Fact-checked by Grok 2 weeks ago

Electronic band structure

In , the electronic band structure describes the arrangement of allowed energy levels for electrons in a crystalline solid, resulting from their interaction with the periodic potential of the atomic lattice. This structure manifests as continuous ranges of energy, known as , separated by forbidden regions called band gaps, which fundamentally dictate the material's electronic properties. The concept is essential for understanding how solids conduct electricity, with band structures visualized in (k-space) to reveal the of energies as a function of the wave vector k. The theoretical foundation of electronic band structure is provided by , which states that the wavefunction of an in a periodic potential can be expressed as a modulated by a : ψ_k(r) = u_k(r) e^{i k · r}, where u_k(r) has the same periodicity as the . This theorem, derived from the for s in a crystal, implies that electron states are labeled by both an energy band index and a k within the first of the . The resulting band structure emerges from solving the periodic , often using approximations like the , where band gaps open at Brillouin zone boundaries due to Bragg scattering by the . The filling of these bands relative to the determines whether a behaves as a metal, , or . In metals, overlap or the conduction is partially filled, allowing electrons to move freely and conduct efficiently. feature a large (typically > 3–5 eV) between a fully occupied and an empty conduction , preventing electron excitation at . have a small (0.1–3 eV), enabling thermal or impurity-induced promotion of electrons to the conduction , which underpins their tunable in devices like transistors. Beyond basic classification, band structure influences advanced properties such as carrier effective mass (from band curvature) and optical responses, with direct versus indirect gaps affecting light emission and efficiency.

Origins and Fundamentals

Why bands and band gaps occur

In isolated atoms, electrons occupy discrete energy levels determined by quantum mechanical solutions to the . When multiple atoms are brought together to form a , the overlap of their wavefunctions perturbs these levels, causing each discrete atomic energy level to split into a of closely spaced states known as an energy band. This broadening occurs because the interaction between neighboring atoms allows electrons to delocalize across the , transforming sharp atomic orbitals into extended molecular-like orbitals that span a range of energies. The periodic potential of the lattice plays a central role in this splitting process. For two adjacent atoms, (such as the 1s orbitals in a pair of atoms) experience a from the interaction, lifting the degeneracy and producing two distinct levels: a lower-energy bonding state (symmetric wavefunction) and a higher-energy antibonding state (antisymmetric wavefunction). The magnitude of this splitting increases as the atoms approach each other and depends on the orbital type, with larger orbitals (e.g., 2p over 1s) showing greater overlap and thus larger separations. Extending this to a with N atoms (typically on the order of $10^{23}), each original atomic level splits into N sublevels, which become so closely spaced that they form a continuous , with the bandwidth scaling with the strength of interatomic interactions. The further drives band formation by prohibiting more than one electron per (considering ), ensuring that as electrons fill the lowest available states, they occupy progressively higher sublevels within the split bands rather than collapsing into a single . This , combined with the antisymmetry of the multi-electron wavefunction, forces the energy levels to spread out, creating the dense structure of bands observed in solids. Without it, all electrons could occupy the lowest state, precluding the extended electronic configurations essential for material properties like . Band gaps arise as forbidden energy regions between these bands, primarily due to the periodic lattice potential causing destructive interference of electron waves at the boundaries of the Brillouin zone. At these boundaries, where the wavevector k satisfies Bragg's condition (e.g., k = \pi/a for a one-dimensional lattice with spacing a), electron waves reflect off the lattice planes, forming standing waves. One standing wave has nodes at the atomic cores (higher energy, as electrons avoid ion cores), while the other has antinodes there (lower energy); the energy separation between these states creates the gap, blocking electron propagation at those energies. A concrete example is a one-dimensional chain of identical atoms separated by distance a, modeled via the tight-binding approximation. The electron energy disperses as E(k) = E_0 - 2t \cos(ka), where E_0 is the atomic on-site energy and t is the hopping amplitude between neighbors, forming a band ranging from E_0 - 2t to E_0 + 2t over the Brillouin zone k \in [-\pi/a, \pi/a]. A band gap opens at the zone boundary k = \pi/a (where E(k) = E_0) when the periodic potential is included, as the nearly degenerate states at k and k + 2\pi/a mix, pushing energies apart by an amount proportional to the potential strength. This phenomenon is captured quantitatively in the Kronig-Penney model, which treats a 1D periodic array of delta-function potentials V(x) = \sum_n P \delta(x - na), where P characterizes the potential strength. The resulting is \cos(ka) = f(E), with f(E) = \cos(\kappa a) + (P/\kappa a) \sin(\kappa a) and \kappa = \sqrt{2mE}/\hbar. Allowed energies correspond to |f(E)| \leq 1, while band gaps form where |f(E)| > 1, explicitly showing forbidden regions at zone edges.

