Fact-checked by Grok 2 weeks ago

Variable-range hopping

Variable-range hopping (VRH) is a theoretical model describing electrical conduction in disordered solids, such as amorphous semiconductors and lightly doped crystalline insulators, at low temperatures, where charge carriers transport via thermally activated hops between localized electronic states, with the characteristic hopping distance increasing as temperature decreases to optimize the trade-off between spatial separation and energy mismatch relative to the . This mechanism dominates when the is insufficient for carriers to overcome barriers to nearest-neighbor sites, favoring longer-range hops to more energetically favorable distant sites within the localized state manifold. VRH was first proposed by British Nevill F. Mott in 1969 to explain conduction in non-crystalline materials exhibiting activated but weakly temperature-dependent resistivity. In Mott's original formulation, assuming a constant density of states N(E_F) at the Fermi energy E_F and exponential localization of wavefunctions with inverse length \alpha, the DC conductivity in three dimensions follows the relation \sigma(T) = \sigma_0 \exp\left[ -\left( \frac{T_0}{T} \right)^{1/4} \right], where \sigma_0 is a prefactor on the order of the extended-state conductivity, T is temperature, and the characteristic temperature T_0 = \frac{18}{k_B N(E_F) \alpha^3} incorporates the localization length \alpha and Boltzmann constant k_B. This T^{-1/4} dependence generalizes to T^{-1/(d+1)} in d dimensions, distinguishing VRH from nearest-neighbor hopping, which yields a simple Arrhenius form \sigma \propto \exp(-E_a / k_B T) with fixed activation energy E_a. In 1975, A. L. Efros and B. I. Shklovskii refined the model by incorporating long-range Coulomb interactions between localized electrons, which suppress the density of states near E_F over an energy scale of order e^2 \alpha / \kappa (with dielectric constant \kappa), creating a "Coulomb gap" and altering the hopping energetics. The Efros-Shklovskii (ES) VRH regime predicts a dimension-independent dependence \sigma(T) = \sigma_0 \exp\left[ -\left( \frac{T_0}{T} \right)^{1/2} \right], with T_0 \approx \frac{e^2}{\kappa k_B a}, where a is a characteristic localization radius, typically observable at lower s or higher than Mott VRH, with crossovers between the two regimes depending on parameters like and . VRH conduction has been experimentally verified in diverse systems, including chalcogenide glasses, impurity bands in doped semiconductors, granular metals, and colloidal films, where it governs below characteristic temperatures often in the range of 1–100 K. Extensions of VRH incorporate magnetic fields, which can suppress the gap and induce Mott-like behavior, as well as finite-size effects in low-dimensional nanostructures.

Physical Background

Hopping Conduction Mechanism

In disordered materials, such as amorphous semiconductors and doped insulators, charge transport at low temperatures primarily occurs through hopping conduction, where charge carriers move between localized states rather than extended states. This dominates when the thermal energy is insufficient to excite carriers into delocalized states above the mobility edge, leading to thermally activated transitions between sites within the localized state regime. The process involves phonon-assisted tunneling, allowing carriers to overcome energy barriers between sites of varying energies and positions. The physical basis for hopping arises from the inherent in these systems, which causes wavefunctions to localize below the mobility edge—a separating extended and localized states. In such insulators or semiconductors, random potentials from impurities or structural irregularities result in an of wavefunctions, characterized by a localization $1/\alpha, preventing diffusive band-like motion and confining carriers to finite regions. This localization phenomenon, fundamentally tied to , ensures that direct overlaps between distant states are negligible, making sequential hops the primary transport pathway. As a baseline, nearest-neighbor hopping (NNH) describes transport where carriers primarily jump to the closest localized sites, with the dictated by the energy difference between the initial site i and the neighboring site j. The transition rate for such hops is captured by the Miller-Abrahams formula, which accounts for both the spatial overlap and the energetic barrier: \begin{align} \nu_{ij} &= \nu_0 \exp(-2\alpha r_{ij}) \exp\left(-\frac{E_j - E_i}{kT}\right) \quad \text{for } E_j > E_i, \\ \nu_{ij} &= \nu_0 \exp(-2\alpha r_{ij}) \quad \text{for } E_j \leq E_i, \end{align} where \nu_0 is the phonon frequency (attempt rate), \alpha is the inverse localization length, r_{ij} is the inter-site distance, E_i and E_j are the site energies, k is Boltzmann's constant, and T is the . This rate emphasizes that downhill hops (lower ) occur more readily without beyond tunneling. At lower temperatures, where kT becomes comparable to or smaller than the typical spacing between nearest-neighbor sites, the NNH model becomes inefficient due to high barriers. Carriers then favor variable-range hops to more distant sites that offer smaller mismatches, trading increased tunneling probability (via larger r_{ij}) for reduced . This optimization balances spatial and energetic costs, enabling conduction over varying ranges that grow with decreasing , thus transitioning from fixed-range NNH to variable-range hopping (VRH).

