝜎 = 1. q (n 𝜇n+p 𝜇p). It must be pointed out ...
  14. [14]
    [PDF] Conduction in Semiconductors -Review
    – n/p drift current density: Jxn = σ n Ex = qμ n. nnEx, Jxp = σ p Ex = qμ p. ppEx. – total drift current density in x direction Jx = q(μn n + μp p) Ex = σ Ex.
  15. [15]
    [PDF] Zur Elektronentheorie der Metalle - Gilles Montambaux
    SOMMERFELD: Zur Elektronentheorie der Metalle. 377 weil andernfalls alle Elektronen das Metall ver- lassen und dieses seinen metallischen Charakter.
  16. [16]
    [PDF] EE 116 Lectures 12-15 - Diffusion of carriers
    Jan 19, 2025 · So What is the Seebeck Coefficient? • Seebeck coefficient can be thought of as the heat per carrier per degree K (specific heat per carrier) ...
  17. [17]
    [PDF] derivation of einstein relation
    34a. DERIVATION OF EINSTEIN RELATION. In equilibrium, the density of particles having temperature T in an electric potential U is. N = No exp.
  18. [18]
  19. [19]
    [PDF] Observing the Hall Effect, Part II: Germanium (Electrical Transport ...
    For example in GaAs, the electron mobility at room temperature is. ~8500 cm2/Vs, whereas for holes the mobility is ~400 cm2/Vs. In Ge, the numbers are ~3900 ...<|separator|>
  20. [20]
    [PDF] Silicon GaAs properties and universal constants
    Silicon's band gap is 1.124 eV, electron mobility is 1400 cm2/V·s. GaAs's band gap is 1.42 eV, electron mobility is 8800 cm2/V·s. Universal constants include ...
  21. [21]
    Enhanced Conductivity of Multilayer Copper–Carbon Nanofilms via ...
    Feb 5, 2025 · However, significant electron–electron scattering in copper results in relatively low electron mobility, about 43 cm2 V−1 s−1, which limits the ...
  22. [22]
    Proximity screening greatly enhances electronic quality of graphene
    Aug 20, 2025 · Transport in high-quality hBN-encapsulated graphene is primarily limited by edge scattering and background fluctuations ('Sources of disorder' ...
  23. [23]
    Deterministic grayscale nanotopography to engineer mobilities in ...
    Aug 13, 2024 · Our results show that monolayer MoS2 FETs under tensile strain achieve an 8-fold increase in on-state current, reaching mobilities of 185 cm²/Vs ...
  24. [24]
    Unveiling high-mobility hot carriers in a two-dimensional conjugated ...
    May 13, 2025 · Here we demonstrate highly mobile hot carriers in crystalline two-dimensional conjugated coordination polymer Cu 3 BHT (BHT, benzenehexathiol) films.
  25. [25]
    Extremely High Electron Mobility in GeSn Epitaxial Films by ...
    Apr 21, 2025 · Very high electron mobilities up to 6,200 and 1,500 cm2V−1s−1 are achieved in the strain-relaxed Ge0.88Sn0.12 film at 50 and 300 K, respectively ...
  26. [26]
  27. [27]
  28. [28]
  29. [29]
    [PDF] Investigation of electron mobility and saturation velocity limits in ...
    Jun 29, 2017 · The polar-optical- phonon limited mobility and the saturation velocity show a. 5% variance with the orientation mostly due to the anisot ...
  30. [30]
    Velocity Saturation - an overview | ScienceDirect Topics
    Velocity saturation refers to the phenomenon where the electron drift velocity in a MOSFET device saturates at high electric fields due to phonon scattering, ...
  31. [31]
    The Gunn effect - ResearchGate
    Aug 6, 2025 · The basic physical mechanism responsible for the effect is the progressive transfer of hot electrons from high mobility valleys to low mobility ...
  32. [32]
    [PDF] Enhancement-mode InP n-channel metal-oxide-semiconductor field ...
    Compared to GaAs, InP is widely believed to be a more forgiving material with respect to Fermi-level pinning and has a higher electron saturation velocity (2⫻ ...Missing: silicon | Show results with:silicon
  33. [33]
    3.2.6 Saturation Velocity - IuE
    The saturation velocity is extracted for field values of 100 kV/cm and above for the GaAs and InP based materials. It is considered a material property here.
  34. [34]
    Electron-Phonon Scattering - an overview | ScienceDirect Topics
    In general, the dominant scattering mechanisms are changed by increasing temperature to acoustic phonons scattering and eventually optical phonons scattering at ...
  35. [35]
    [PDF] Acoustic Phonon Scattering | nanoHUB
    Now, we would interpret δE. F as the deformation potential electron-. Phonon interaction. This leads to: u. ⋅∇=∆. [. ]rqi q rqi q q q q e a eaeq. mN . , .
  36. [36]
    Electron–acoustic phonon coupling in single crystal CH3NH3PbI3 ...