Assumptions and limitations of band structure theory

The electronic band structure theory relies on several key assumptions to simplify the complex of s in solids. Foremost is the adiabatic approximation, or Born-Oppenheimer approximation, which assumes that the nuclei are fixed in position relative to the much faster-moving s, allowing the electronic wavefunction to be solved independently of nuclear motion. Another core assumption is the use of a single-particle , where each experiences an average field generated by the nuclei and all other s, enabling the treatment of s as quasiparticles in a periodic lattice. The theory further presumes an infinite, perfectly periodic crystal lattice without defects or boundaries, which justifies the application of and leads to continuous energy bands. These assumptions were foundational in Bloch's 1928 doctoral thesis, which introduced the quantum mechanical description of s in periodic potentials. A critical component is the independent electron approximation, which neglects direct electron-electron correlations beyond the mean-field level, treating interactions through an averaged potential rather than explicit many-body effects. This simplification allows the to be solved for individual electrons, forming the basis for band formation. However, it inherently limits the theory's accuracy in systems where correlations play a dominant role. Despite its successes, band structure theory has notable limitations stemming from these assumptions. It breaks down in strongly correlated systems, such as Mott insulators like , where electron-electron repulsion prevents metallic conduction despite partial band filling, as electron interactions exceed the mean-field description—a failure first highlighted by Nevill Mott in 1949. Basic implementations also neglect dynamic effects like electron-phonon coupling and spin-orbit interactions, which can significantly alter band dispersions and gaps in materials with strong lattice vibrations or heavy elements. The infinite crystal approximation further restricts applicability to low-dimensional or finite systems, such as nanowires or surfaces, where boundary effects disrupt periodicity. The theory performs best in weakly correlated metals and semiconductors but often fails in transition metals due to d-electron correlations. Although band theory correctly identifies insulators via band gaps, it typically underestimates these gaps by 30-50% in standard approximations like , necessitating advanced methods such as corrections for quantitative accuracy. Early limitations were recognized in the decades following Bloch's work, particularly as experimental discrepancies in correlated oxides emerged in the mid-20th century.

