Fact-checked by Grok 2 weeks ago

Nearly free electron model

The nearly free electron model (NFEM) is a quantum mechanical framework in solid-state physics that approximates the behavior of conduction electrons in crystalline solids as plane waves propagating freely but weakly perturbed by the periodic electrostatic potential arising from the ionic lattice. This model builds directly on the gas theory by incorporating a small periodic potential V(\mathbf{r}) = \sum_{\mathbf{G}} V_{\mathbf{G}} e^{i \mathbf{G} \cdot \mathbf{r}}, where \mathbf{G} are vectors, leading to the formation of bands separated by band gaps, particularly at the boundaries. Unlike the , which predicts a continuous parabolic dispersion relation E(\mathbf{k}) = \frac{\hbar^2 k^2}{2m} without gaps, the NFEM uses perturbation theory to show how lattice scattering mixes states with wavevectors differing by \mathbf{G}, opening gaps of magnitude approximately $2|V_{\mathbf{G}}| and enforcing Bloch's theorem that electron wavefunctions take the form \psi_{\mathbf{k}}(\mathbf{r}) = u_{\mathbf{k}}(\mathbf{r}) e^{i \mathbf{k} \cdot \mathbf{r}} with periodic u_{\mathbf{k}}(\mathbf{r}). Developed as part of Felix Bloch's seminal 1928 doctoral work, the NFEM provided the first rigorous quantum explanation for the in periodic potentials, resolving classical puzzles about electron motion in crystals and laying the foundation for understanding metals, insulators, and semiconductors. Bloch's analysis demonstrated that even weak potentials suffice to produce the observed periodicity in reciprocal space, with energy E(\mathbf{k}) = E(\mathbf{k} + \mathbf{G}), which is essential for the model's predictive power. This approach assumes the potential is sufficiently weak (V_{\mathbf{G}} \ll E_F, where E_F is the ) to justify first-order degenerate near zone edges, making it particularly applicable to simple metals like metals where valence electrons experience minimal scattering. The model's key insights include the distortion of Fermi surfaces near zone boundaries, which explains anomalies in electrical conductivity and , and it complements the tight-binding model for stronger potentials by providing a unified picture of band formation across materials. In practice, the NFEM enables calculations of band structures using methods like the pseudopotential approximation, influencing modern applications in design and . Its limitations, such as neglecting strong electron-electron interactions, are addressed in more advanced theories like , but it remains a cornerstone for interpreting experimental data from .

Background and Motivation

Historical Context

The nearly free electron model emerged in the 1920s and 1930s as was extended to describe electron behavior in solid-state materials, building on the realization that electrons in crystals could be treated as waves propagating through periodic potentials. This development was spurred by Louis de Broglie's 1924 hypothesis proposing wave-particle duality for matter, which suggested that electrons exhibit wave-like properties with wavelengths inversely proportional to their momentum. De Broglie's idea laid the groundwork for applying wave mechanics to solids, influencing subsequent work on electron dynamics in lattices. Shortly after, Erwin Schrödinger's 1926 formulation of the wave equation provided the mathematical framework for solving electron motion in potential fields, including those of crystalline structures. In the late 1920s, advanced early applications of to metals by refining the classical with Fermi-Dirac statistics, treating electrons as a free gas to explain electrical conductivity and other transport properties. This served as a precursor but overlooked the periodic lattice potential, prompting further refinements. The pivotal breakthrough came in 1928 with Felix Bloch's doctoral thesis at the University of under , where he demonstrated that electron wave functions in a periodic potential take the form of plane waves modulated by the lattice periodicity—now known as Bloch waves. Bloch's work directly applied Schrödinger's equation to one-dimensional crystals, revealing how lattice periodicity leads to energy band structures. Bloch's theorem, central to his 1928 publication in Zeitschrift für Physik, established the foundational principle that electron states in crystals are extended waves compatible with , enabling the prediction of allowed and forbidden energy ranges. This insight resolved longstanding puzzles in metal conductivity and paved the way for the 's formalization in the 1930s, as researchers like Alan Wilson extended it to three dimensions and distinguished conductors from insulators based on band filling. The model's evolution marked a shift from simplistic approximations to more accurate treatments incorporating weak periodic perturbations, profoundly shaping .