Localized States in Disordered Systems

In disordered systems, such as with random potentials, wavefunctions can become localized due to quantum effects, preventing classical and leading to insulating behavior. This phenomenon, known as , was first theoretically described in three dimensions by in 1958, where he modeled transport in an "impurity band" of a with randomly placed impurities, showing that sufficiently strong disorder confines wavefunctions to finite regions rather than allowing them to extend throughout the material. The localization arises from the random potential landscape created by , which disrupts the coherence of electron propagation; in three-dimensional systems, wavefunctions decay exponentially away from their centers, with a localization that decreases as disorder strength increases. Below a critical energy threshold, known as the mobility edge, all states are localized, while above it, states become extended and conductive; this energy-dependent transition was proposed by N.F. Mott to distinguish the regime where can propagate freely from one where they are trapped. At low temperatures, the typically lies within the localized state regime in moderately disordered materials, suppressing metallic conduction and favoring thermally activated processes. In disordered semiconductors, localized states emerge prominently due to structural imperfections, including impurities, defects, and the absence of long-range order in amorphous structures. Impurities introduce potential fluctuations that trap electrons in bound states within the bandgap, while defects such as dangling bonds or coordination irregularities in amorphous networks, like (a-Si:H), create deep gap states that further contribute to localization. The amorphous structure itself, lacking periodic translation , broadens the into tails of localized states extending into the gap, arising from variations in local bonding and atomic positions. The (DOS) in these systems reflects this disorder, often exhibiting exponential tails near the band edges due to the statistical distribution of potential fluctuations from defects and amorphous topology. For the Mott variable-range hopping model, the DOS is typically assumed to be nearly constant near the within the localized regime, providing a of states available for thermal activation, though real materials show these exponential tails dominating transport properties at low energies.

Mott Variable-Range Hopping

Core Assumptions

The Mott variable-range hopping (VRH) model describes charge transport in disordered materials at low temperatures, where electrons hop between localized states over varying distances to minimize the barrier. Proposed by Nevill Mott in 1969 as an extension of his work on conduction in amorphous semiconductors, the model emphasizes that at sufficiently low temperatures, nearest-neighbor hopping becomes improbable due to high energies, favoring longer-range hops that trade increased spatial separation for reduced energy mismatches. A foundational assumption is a constant density of states g(E) = g_0 near the Fermi energy E_F, implying no significant pseudogap or depletion in available states at the relevant energies, which allows for the statistical estimation of optimal hopping paths. This constant g_0 enables the model to predict the availability of states within an energy window of order kT around E_F, without correlations depleting the spectrum. The model explicitly neglects electron-electron interactions, treating hopping as a single-particle process where occupied and unoccupied states exchange electrons independently of Coulomb effects. Hopping is phonon-assisted to conserve during transitions between non-resonant states, with phonons providing or absorbing the necessary |E_i - E_j| \approx kT for hops spanning distances much larger than the average inter-site separation. The localized states arise from spatial in the material, such as random potentials or structural irregularities, leading to exponentially decaying wavefunctions \psi(r) \sim \exp(-\alpha r), where \alpha^{-1} is the localization length characterizing the spatial extent of each state. This decay governs the tunneling probability, exponentially suppressing hops beyond a few localization lengths. The model applies in the low-temperature regime where kT is much smaller than the typical energy spacings between localized states near E_F, rendering direct thermal excitation to extended states negligible, yet allowing variable-range hops that optimize the trade-off between spatial and energetic barriers.