Core Concepts

Crystalline symmetry and Bloch wavevectors

The periodicity of the crystal imposes profound constraints on the electronic wavefunctions, fundamentally shaping the electronic band structure. In a crystalline solid, the potential experienced by electrons varies periodically with the translation vectors \mathbf{R}, such that V(\mathbf{r} + \mathbf{R}) = V(\mathbf{r}). This leads to , which asserts that the eigenfunctions of the can be expressed in the form \psi_{\mathbf{k}}(\mathbf{r}) = u_{\mathbf{k}}(\mathbf{r}) e^{i \mathbf{k} \cdot \mathbf{r}}, where u_{\mathbf{k}}(\mathbf{r}) is a with the periodicity, u_{\mathbf{k}}(\mathbf{r} + \mathbf{R}) = u_{\mathbf{k}}(\mathbf{r}), and \mathbf{k} is the Bloch wavevector. This form combines a plane-wave-like propagation modulated by the potential, ensuring compatibility with the crystal's . The Bloch wavevector \mathbf{[k](/page/K)} labels distinct electronic states and resides in the , which is defined by basis vectors \mathbf{b}_i satisfying \mathbf{a}_i \cdot \mathbf{b}_j = 2\pi \delta_{ij}, where \mathbf{a}_i are the real-space vectors. In one dimension, for a a, \mathbf{[k](/page/K)} spans the first Brillouin zone from -\pi/a to \pi/a, representing the unique set of wavevectors modulo reciprocal vectors. In three dimensions, this extends to the full reciprocal space, but due to the periodicity of the energy dispersion E(\mathbf{k} + \mathbf{G}) = E(\mathbf{k}) for any reciprocal vector \mathbf{G}, all states can be reduced to the first Brillouin zone without loss of information. This periodicity arises directly from Bloch's theorem, as shifting \mathbf{[k](/page/K)} by \mathbf{G} merely rephases the periodic part u_{\mathbf{k}}. The first is constructed as the Wigner-Seitz cell in reciprocal space, the region closest to the \Gamma (where \mathbf{k} = 0) bounded by planes perpendicular to lines connecting points and bisecting those lines. Zone boundaries correspond to points where Bragg reflection conditions are met, \mathbf{k} \cdot \mathbf{G} = n \pi for integer n, leading to degeneracies that split into band gaps under the periodic potential. Higher Brillouin zones are obtained by translating the first zone by vectors, and their contents are folded back into the first zone via equivalence under \mathbf{G}-shifts, a process analogous to Umklapp contributions in scattering that enforce the reduced description. In practice, high-symmetry points and lines within the dictate the locations of band extrema, exploiting the crystal's symmetries. For the face-centered cubic (FCC) lattice, common in semiconductors like and , key points include \Gamma at the zone center, X at the zone face centers (e.g., (2\pi/a)(0, 1/2, 1/2)), and L along the hexagonal faces (e.g., (2\pi/a)(1/2, 1/2, 1/2)); these often host conduction band minima or valence band maxima, as verified in detailed band structure calculations. This symmetry-reduced labeling simplifies the analysis of electronic properties across diverse crystal structures.

Density of states

The density of states, denoted as g(E), represents the number of electronic states available per unit energy interval dE between E and E + dE, normalized per unit volume of the material. This function is fundamental to characterizing the distribution of allowed electron energies in a solid and plays a key role in determining thermodynamic and transport properties. For free electrons in three dimensions, the can be derived by considering the uniform distribution of states in . The number of states in a of radius k and thickness dk is proportional to the volume $4\pi k^2 dk, divided by the volume per state (2\pi/L)^3 for a of side L, yielding a k-space density that transforms to via the E = \frac{\hbar^2 k^2}{2m}. Including a spin degeneracy factor of 2 for the two possible electron spin states, the result is g(E) \propto \sqrt{E}. In the presence of a periodic crystal potential, for band structure is obtained by integrating over the first . For a given E, it is given by summing over band indices n as g(E) = \sum_n \frac{1}{(2\pi)^3} \int_{\text{BZ}} d^3k \, \delta(E - E_n(\mathbf{k})), where the integral counts states where the band E_n(\mathbf{k}) matches E, and the factor $1/(2\pi)^3 arises from the normalization in reciprocal space per unit real-space volume (spin degeneracy is often included separately as a factor of 2). This expression generalizes the free-electron case by accounting for the folding of states into the due to the lattice periodicity. Due to the periodic potential, the energy surfaces E_n(\mathbf{k}) = constant can exhibit critical points where the gradient \nabla_{\mathbf{k}} E_n(\mathbf{k}) = 0, leading to singularities in g(E). These are known as Van Hove singularities, manifesting as peaks, steps, or divergences in the at energies corresponding to saddle points, minima, or maxima in the band dispersion. Such features arise from the reduced at these points, causing an accumulation of states. A specific example occurs near the conduction band minimum under the parabolic band approximation, where the is E(\mathbf{k}) = E_c + \frac{\hbar^2 k^2}{2m^*} with effective mass m^*. Here, the simplifies to g(E) = \frac{1}{2\pi^2} \left( \frac{2m^*}{\hbar^2} \right)^{3/2} \sqrt{E - E_c} for E > E_c, incorporating spin degeneracy and reflecting the quadratic energy-momentum relation. The provides the framework for applying Fermi-Dirac statistics to compute occupied states and derive properties such as electronic specific heat and electrical conductivity.