Free Electron Model and Its Shortcomings

The conceptualizes the valence electrons in a metal as a gas of non-interacting particles confined to a potential-free box, allowing them to move freely throughout the solid. This approach stems from early classical electron gas theories and treats the positive cores as a uniform neutralizing background, neglecting any structure. In the quantum mechanical formulation, the energy for these s is parabolic, given by E = \frac{\hbar^2 k^2}{2m}, where \hbar is the reduced Planck's constant, k is the wavevector, and m is the . This results in a continuum of energy states filled up to the according to the . The model achieves notable successes in describing basic metallic properties through the Drude-Sommerfeld theory, which refines the classical approach with quantum statistics. It accurately predicts the electrical of metals, \sigma = \frac{ne^2 \tau}{m}, where n is the , e the charge, and \tau the relaxation time, aligning well with experimental values when using Fermi velocities. Additionally, Sommerfeld's quantum refinement correctly explains the low-temperature electronic as linear in temperature, C_V = \gamma T, where \gamma is the Sommerfeld coefficient, resolving the classical overestimation. Despite these strengths, the free electron model has critical shortcomings that limit its applicability to real solids. By assuming a constant zero potential, it disregards the periodic lattice potential from ion cores, leading to unphysical continuous energy bands without gaps and failing to predict the existence of insulators or semiconductors. For instance, it cannot account for energy band gaps at boundaries, where effects from the would otherwise cause discontinuities in the . Furthermore, the model predicts a perfectly spherical , which contradicts experimental observations of distortions due to effects in many metals. These deficiencies highlight the need for refinements that incorporate weak periodic perturbations.

Core Principles

Bloch's Theorem

Bloch's theorem provides the foundational framework for describing wave functions in a crystal lattice characterized by a periodic potential. Specifically, for a potential V(\mathbf{r}) that satisfies V(\mathbf{r} + \mathbf{R}) = V(\mathbf{r}) for all lattice vectors \mathbf{R}, the solutions to the time-independent take the form \psi_{n\mathbf{k}}(\mathbf{r}) = u_{n\mathbf{k}}(\mathbf{r}) e^{i \mathbf{k} \cdot \mathbf{r}}, where the index n labels different energy bands, the function u_{n\mathbf{k}}(\mathbf{r}) is periodic with the lattice periodicity such that u_{n\mathbf{k}}(\mathbf{r} + \mathbf{R}) = u_{n\mathbf{k}}(\mathbf{r}), and the wave vector \mathbf{k} lies within the first of the . This form was first established by in his seminal 1928 analysis of electron motion in crystal lattices. The implications of are profound for understanding electron behavior in solids: the wave functions represent plane waves modulated by that reflects the , ensuring that the overall wave function transforms under lattice translations by e^{i \mathbf{k} \cdot \mathbf{R}}. This modulation leads to the conservation of quasi-momentum, where \hbar \mathbf{k} serves as the crystal momentum, conserved modulo vectors. Consequently, the theorem proves that electrons in a periodic potential behave as waves carrying crystal momentum \hbar \mathbf{k}, enabling the classification of electronic states by their \mathbf{k} within the and facilitating the band description essential to the nearly free electron model. The derivation of begins with the time-independent for an in a periodic potential: \left[ -\frac{\hbar^2}{2m} \nabla^2 + V(\mathbf{r}) \right] \psi(\mathbf{r}) = E \psi(\mathbf{r}), subject to over a large volume. By considering the of the , which commutes with the translation operator T_{\mathbf{R}} such that T_{\mathbf{R}} H = H T_{\mathbf{R}}, the eigenfunctions can be chosen to be simultaneous eigenfunctions of both H and T_{\mathbf{R}}. This yields T_{\mathbf{R}} \psi(\mathbf{r}) = e^{i \mathbf{k} \cdot \mathbf{R}} \psi(\mathbf{r}), leading directly to the Bloch form \psi(\mathbf{r}) = e^{i \mathbf{k} \cdot \mathbf{r}} u(\mathbf{r}) with u(\mathbf{r}) periodic. In the absence of the potential, where V(\mathbf{r}) = 0, the periodic function u_{n\mathbf{k}}(\mathbf{r}) becomes constant, recovering the plane waves.