    Feb 8, 2017 · For instance, a T−3/2 dependence has been reported for the mobility, which is characteristic of electron–acoustic phonon scattering. Furthermore ...Missing: μ_ac | Show results with:μ_ac
  37. [37]
    Modeling of electron mobility in GaN materials - ResearchGate
    Aug 8, 2025 · The mobility due to lattice scattering is proportional to T −3/2 , while the mobility due to impurity scattering is proportional to N i −1 T 3/2 ...
  38. [38]
    Theory of quasistatically screened electron-polar optical phonon ...
    The scattering of electrons with polar optical phonons (POP) is an important mechanism that limits electronic transport and determines electron mobility in ...
  39. [39]
    Theory of Quasi-Statically Screened Electron-Polar Optical Phonon ...
    May 16, 2025 · The scattering of electrons with polar optical phonons (POP) is an important mechanism that limits electronic transport and determines electron ...
  40. [40]
    [PDF] Untitled - MAK HILL Publications
    Where, bo, is the polar optical phonon energy. op. Non-polar optical phonon scattering: Non-polar optical phonon scattering is similar to deformation potential.
  41. [41]
    [PDF] arXiv:2210.01746v1 [cond-mat.mtrl-sci] 4 Oct 2022
    Oct 4, 2022 · In the high-temperature regime, the average po- lar optical phonon T-dependence is reduced to µ ∝ T−1.5 and the interquartile range ...
  42. [42]
    Acoustic Phonon - an overview | ScienceDirect Topics
    The relaxation time τ is determined by various scattering mechanisms, such as acoustic phonon ... (232) 1 τ ∝ T ( ħ ω c ) 1 / 2. Ito et al. (1966) ...
  43. [43]
    Phonon-limited electron mobility in graphene calculated using tight ...
    Sep 5, 2012 · We present a calculation of the electron-phonon scattering rates in ideal monolayer graphene using the third-nearest-neighbor (3NN) ...<|control11|><|separator|>
  44. [44]
    [PDF] Phonon-limited electron mobility in graphene calculated using tight ...
    Phonon-limited electron mobility in graphene is calculated using tight-binding Bloch waves, with values around 107 cm2=Vs (50K) and 106 cm2=Vs (300K) at 10^12 ...Missing: μ_ac ∝
  45. [45]
    Theory of Impurity Scattering in Semiconductors | Phys. Rev.
    Another probable source of resistance is scattering by ionized impurity centers. This resistance can be calculated by using the Rutherford scattering formula.
  46. [46]
    Theory of the screened Coulomb field generated by impurity ions in ...
    Apr 19, 2006 · This theory examines the effect of the screened Coulomb field generated by impurity ions on various semiconductor phenomena.
  47. [47]
    Improved surface-roughness scattering and mobility models for multi ...
    Jun 23, 2017 · The parameters of the surface roughness spectrum are Δrms = 0.21 nm and correlation length Λ = 1.4 nm. Calculations obtained with Fourier modes ...
  48. [48]
    (PDF) Alloy scattering in p‐type AlxGa1-xAs - ResearchGate
    Aug 8, 2025 · Alloy scattering is very important, and in fact reduces the hole mobility from 150 to less than 90 cm2/V s at x=0.5. The main parameter in the ...
  49. [49]
    Electron mobility in GaAs due to piezoelectric scattering - IOPscience
    Theoretical studies so far on piezoelectric scattering have resulted in a mobility behaviour of mu approximately T-1/2 in which mu continues to fall ...
  50. [50]
    Effect of electron‐electron scattering on mobility in GaAs
    Apr 1, 1982 · The influence of electron‐electron scattering on mobility in GaAs is studied over the temperature range 80–300 K using the variation ...
  51. [51]
    Electron-electron scattering and mobilities in semiconductors and ...
    Nov 15, 1986 · Electron-electron scattering significantly reduces mobility in semiconductors and quantum wells, especially when the collision rate is faster ...Missing: 3/2} / | Show results with:3/2} /
  52. [52]
    [PDF] Chapter 24 - Fermi's Golden Rule
    Here the density of states g(Ek) is per unit volume, in units 1/(eV.m3), as is standard in semiconductor physics. The scattering rate is linearly proportional ...Missing: mobility | Show results with:mobility
  53. [53]
    Ionized-Impurity Scattering Mobility of Electrons in Silicon | Phys. Rev.
    The curves have been used in an experimental test of the applicability of the Brooks-Herring formula to the ionized-impurity scattering of electrons in silicon ...Missing: original | Show results with:original
  54. [54]
    [PDF] The dopant density and temperature dependence of electron ...
    Electron mobility at 300 K was deduced from resis- tivity and junction ... and electron mobility data for phosphorus-doped silicon are presented. 4.1 ...
  55. [55]
    [PDF] Temperature dependence of minority and majority carrier mobilities ...
    We show that majority carrier mobility in heavily doped n- and p-GaAs is essentially temperature independent while minority carrier mobility has a roughly 1/T ...
  56. [56]
    Recent Advances in the Carrier Mobility of Two-Dimensional Materials
    Jun 11, 2020 · Here, we review significant recent advances and important new developments in the carrier mobility of 2D materials based on theoretical investigations.Introduction · Methods · Transport Properties of 2D... · Future Prospects and...