Derivation of Conductivity Law

The derivation of the conductivity law in Mott's variable-range hopping (VRH) model relies on a percolation theory framework to identify the dominant hopping paths that connect localized states across a macroscopic sample at low temperatures. In disordered systems with localized states, conduction occurs via phonon-assisted tunneling between sites, where the hopping rate between two sites separated by distance R and energy difference W is given by \nu \propto \exp\left(-2\alpha R - \frac{W}{kT}\right), with \alpha = 1/\xi the inverse localization \xi, k Boltzmann's , and T . The overall \sigma is dominated by the bottleneck hops in a percolating , corresponding to the maximum value of the exponent \beta = 2\alpha R + W/kT along the optimal path; minimizing this \beta determines the temperature dependence. Under the core assumption of a constant density of states g_0 near the Fermi level, the typical energy W required for a hop over distance R is set by the condition that there is approximately one accessible state within a sphere of radius R and an energy window of width W. The volume of the sphere is V = \frac{4}{3} \pi R^3, so the number of states is g_0 V W \approx 1, yielding W \approx \frac{1}{g_0 V} = \frac{3}{4\pi g_0 R^3}. This relation reflects the percolation criterion in the simplest approximation, where the critical number of overlapping sites is order unity; more refined treatments adjust this to a percolation constant B_c \approx 2.8 for three-dimensional networks, but the form remains similar. Substituting into the exponent gives \beta(R) = 2\alpha R + \frac{3}{4\pi g_0 k T R^3}. To find the optimal hop distance R_\mathrm{opt} that minimizes \beta, differentiate with respect to R: \frac{d\beta}{dR} = 2\alpha - \frac{9}{4\pi g_0 k T R^4} = 0, which solves to R_\mathrm{opt}^4 = \frac{9}{8\pi \alpha g_0 k T}, \quad R_\mathrm{opt} = \left( \frac{9}{8\pi \alpha g_0 k T} \right)^{1/4}. At this optimum, the two terms in \beta satisfy $2\alpha R_\mathrm{opt} = 3 \frac{W}{kT}, so \beta_\mathrm{min} = 2\alpha R_\mathrm{opt} + \frac{2\alpha R_\mathrm{opt}}{3} = \frac{8}{3} \alpha R_\mathrm{opt}. Substituting R_\mathrm{opt} yields \beta_\mathrm{min} \propto (1/T)^{1/4}. The minimal exponent is thus \beta_\mathrm{min} = \left( T_0 / T \right)^{1/4}, where the characteristic temperature T_0 encapsulates the material parameters. In the standard form accounting for the percolation threshold, T_0 = \frac{18 \alpha^3}{g_0 k}, leading to the conductivity \sigma = \sigma_0 \exp\left[ -\left( \frac{T_0}{T} \right)^{1/4} \right], with prefactor \sigma_0 depending on microscopic details like phonon frequency and overlap integrals. This T_0 arises from refining the simple unity condition to g_0 k T V \approx B_c, where B_c^{4/3} contributes the factor of approximately 18 in three dimensions. The T^{-1/4} scaling emerges directly from the dimensional optimization in three-dimensional space, balancing the exponential decay in distance against the thermal accessibility of states over larger volumes at lower temperatures.