Periodic Potential Approximation

The periodic potential approximation in the nearly free electron model posits that the crystal potential experienced by conduction electrons arises primarily from the ionic lattice and can be treated as a weak on otherwise states. This approximation leverages , which guarantees that the potential is periodic with the lattice and thus expandable in a over reciprocal lattice vectors. The potential V(\mathbf{r}) is mathematically expressed as a Fourier series: V(\mathbf{r}) = \sum_{\mathbf{G}} U_{\mathbf{G}} e^{i \mathbf{G} \cdot \mathbf{r}}, where \mathbf{G} are the vectors and the Fourier coefficients U_{\mathbf{G}} represent the strength of the periodic components. The "nearly free" designation stems from the condition that these coefficients satisfy |U_{\mathbf{G}}| \ll E_F, where E_F is the , ensuring the potential energy is much smaller than the of the conduction electrons. This disparity justifies a perturbative approach, where the potential introduces only minor corrections to the dispersion relation, primarily manifesting as band gaps at boundaries. In the zeroth-order empty lattice approximation, the potential is set to zero (V(\mathbf{r}) = 0), yielding unperturbed energy bands given by E_n(\mathbf{k}) = \frac{\hbar^2}{2m} (\mathbf{k} + \mathbf{G}_n)^2, where n labels the band and \mathbf{G}_n are vectors. These bands fold back into the first and touch degenerately at zone boundaries without gaps, providing the baseline for subsequent weak potential effects. This approximation holds well for metals, where conduction s have high kinetic energies, but breaks down in insulators, where the ionic potential is not weak relative to electron energies, necessitating alternative models like tight-binding to account for localized states.

Mathematical Formulation

Hamiltonian Setup

The nearly free electron model begins with the time-independent for a electron in a crystal , given by H \psi = \varepsilon \psi, where H is the and \psi is the wavefunction with energy eigenvalue \varepsilon. The is expressed as H = \frac{p^2}{2m} + V(\mathbf{r}), consisting of the term \frac{p^2}{2m} (with p as the and m the ) and a potential V(\mathbf{r}) arising from the periodic arrangement of ions in the . This potential V(\mathbf{r}) is periodic with the lattice periodicity, satisfying V(\mathbf{r} + \mathbf{R}) = V(\mathbf{r}) for any \mathbf{R}. The model employs the independent electron approximation, treating electrons as non-interacting particles moving in the mean-field potential V(\mathbf{r}) generated by the fixed ions and the average electron distribution, thereby neglecting explicit electron-electron interactions. This simplification allows the to be reduced to solving the single-particle for each electron independently. The periodic nature of V(\mathbf{r}) can be expanded in a as V(\mathbf{r}) = \sum_{\mathbf{G}} V_{\mathbf{G}} e^{i \mathbf{G} \cdot \mathbf{r}}, where the sum is over vectors \mathbf{G}, and V_{\mathbf{G}} are the Fourier coefficients representing the strength of the periodic potential components. To solve the , the wavefunction is expanded in the basis of plane waves consistent with , taking the form \psi_{\mathbf{k}}(\mathbf{r}) = \sum_{\mathbf{G}} C_{\mathbf{k} - \mathbf{G}} e^{i (\mathbf{k} - \mathbf{G}) \cdot \mathbf{r}}, where \mathbf{k} is a wavevector in the first and the sum is over vectors \mathbf{G}. Substituting this expansion into the yields a set of coupled linear equations for the coefficients C_{\mathbf{k} - \mathbf{G}}, which can be assembled into a matrix equation known as the secular equation. This infinite matrix is truncated in practice by considering only the dominant Fourier components of the potential. In the plane-wave basis, the kinetic energy operator is diagonal, with eigenvalues \lambda_{\mathbf{k}} = \frac{\hbar^2 k^2}{2m} for each e^{i \mathbf{k} \cdot \mathbf{r}}, reflecting the free-electron dispersion before the periodic potential mixes states at wavevectors differing by vectors. This diagonal form simplifies the setup, as the off-diagonal elements introduced by V(\mathbf{r}) account for the and mixing of plane waves.