  57. [57]
    Phonon-limited carrier mobility modeling of two-dimensional ...
    In this case, tensile strain should reduce the electron mobility of 2D semiconductors. However, recent experimental studies have demonstrated that tensile ...<|control11|><|separator|>
  58. [58]
    V. On the effect of the presence of metals and metalloids ... - Journals
    On the effect of the presence of metals and metalloids upon the electric conducting power of pure copper. Augustus Matthiessen.
  59. [59]
    Augustus Matthiessen - Engineering and Technology History Wiki
    Jan 18, 2016 · "Matthiessen's rule" attempted to approximate the relationship between electrical resistivity, ideal resistivity and residual resistivity, ...Missing: origin | Show results with:origin
  60. [60]
    Validity of Matthiessen's rule for calculating electron mobility in Ga 1
    Oct 1, 1985 · Hall electron mobilities in Ga1−xAlxAs alloys for various compositions have been analyzed using the Matthiessen's rule and Monte Carlo ...
  61. [61]
    [PDF] Scattering of electrons by ionized impurities in semiconductors - arXiv
    The studies of Conwel and Weisskopf [1], Brooks. [2] and Herring (his work is unpublished), as well as Ridley [3] have already become the classical research ...Missing: original | Show results with:original
  62. [62]
    Electron mobility of heavily doped semiconductors including multiple ...
    Aug 15, 2023 · A theoretical treatment of the multiple scattering problem for electrons in heavily doped semiconductors is developed for the purpose of resolving a long- ...
  63. [63]
  64. [64]
    Influence of Dopant Species on Electron Mobility in Heavily Doped ...
    Apr 18, 2025 · We present a new theoretical approach to study the influence of the dopant species on the electron drift mobility in doped semiconductors ...
  65. [65]
  66. [66]
    The study of electron mobility on ultra-scaled silicon nanosheet FET
    In this paper, we investigate the impact of phonon (ph) and surface roughness (SR) scattering on the electron mobility of n-type silicon NSFETs.
  67. [67]
    Electron mobility enhancement in nanoscale silicon-on-insulator ...
    Electron mobility in nanoscale silicon-on-insulator (SOI) layers with a doping concentration ranging from 2 × 1017 cm−3 to 1 × 1019 cm−3 is thoroughly studied.
  68. [68]
    Revisiting multiple trapping and release electronic transport in ...
    Aug 7, 2024 · Multiple trapping and release (MTR) is a typical transport mechanism of electron carriers in amorphous and other disordered semiconductors where ...
  69. [69]
    [PDF] Transport, Interfaces, and Modeling in Amorphous Silicon Based ...
    In three different papers, we have reviewed the overall picture for hole and electron drift-mobilities in amorphous and microcrystalline silicon [1,10]. One ...
  70. [70]
    Charge transport in non-crystalline semiconductors - SpringerLink
    Aug 19, 2007 · Some experimental work is described which distinguishes the two types of charge transport and shows what happens at the transition between them.
  71. [71]
    Coulomb gap and low temperature conductivity of disordered systems
    Journal of Physics C: Solid State Physics, Volume 8, Number 4Citation A L Efros and B I Shklovskii 1975 J. Phys. C: Solid State Phys. 8 L49DOI 10.1088/0022 ...
  72. [72]
    Charge Transport in Organic Semiconductors | Chemical Reviews
    Transport measurements on molecular single crystals indicate that charge mobilities as high as a few hundreds cm2/V·s can be attained at low temperature (up to ...
  73. [73]
    Observation of variable-range hopping conduction in a Ge Sb S ...
    Mott's variable-range hopping was found to be dominant in the whole temperature range from 300 K to 10 K. Co-induced gap states in the middle of the mobility ...
  74. [74]
    [PDF] Hall Effect Measurement Handbook - Quantum Design
    The mobility is fundamentally a measure of the velocity of the carriers in the material and is often the most important material parameter measured by the Hall ...
  75. [75]
    The Hall Effect | NIST - National Institute of Standards and Technology
    Apr 15, 2010 · The objective of the Hall measurement in the van der Pauw technique is to determine the sheet carrier density ns by measuring the Hall voltage ...
  76. [76]
  77. [77]
    [PDF] MOS Transistor
    A MOSFET is a field-effect transistor, a basic building block of circuits, that acts as an on-off switch controlled by gate voltage.
  78. [78]
    Mobility overestimation due to gated contacts in organic field-effect ...
    Mar 10, 2016 · Charge mobility, extracted from current–voltage curves, is an important parameter for evaluating the performance of organic field-effect ...
  79. [79]
    [PDF] Mobility Measurement and Degradation Mechanisms of MOSFETs ...
    the mobility obtained from the conventional split-C–V method is significantly lower than that obtained from the Hall-effect method (by about 40–50%), which ...
  80. [80]
    Two- and four-probe field-effect and Hall mobilities in transition ...
    Regardless of the electrical characteristics of the contacts, the field-effect mobility is much lower than the Hall mobility. Electrical contacts in the WS2 or ...<|control11|><|separator|>