Efros-Shklovskii Variable-Range Hopping

Role of Electron-Electron Interactions

In disordered systems with localized electron states, electron-electron repulsion via interactions plays a pivotal role in modifying the single-particle () near the E_F, leading to the formation of a soft gap. This gap arises as a deviation from the constant assumed in earlier models without interactions, suppressing the availability of states at E_F and altering low-temperature transport properties. The physical origin of the Coulomb gap stems from the energy cost associated with adding or removing an from a localized site in the presence of charged neighboring sites. When an electron is added to an empty site or removed from an occupied one, the repulsion from surrounding charged centers increases the , effectively depleting the DOS at E_F. This many-body ensures that the lowest-energy single-particle excitations require finite , creating a quadratic suppression in the DOS: g(E) \propto |E - E_F|^2 for |E - E_F| < E_c, where E_c characterizes the gap width. The Efros-Shklovskii (ES) model, which incorporates these interactions, assumes a disordered where wavefunctions are strongly localized due to potential fluctuations, with the localization much smaller than the average inter- distance, and the is half-filled (compensated) such that both and holes are present in the localized states. These assumptions allow for a self-consistent treatment of the potential from random charged impurities, validating the quadratic DOS form in three dimensions. The model was developed by Alexei Efros and Boris Shklovskii in to address discrepancies between observed low-temperature conductivity in doped semiconductors and predictions from non-interacting hopping theories. The width of the Coulomb gap, E_c \approx e^2 / \kappa r_0, where e is the electron charge, \kappa is the dielectric constant, and r_0 is the average inter-electron distance (inversely related to the unperturbed DOS g_0), scales with the strength of disorder: stronger localization from increased disorder reduces the effective screening and enlarges E_c. This interaction-dominated regime is valid when the Coulomb energy exceeds the disorder-induced broadening of states, typically in moderately disordered insulators at low temperatures.

Derivation Accounting for Coulomb Gap

In the Efros-Shklovskii (ES) model, the derivation of the variable-range hopping (VRH) conductivity law incorporates the quadratic density of states (DOS) arising from the Coulomb gap, which suppresses states near the Fermi level \mu. Unlike the constant DOS assumed in the Mott model, the interaction-modified DOS in three dimensions takes the form g(\epsilon) = \frac{\kappa^3}{2\pi e^6} |\epsilon - \mu|^2, where \kappa is the dielectric constant and e is the electron charge; this quadratic dependence g(E) \propto |E - \mu|^2 significantly alters the available states for hopping. The percolation approach to transport considers hops between localized states in the interaction-modified energy landscape, where the effective energy window W for viable transitions is set by the Coulomb energy scale W \approx e^2 / (\kappa R), balancing the electrostatic repulsion between the charged initial site and the target site at distance R. This replaces the thermally random energy fluctuations of the Mott picture with a distance-dependent barrier. The hopping rate is then governed by the combined tunneling and thermal activation factors, leading to an exponent that must be minimized for the dominant percolation path. To derive the optimal hop, the total exponent \phi in the conductivity prefactor is expressed as \phi = 2\alpha R + \frac{e^2}{\kappa R k_B T}, where \alpha = 1/\xi is the inverse localization length \xi, k_B is Boltzmann's constant, and T is ; the first term accounts for tunneling probability \exp(-2\alpha R), while the second reflects the thermal overcoming of the \exp\left(-\frac{e^2}{\kappa R k_B T}\right). Minimizing \phi with respect to R yields \frac{d\phi}{dR} = 2\alpha - \frac{e^2}{\kappa (k_B T) R^2} = 0, so R_\mathrm{opt} = \left( \frac{e^2}{2\alpha \kappa k_B T} \right)^{1/2} \propto T^{-1/2}. Substituting back gives the minimum \phi_\mathrm{min} = 2 \sqrt{2\alpha \frac{e^2}{\kappa k_B T}} = \left( \frac{T_\mathrm{ES}}{T} \right)^{1/2} , where the characteristic temperature is T_\mathrm{ES} = \frac{2.8 e^2 \alpha}{\kappa k_B} (with the numerical prefactor 2.8 obtained from detailed ). The condition for percolation requires that the number of accessible states within the sphere of radius R_\mathrm{opt} and energy window W \approx e^2 / ([\kappa](/page/Kappa) R_\mathrm{opt}) is on the order of a B_c \approx 10-20, ensuring network connectivity. With the quadratic DOS, the integrated number of states is N \approx \frac{4\pi}{3} R^3 \int_{-W}^{W} g(E) \, dE \propto R^3 W^3 \propto R^3 \left( \frac{1}{R} \right)^3 \propto \mathrm{constant}, independent of R and thus of ; this contrasts with the Mott case and confirms that the optimization alone sets the scale without additional constraints from state scarcity. Notably, T_\mathrm{ES} depends only on fundamental parameters like \xi, \kappa, and e, rendering the ES law independent of the unperturbed DOS g_0. The resulting follows the ES VRH law: \sigma = \sigma_0 \exp\left[ -\left( \frac{T_\mathrm{ES}}{T} \right)^{1/2} \right], where \sigma_0 is a temperature-independent prefactor; this \sqrt{T} dependence emerges directly from the T^{-1/2} scaling of R_\mathrm{opt} and the quadratic integration. At low temperatures T < T_c, where T_c marks the crossover from Mott to ES regimes by equating the respective exponents (typically T_c \sim T_\mathrm{M}^2 / T_\mathrm{ES}, with T_\mathrm{M} the Mott scale), the ES law dominates due to the enhanced role of long-range effects in the gapped .