Perturbation Theory Application

In the nearly free electron model, the effects of the weak periodic potential are incorporated perturbatively into the using an expansion in the plane-wave basis. This leads to a formulation for the , where the coefficients C_{\mathbf{k}} satisfy the secular equation (\lambda_{\mathbf{k}} - \varepsilon) C_{\mathbf{k}} + \sum_{\mathbf{G}} U_{\mathbf{G}} C_{\mathbf{k} - \mathbf{G}} = 0, with \lambda_{\mathbf{k}} = \frac{\hbar^2 k^2}{2m} denoting the unperturbed , \varepsilon the perturbed , and U_{\mathbf{G}} the Fourier coefficients of the potential for vectors \mathbf{G}. Away from the boundaries, the unperturbed states with wavevectors \mathbf{k} and \mathbf{k} - \mathbf{G} have sufficiently different (|\lambda_{\mathbf{k}} - \lambda_{\mathbf{k} - \mathbf{G}}| \gg |U_{\mathbf{G}}|), allowing the use of non-degenerate to compute small corrections to the . In contrast, near the zone boundaries, degeneracy arises when \mathbf{k} = \mathbf{G}/2, making \lambda_{\mathbf{k}} = \lambda_{\mathbf{k} - \mathbf{G}} and requiring degenerate to resolve the strong coupling between the two states. In the degenerate case at the zone boundary, the secular equation truncates to a two-by-two for the coefficients C_{\mathbf{k}} and C_{\mathbf{k} - \mathbf{G}}: \begin{pmatrix} \lambda_{\mathbf{k}} - \varepsilon & U_{\mathbf{G}} \\ U_{\mathbf{G}}^* & \lambda_{\mathbf{k} - \mathbf{G}} - \varepsilon \end{pmatrix} \begin{pmatrix} C_{\mathbf{k}} \\ C_{\mathbf{k} - \mathbf{G}} \end{pmatrix} = 0. Since \lambda_{\mathbf{k}} = \lambda_{\mathbf{k} - \mathbf{G}} at degeneracy, the eigenvalues are \varepsilon = \lambda_{\mathbf{k}} \pm |U_{\mathbf{G}}|, with normalized eigenvectors given by the even and odd combinations C_{\mathbf{k}} = \pm C_{\mathbf{k} - \mathbf{G}} = 1/\sqrt{2}. This two-state mixing lifts the degeneracy and determines the splitting at the boundary.

Key Results

Band Gap Formation

In the nearly free electron model, band gaps emerge as a consequence of electron wave by the periodic potential, resulting in destructive that prohibits electron propagation in specific energy ranges. This phenomenon is physically analogous to Bragg reflection in X-ray diffraction, where incident waves constructively interfere upon reflection from crystal planes, leading to forbidden directions; similarly, at certain wavevectors, electron waves backscattered by the form standing waves with nodes at cores or between them, creating regions of zero probability and thus forbidden energies. These gaps specifically open at the boundaries of the , where the free-electron energy levels for wavevectors \mathbf{k} and \mathbf{k} - \mathbf{G} (with \mathbf{G} a vector) become degenerate, allowing the weak periodic potential to couple and split these states. Away from these boundaries, inside the zones, the energy closely follows the parabolic free-electron form E \propto k^2, perturbed only slightly by the . The magnitude of the primary at a zone is $2 |U_{\mathbf{G}}|, where U_{\mathbf{G}} represents the component of the lattice potential for the relevant \mathbf{G}. Higher-order band gaps arise at boundaries of higher Brillouin zones, corresponding to larger reciprocal lattice vectors \mathbf{G}, through processes involving multiple scatterings. These gaps are generally smaller because the Fourier components |U_{\mathbf{G}}| diminish with increasing |\mathbf{G}|, reflecting the shorter-range nature of the potential's higher harmonics. This selective gap opening at zone edges underscores the model's prediction of nearly free-electron behavior modulated by weak lattice effects.

Energy Dispersion in Bands

In the nearly free electron model, the energy dispersion relation \epsilon(\mathbf{k}) deviates from the free electron parabola only in regions close to the Brillouin zone boundaries, where the periodic potential introduces significant perturbations. Away from these boundaries, the dispersion approximates the free electron form \epsilon(\mathbf{k}) \approx \frac{\hbar^2 k^2}{2m}, with m the electron mass and \mathbf{k} the wavevector, reflecting the weak influence of the lattice potential on plane-wave-like states. Near the zone boundaries, the dispersion splits into two branches: \epsilon(\mathbf{k}) \approx \lambda_{\mathbf{k}} \pm |U_{\mathbf{G}}| + higher-order terms, where \lambda_{\mathbf{k}} represents the unperturbed energy at the degenerate points (typically the average free-electron energy), U_{\mathbf{G}} is the component of the potential for reciprocal lattice vector \mathbf{G}, and the \pm terms account for the lifting of degeneracy. This splitting results in band gaps at the boundaries, as briefly referenced in the model's core predictions. Higher-order corrections further refine the curvature but remain small for weak potentials. In the reduced zone scheme, band folding maps the extended free-electron dispersion into the first by translating branches with \mathbf{k} + \mathbf{G}, producing multiple bands derived from the original parabola. These folded branches are separated by gaps at edges, forming a series of continuous bands within the zone, with the lowest band encompassing states from |\mathbf{k}| < \pi/a (for lattice constant a) and higher bands incorporating folded segments from larger k. A key consequence of this is the variation in effective mass m^* = \hbar^2 \left( \frac{\partial^2 \epsilon}{\partial k^2} \right)^{-1} near the band gaps, where the flattened curvature—due to the —leads to an increased m^* compared to the free-electron value, altering electron transport properties. This flattening is most pronounced at the band edges, where m^* can diverge in the lowest-order approximation. For a simple cubic lattice, the first few bands illustrate this structure qualitatively: the lowest (valence-like) band follows the free-electron parabola from the zone center \Gamma to the boundary X (at k_x = \pi/a), flattening near X before the gap; the second band starts above this gap at X, folds back toward \Gamma with reduced curvature, and continues to the next boundary, while higher bands repeat this pattern with progressively larger average energies, all within the cubic . This folding preserves the overall while introducing the banded structure essential for insulators and semiconductors.

Physical Justifications

Weak Potential Validity

The validity of the nearly free electron model hinges on the perturbative assumption that the periodic lattice potential acts as a weak on the free-electron , specifically when the magnitude of the Fourier components of the potential, |V_{\mathbf{G}}|, is much smaller than the E_F, which is typically on the order of a few for metals with delocalized conduction s. This condition ensures that the electron wavefunctions remain close to plane waves, with only minor distortions near boundaries where band gaps form. In such systems, the dominates, allowing first-order to accurately describe the band structure modifications induced by the weak periodic potential. This weak potential criterion is well satisfied in simple metals like the alkali metals, where conduction electrons are highly delocalized. For example, in sodium, the relevant component of the , such as |V_{110}|, is approximately 0.23 eV, which is significantly smaller than the of about 3.2 eV, thereby validating the near-free electron behavior and enabling accurate predictions of properties like the . In contrast, the model breaks down in transition metals, where stronger potentials arising from localized d-band electrons lead to substantial hybridization and deviations from free-electron-like dispersion, necessitating more advanced approaches like the tight-binding model. The effective weakness of the potential is further justified by the screening of cores, which reduces the amplitude of the real-space potential V(r) experienced by electrons. Within the periodic potential framework, this screening arises from the redistribution of conduction electrons around the positively charged cores, effectively softening the ionic interactions. Additionally, the Born-Oppenheimer approximation underpins the model's treatment of the by separating the fast electronic motion from the slower ionic vibrations, assuming fixed positions to define the static periodic potential for electron dynamics. This separation provides a quantitative basis for estimating the potential's weakness, as the electronic energies far exceed typical vibrational energies, ensuring the adiabatic validity of the approximation in metallic systems.