Comparisons and Extensions

Differences Between Mott and Efros-Shklovskii Models

The Mott variable-range hopping (VRH) model and the Efros-Shklovskii (ES) VRH model differ fundamentally in their treatment of electron interactions and the resulting temperature dependence of conductivity in disordered systems. In the Mott model, which assumes a constant density of states near the Fermi level and neglects electron-electron interactions, the conductivity follows \sigma \propto \exp\left(- (T_0 / T)^{1/4}\right) in three dimensions, reflecting a balance between thermal activation energy and hopping distance. In contrast, the ES model incorporates long-range Coulomb interactions, leading to a soft Coulomb gap in the density of states and a steeper low-temperature dependence, \sigma \propto \exp\left(- (T_{ES} / T)^{1/2}\right), which dominates when interactions suppress states near the Fermi energy. This exponent difference—1/4 for Mott versus 1/2 for ES—results in ES VRH exhibiting a more pronounced resistivity increase at low temperatures, making it observable in regimes where Mott's shallower dependence would otherwise apply. The characteristic parameters further highlight these distinctions. The Mott parameter T_0 scales as T_0 \propto 1 / (g_0 \alpha^3), where g_0 is the at the and \alpha is the inverse localization length, emphasizing dependence on the availability of states and wavefunction overlap. Conversely, T_{ES} in the ES model is independent of g_0 and given by T_{ES} \propto e^2 / (\kappa \alpha k_B), relying instead on the constant \kappa that screens interactions, reflecting the model's focus on interaction strength over state density. These parameter dependencies underscore how Mott VRH is sensitive to disorder-induced state localization, while ES VRH prioritizes electrostatic effects in correlated systems. Regarding validity regimes, the Mott model applies at higher temperatures or in weakly interacting systems where the thermal energy exceeds the Coulomb interaction scale, such as in amorphous semiconductors with minimal doping. The ES model becomes relevant at lower temperatures in strongly interacting environments, particularly doped semiconductors where the Coulomb gap forms due to electron-electron repulsion, suppressing hopping within a narrow energy window near the . A crossover between the regimes occurs below a characteristic T_c \approx (T_0^2 / T_{ES})^{1/3}, where ES VRH governs as interactions dominate; above T_c, Mott behavior prevails. Both models assume three-dimensional geometry, limiting their direct applicability to lower dimensions; extensions to two dimensions yield modified exponents, such as T^{-1/3} for Mott VRH, due to altered percolation paths in reduced dimensionality.

Experimental Observations and Applications

Experimental observations of variable-range hopping (VRH) are characterized by distinct signatures in the temperature dependence of electrical . In Mott VRH, plotting the logarithm of conductivity against T^{-1/4} yields a straight line at low temperatures, indicating three-dimensional hopping between localized states without strong electron-electron interactions. Similarly, for Efros-Shklovskii (ES) VRH, a linear relationship emerges when plotting \ln \sigma versus T^{-1/2}, reflecting the influence of the gap in the . These signatures have been observed in various disordered systems, such as ES VRH in heavily doped and at millikelvin temperatures, where conductivity spans several orders of magnitude. Seminal theoretical work and supporting experiments on chalcogenide glasses in the late demonstrated low-temperature following the predicted T^{-1/4} dependence. Confirmation of the ES model came in the 1970s and 1980s through measurements of variable-range in doped semiconductors, including n-type InSb and GaAs-AlGaAs heterostructures, where the T^{-1/2} law and positive aligned with predictions accounting for interactions. These early works established VRH as a dominant in insulators near the metal-insulator transition. VRH plays a key role in applications across . In , such as thin-film transistors and thermoelectric devices, charge transport follows VRH models, enabling with mobilities tuned by disorder. Granular metals exhibit VRH-dominated conduction below the , useful for modeling nanoscale interconnects and sensors. At low temperatures, VRH governs the behavior of neutron-transmutation-doped thermistors, providing high sensitivity for cryogenic temperature sensing down to millikelvin ranges. Additionally, in polymer nanocomposites, VRH integrates with to describe thresholds in carbon black-filled systems, informing the design of conductive fillers for . Challenges in VRH observations include deviations from ideal behavior due to external factors. Phonon drag can enhance thermopower in VRH regimes, altering effective at intermediate temperatures in materials like chalcogenide spinels. induce positive , shrinking the effective hopping range and causing nonlinearities, as seen in granular systems under moderate fields. Post-2000 studies have extended VRH to novel systems, revealing two-dimensional (2D) variants. In graphene and its derivatives, 2D Mott or ES VRH explains low-temperature transport in disordered flakes, with T^{-1/3} or T^{-1/2} dependencies observed in epitaxial and reduced oxide samples. Topological insulators like Bi₂Se₃ nanowires show robust 2D ES VRH, linking hopping to in ultra-narrow geometries. In quantum dot films, such as CdSe assemblies, crossovers between Mott and ES regimes highlight density-of-states effects, advancing colloidal nanocrystal . Recent work on , including reactively sputtered oxides, demonstrates ES VRH explaining anomalous low-temperature resistivity in amorphous multicomponent films. More recent studies as of 2025 have observed VRH in Zintl phases like EuIn₂P₂ and compacted VO₂ nanopowders, further expanding its relevance to for .