Electron Screening Effects

In the nearly free electron model, electron screening effects play a crucial role in weakening the periodic potential experienced by conduction electrons, making the weak-potential approximation viable. Conduction electrons, behaving as a degenerate , respond to the positive charges of ions by redistributing to shield or neutralize them, thereby damping the long-range interactions. This screening reduces the effective strength of the ionic potential at distances relevant to electron wavefunctions, which typically span multiple sites in metals. The Thomas-Fermi approximation provides a semi-classical description of this shielding, where the conduction electron density adjusts to maintain local charge neutrality around each . In this model, the screened potential decays exponentially as \phi(r) = \frac{q}{r} e^{-k_{TF} r}, with the Thomas-Fermi screening wavevector k_{TF} = \sqrt{4\pi e^2 D(E_F)}, where D(E_F) is the at the . For typical metallic electron densities around $10^{22} cm^{-3}, the screening length $1/k_{TF} is on the order of a few angstroms, much shorter than interatomic distances, effectively the ionic potential V(r) at long range and preventing strong . This mechanism ensures that the potential perturbation remains weak compared to the free-electron , justifying the perturbative treatment in the model. Core electrons contribute to screening by occupying tightly bound orbitals close to the , forming a stable cloud that partially neutralizes the ionic charge seen by electrons, effectively lowering the charge from Z to an effective Z_eff < Z. This inner-shell screening diminishes the overall ionic potential for conduction electrons, further supporting the nearly free electron regime in simple metals where electrons are loosely bound. The effective potential experienced by conduction electrons can thus be expressed as V_{eff}(r) = V_{ion}(r) - V_{electron}(r), where the electron contribution arises from both core and conduction clouds, partially neutralizing the bare ionic potential. In the model, which idealizes the solid as a positive background with delocalized s, this screening leads to an oscillatory but overall weak potential due to —density modulations decaying as $1/r^3 beyond the screening length, driven by the sharp . These effects highlight how collective electron responses maintain the weakness of the periodic potential in nearly free electron systems.

Applications and Limitations

Band Structure Predictions

The nearly free electron model is widely applied to predict the electronic band structures of simple metals, particularly metals like sodium, , and , as well as metals such as , silver, and , where the periodic potential acts as a weak on states. In these materials, the model accounts for the formation of nearly bands by incorporating small energy gaps at boundaries, leading to dispersions that closely resemble parabolas away from these boundaries. (ARPES) experiments on these metals confirm the model's predictions, revealing band dispersions with minimal deviations from behavior and low quasiparticle scattering rates, as observed in sodium where the effective mass enhancement is only about 1.2 compared to the bare . A representative example is the band structure of sodium, a body-centered cubic with one per atom. The model predicts small band gaps of approximately 0.5 eV at key boundaries, such as the L point, arising from Bragg scattering of waves by the weak ionic potential. These gaps slightly distort the otherwise parabolic energy dispersion and are consistent with sodium's measured electrical resistivity, where electron-phonon scattering near the zone boundaries contributes to the temperature-dependent transport properties without significantly altering the metallic conductivity. Beyond basic dispersions, the nearly free electron model provides detailed predictions for the Fermi surface topology in these metals, yielding nearly spherical surfaces that fill a significant portion of the , with minor necks or bulges at zone faces due to the small gaps; for sodium, the Fermi wavevector is about 0.92 ⁻¹, placing the surface close to but inside the zone boundary. The resulting features van Hove singularities—logarithmic peaks or step-like features—near the energies of these band gaps, which influence higher-order electronic properties like electron-phonon coupling although they lie above the in simple metals. Furthermore, the nearly free electron model serves as the foundational framework for pseudopotential theory, enabling calculations of band structures in simple metals by replacing the strong core potentials with weak, screened pseudopotentials that preserve the free electron-like behavior. This approach has been instrumental in quantitative predictions for alkali metals, facilitating computations of cohesive energies, phonon spectra, and transport coefficients with high accuracy.

Comparisons to Other Models

The nearly free electron (NFE) model contrasts with the tight-binding model in its foundational assumptions and regimes of applicability. The NFE model assumes a weak periodic potential, treating electrons as nearly delocalized plane waves suitable for metals where valence electrons experience weak scattering from the , leading to large bandwidths. In contrast, the tight-binding model posits strong atomic potentials that localize electrons in atomic orbitals, making it appropriate for insulators and semiconductors with narrow bands formed by overlapping localized states. This distinction arises because the NFE approach excels in systems with delocalized conduction electrons, such as simple metals, while tight-binding better captures the directional bonding and localization in covalent solids. Compared to (DFT), the NFE model serves as a perturbative starting point for understanding band formation in weakly interacting systems, relying on analytical approximations to a free electron gas perturbed by a potential. DFT, however, provides a numerical framework that solves the Kohn-Sham equations self-consistently to obtain exact ground-state properties in principle, incorporating electron-electron interactions via exchange-correlation functionals without relying on . While the NFE model offers conceptual simplicity for simple metals, DFT enables detailed computations for complex materials, though standard local-density approximations in DFT often overestimate bandwidths relative to experiment in nearly free electron metals. The NFE model has notable limitations, particularly in systems with strong electron correlations, such as Mott insulators, where electron motion is hindered by Coulomb interactions beyond the weak-potential assumption, leading to insulating behavior not captured by perturbative treatments. It also fails for strongly covalent bonds, where localization dominates, and can overestimate bandwidths in materials with moderate correlations. In practice, hybrid approaches mitigate these issues by combining models; for instance, in semiconductors like silicon, the tight-binding model describes the localized valence bands formed by sp³ hybrid orbitals, while the NFE model approximates the more delocalized conduction bands.

References

  1. [1]
    Nearly free electron model - Open Solid State Notes
    In this lecture, we will analyze how electrons behave in solids using the nearly-free electron model. This model considers electrons as plane waves (as in ...
  2. [2]
    [PDF] Solid State Physics NEARLY FREE ELECTRON MODEL (Contd)
    The nearly free electron model modifies the free electron picture by opening up small gaps near the zone boundaries. The free electron approximation is good ...
  3. [3]
    Über die Quantenmechanik der Elektronen in Kristallgittern
    Die Bewegung eines Elektrons im Gitter wird untersucht, indem wir uns dieses durch ein zunächst streng dreifach periodisches Kraftfeld schematisieren.
  4. [4]
    [PDF] Lecture 5 - Nearly Free Electron Model
    Nearly free electron model: weak perturbation of electrons by periodic potential of ions. Lecture 5. 5. Nearly Free Electrons. Consider the effects due to a ...
  5. [5]
    [PDF] FELIX BLOCH - National Academy of Sciences
    Why the free electron ap- proach worked for metals and why the electrons ... Bloch's last original papers were connected with super- conductivity and ...
  6. [6]
    Arnold Sommerfeld develops the free-electron theory of metals
    Nov 23, 2017 · In 1927, Arnold Sommerfeld applies quantum statistics to the Drude model of electrons in metals and develops the free-electron theory of metals ...Missing: gas | Show results with:gas
  7. [7]
    [PDF] 3. Electron Dynamics in Solids kz k k - DAMTP
    Much of the basic theory of band structure was laid down by Felix Bloch in 1928 as part of his doctoral thesis. As we have seen, Bloch's name is attached to ...
  8. [8]
  9. [9]
    [PDF] Section 7: Free electron model
    A free electron model is the simplest way to represent the electronic structure of metals. Although the free electron model is a great oversimplification of ...
  10. [10]
    [PDF] τ τ ε - bingweb - Binghamton University
    Feb 3, 2020 · His theory assumes that electrons are formed of a classical gas. Such a classical model survives even after the quantum mechanics appears in ...
  11. [11]
    [PDF] Energy bands - Rutgers Physics
    The free electron model gives us a good insight into many properties of metals, such as the heat capacity, thermal conductivity and electrical conductivity.Missing: history | Show results with:history<|control11|><|separator|>
  12. [12]
    [PDF] Lecture Notes for Solid State Physics (3rd Year Course 6) Hilary ...
    Jan 9, 2012 · ... Ashcroft and Mermin. Its selection of topics and organization may seem a bit strange in ... 4.5 Shortcomings of the Free Electron Model .
  13. [13]
    [PDF] Über die Quantenmechanik der Elektronen in Kristallgittern
    Die Bewegung eines Elektrons im Gitter wird untersucht, indem wir uns dieses durch ein zun~chst streng dreifaeh periodisches Kraftfeld schematisieren. Unter.
  14. [14]
    [PDF] Solid State Physics - DAMTP
    This result is known as Bloch's Theorem. Here we prove the theorem for our one-dimensional system; we will revisit it in Section 2.3.1 in higher dimensions.
  15. [15]
    [PDF] 6.730 Physics for Solid State Applications
    BLOCH. For wavefunctions that are eigenenergy states in a periodic potential… or. Page 3. Proof of Bloch's Theorem. Step 1: Translation operator commutes with ...Missing: textbook | Show results with:textbook
  16. [16]
    [PDF] Electrons in a weak periodic potential Assumptions: - IISc Physics
    Nearly free electron approximation – Empty lattice approximation: Take the Sommerfeld free electron states – treat the lattice as a weak perturbation - free ...
  17. [17]
    [PDF] Nearly-Free-Electron Model
    Substituting the value for the coefficients into the wave function, and we've at last found a first order solution for our electron in a periodic potential.Missing: Felix paper
  18. [18]
    [PDF] ashcroft mermin
    point seems appropriate the resolution of the difficulties in the free electron model of ... illustrative and very important practical cases of nearly free ...
  19. [19]