References

  1. [1]
    Lecture 8: Mott's Variable Range Hopping | Theory of Solids II
    Lecture notes on Mott variable range hopping theory.
  2. [2]
    Variable-Range Hopping Conduction - SpringerLink
    This chapter deals with hopping conduction at temperatures which are so low that typical resistances between neighboring impurities become larger.
  3. [3]
    Conduction in non-crystalline materials - Taylor & Francis Online
    Conduction in non-crystalline materials. III. Localized states in a pseudogap and near extremities of conduction and valence bands. N. F. Mott Cavendish ...
  4. [4]
    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 ...
  5. [5]
    Mott and Efros-Shklovskii Variable Range Hopping in CdSe ...
    The model of variable range hopping conductivity predicts a crossover between Mott and Efros-Shklovskii as a function of temperature and density of states.
  6. [6]
    Hopping Conductivity in Disordered Systems | Phys. Rev. B
    Oct 15, 1971 · A model in which charge is transported via phonon-induced tunneling of electrons between localized states which are randomly distributed in energy and position.
  7. [7]
    Theory of impurity band hopping conduction - IOPscience
    The authors extend the theory of impurity band hopping conduction to lower temperatures. The theory of Miller and Abrahams (1960) is critically examined.
  8. [8]
    Impurity Conduction at Low Concentrations | Phys. Rev.
    Impurity Conduction at Low Concentrations. Allen Miller* and Elihu Abrahams. Physics Department, Rutgers University, New Brunswick, New Jersey. *Present ...
  9. [9]
    On the theory of hopping conductivity in disordered systems
    Aug 10, 2025 · Starting with the linearised master equation, the authors present a first-principles theory of conductivity for disordered systems.
  10. [10]
    Absence of Diffusion in Certain Random Lattices | Phys. Rev.
    This paper presents a simple model for such processes as spin diffusion or conduction in the "impurity band." These processes involve transport in a lattice ...Missing: theory | Show results with:theory
  11. [11]
    Nobel Prize in Physics 1977
    ### Summary of Mobility Edge and Localized States in Disordered Systems from Mott's Nobel Lecture
  12. [12]
    Defects in amorphous semiconductors: Philosophical Magazine B
    Abstract. The electronic structure of defects in some canonical amorphous semiconductors, namely a-Se, a-As, a-Si: H, a-GaAs, a-Si3N4 and a-SiO2, is reviewed. ...
  13. [13]
    [PDF] Nevill Mott - Nobel Lecture
    The density of states in the conduction band of a non-crystalline material, showing the mobility edge Ec separated by an energy from the band edge. We now ...
  14. [14]
    Half-century of Efros–Shklovskii Coulomb gap: Romance with ...
    Dec 1, 2024 · Efros-Shklovskii variable range hopping transport in nanocluster metallic films. J. Appl. Phys. (May 2012). Crossover between Mott and Efros ...
  15. [15]
    Two-dimensional Mott variable-range hopping transport in a ... - arXiv
    Jan 26, 2018 · ... T^{-1/3} temperature dependence, an evidence for the two-dimensional (2D) Mott VRH transport. The measured low-field magnetoresistance of ...Missing: extensions exponent
  16. [16]
    Variable-range-hopping conduction and low thermal conductivity in ...
    Aug 8, 2025 · ... Mott variable-range hopping is the dominant conduction mechanism. However, at high temperatures (above 350 K), ρ and S decrease ...
  17. [17]
    [PDF] Efros-Shklovskii variable range hopping in reduced graphene oxide ...
    The difference between the Mott and ES-VRH is in the details of their localization. Page 2. 2 parameters, density of states (DOS) and interactions that ...
  18. [18]
    Mott, N.F. (1968) Conduction in Glasses Containing Transition Metal ...
    ABSTRACT: A set of borophosphate glasses doped with alkali and transition metal (TM) ions have been synthesized. The glasses were carried through; annealing, ...Missing: variable range chalcogenide
  19. [19]
    Variable-range hopping charge transport in organic thin-film ...
    Feb 23, 2020 · A common assumption in these approaches is that the charge is spatially and energetically localized in states, wells, or grains of the amorphous ...Missing: core | Show results with:core
  20. [20]
    Effective description of hopping transport in granular metals - arXiv
    Jan 5, 2005 · We develop a theory of a variable range hopping transport in granular conductors based on the sequential electron tunnelling through many grains.Missing: applications | Show results with:applications
  21. [21]
    Low temperature hopping conduction in neutron transmutation ...
    The temperature dependence of variable range hopping resistivity ρ in neutron transmutation doped (NTD) isotopically enriched70Ge:Ga samples is reported.<|control11|><|separator|>
  22. [22]
    Electric conductivity in silicone-carbon black nanocomposites
    The influence of filler volume fraction, temperature and electric field strength, are discussed using percolation and variable range hopping on a fractal as ...
  23. [23]
    Large Thermoelectricity via Variable Range Hopping in Chemical ...
    Large values of up to ∼30 mV/K at room temperature are observed, which are much larger than those observed in other two-dimensional crystals and bulk MoS 2.
  24. [24]
    Variable-range-hopping magnetoresistance | Phys. Rev. B
    Mar 15, 1991 · The hopping magnetoresistance R of a two-dimensional insulator with metallic impurities is considered. In sufficiently weak magnetic fields it increases or ...Missing: confirmation 1970s
  25. [25]
    Variable range hopping and nonlinear transport in monolayer ...
    Sep 13, 2016 · We report experimental results on variable range hopping (VRH) and nonlinear transport in monolayer epitaxial graphene.Missing: original | Show results with:original
  26. [26]
    Evidence of robust 2D transport and Efros-Shklovskii variable range ...
    Aug 10, 2017 · We report the experimental observation of variable range hopping conduction in focused-ion-beam (FIB) fabricated ultra-narrow nanowires of topological ...<|control11|><|separator|>
  27. [27]
    Mott and Efros-Shklovskii Variable Range Hopping in CdSe ...
    The model of variable range hopping conductivity predicts a crossover between Mott and Efros-Shklovskii as a function of temperature and density of states.Missing: original | Show results with:original
  28. [28]
    Thin films made by reactive sputtering of high entropy alloy ...
    Feb 28, 2022 · The amorphous HEO has an electrical conduction interpreted as due to variable range hopping described by the Efros–Shklovskii theory. The ...