    [PDF] 1 The course is taught on basis of lecture notes which are ...
    Ashcroft and Mermin (AM) start out with the free electron model and show that it fails and give then a ... free electron like bands that cross the Fermi level.
  20. [20]
    [PDF] Review of Energy Dispersion Relations in Solids - MIT
    Figure 1.1: One dimensional electron energy bands for the nearly free electron model shown in the extended Brillouin zone scheme. ... 1.2), band gaps Eg and band ...
  21. [21]
    [PDF] Nearly free electron model; perturbation - bingweb
    Here we consider a very simple model (nearly free electron model) in order to understand the fundamental property of the energy gap at the zone boundary. 1.
  22. [22]
    [PDF] Lecture 4 - Nearly Free Electron Model
    4.2 Translational Symmetry – Bloch's Theorem. Bloch's theorem: the wave functions of the electrons in a crystal must be of a special form (the Bloch form) uk ...<|control11|><|separator|>
  23. [23]
    [PDF] reflectivity of nak droplets - AMOS Conference
    For sodium and potassium values of V110,Na = 0.32 eV and V110,K = 0.23 eV according to Butcher [10] and Ashcroft [5] were selected. The results are shown in ...
  24. [24]
    [PDF] Chap 8 Nearly free and tightly bound electrons
    Tight binding model: Energy bands as an extension of atomic orbitals. Page 22. • Alkali metal. • noble metal. • Covalent solid. • d-electrons in transition ...
  25. [25]
    [PDF] (9) Quantum Mechanical Methods: Calculation of the electronic ...
    The nearly free electron model introduces atomic nuclei, which can scatter electrons, and the next stage of improvement is pseudopotentials, which incorporates ...
  26. [26]
    [PDF] MATRL 218/CHEM 227: Class IX — More on electronic ... - UCSB MRL
    1This is referred to as the nearly free electron model. ... screening in a crystal is the so-called Thomas-Fermi formula for the screened Coulomb potential ( ...
  27. [27]
    [PDF] Electronic Structure of the Solid State -How electrons glue crystals ...
    Core electrons are tightly bound to the core. Valence electrons experience the core as screened by the core electrons, i.e. an effect potential -> pseudo ...
  28. [28]
    [PDF] 2 Electron-electron interactions 1 - UF Physics
    neutrality, we therefore assume the electrons move in a neutralizing positive background (“jellium model”). ... moving in the potential created by the impurity ...<|separator|>
  29. [29]
    Lifetime of quasiparticles in the nearly free electron metal sodium
    We report a high-resolution angle-resolved photoemission (ARPES) study of the prototypical nearly free-electron metal sodium.
  30. [30]
    Electronic band structures of the alkali metals and of the noble ...
    In the monovalent metals the electronic band structure is strongly affected by the size of the band gap E s-E p at the Brillouin zone faces, a large gap ...
  31. [31]
    APPLICATION OF THE PSEUDOPOTENTIAL MODEL TO SOLIDS
    Pseudopotentials have also been used to compute electron-phonon couplings in semiconductors and metals. The newer ab initio schemes should yield even better ...
  32. [32]
    [PDF] ψ ϕ x - bingweb
    The tight-binding model is opposite limit to the nearly free electron model. The potential is so large that the electrons spend most of their lives near ionic ...
  33. [33]
    [PDF] Electronic correlation in nearly free electron metals with beyond-DFT ...
    For more than three decades, nearly free-electron elemental metals have been a topic of debate because the computed bandwidths are significantly wider in ...
  34. [34]
    STRONGLY CORRELATED QUANTUM MATERIALS - Anh Ngo Lab
    ... density functional theory (DFT) or the nearly-free-electron model. In strongly correlated materials, the motion of one electron is highly dependent on the ...
  35. [35]
    [PDF] Energy Bands in Solids - Physics Courses
    ... where the potential V (r) is weak. This is known as the nearly free electron. (NFE) model. The matrix form of the Hamiltonian HGG′ (k) is given by. HGG′ (k) ...