Band filling and the

In the at temperature, electrons in a crystalline solid occupy the lowest available energy states according to the , completely filling all quantum states up to the E_F, while leaving all states above E_F empty. The E_F thus represents the highest occupied and serves as the at T = 0 . The boundary defined by E_F in reciprocal space forms the Fermi surface, a constant-energy surface that generalizes the free-electron Fermi sphere to Bloch electrons in a periodic potential. This surface determines key electronic properties: in metals, partially filled bands cross the , enabling conduction, whereas in insulators, a band gap at E_F prevents electrons from reaching the empty states above, resulting in zero at low temperatures. At finite temperatures, the sharp step function of occupation at T = 0 K broadens due to thermal excitations, described by the Fermi-Dirac distribution function: f(E) = \frac{1}{e^{(E - \mu)/k_B T} + 1}, where \mu is the chemical potential (approximately equal to E_F for degenerate electron gases at low T), k_B is Boltzmann's constant, and T is the temperature. This probability function gives the average occupation of a state at energy E, approaching 1 for E \ll \mu and 0 for E \gg \mu. The total number of electrons N in the system is obtained by integrating the product of the g(E) and the Fermi-Dirac distribution over all energies: N = \int_{-\infty}^{\infty} g(E) f(E) \, dE, where g(E) quantifies the number of available states per unit energy interval. In semiconductors, the lies within the band gap for intrinsic materials, but doping shifts E_F toward the conduction or valence band edges depending on the type. Unlike the free-electron model, where continuous energy levels always yield metallic behavior regardless of electron density, the periodic potential in band theory introduces energy gaps that can fully fill bands below E_F, leading to insulating states when no states are available at the .

Classification of Bands

Valence and conduction bands

In electronic band structure, the valence band refers to the highest-energy band that is completely filled with electrons at temperature, primarily arising from molecular orbitals formed by the overlap of atomic orbitals in the crystal lattice. These bonding states contribute to the cohesive forces holding the solid together, with electrons occupying all available states up to the band's maximum energy. The conduction , situated immediately above the valence , is the lowest-energy that remains empty at T=0 K and consists mainly of antibonding or delocalized orbitals, allowing electrons in this to move freely through the and contribute to electrical . The energy separation between the maximum of the valence and the minimum of the conduction defines the gap, denoted as E_g, which represents a forbidden energy range for states in the ideal crystal. The relative positions of the with respect to the determine a material's electrical properties. In metals, these bands overlap or the lies within one of them, enabling high without external excitation. Semiconductors feature a modest of roughly 0.1 to 3 , positioning the within the gap and allowing limited thermal or optical promotion of electrons to the conduction band. Insulators, by contrast, exhibit large band gaps exceeding 3 , rendering electron excitation across the gap negligible under typical conditions. In intrinsic semiconductors, conduction arises from thermal excitation, where sufficient energy at finite temperatures bridges the band gap, elevating electrons from the valence band to the conduction band and leaving mobile holes in the valence band. This process generates equal numbers of electrons and holes, whose transport properties are characterized by effective masses derived from the band curvature near the respective band edges.

Direct and indirect band gaps

In semiconductors, the nature of the band gap—whether —depends on the relative positions in () of the maximum of the valence band and the minimum of the conduction band. A occurs when these extrema align at the same wavevector , such as at the Γ point (=0), enabling momentum-conserving electronic transitions without additional interactions. In contrast, an features these extrema at different -points, for example, the valence band maximum at Γ and the conduction band minimum near the X point, necessitating involvement to conserve during transitions. Gallium arsenide (GaAs) exemplifies a direct band gap material, with a gap energy of 1.42 eV at the Γ point, while silicon (Si) represents an indirect case, with a gap of 1.12 eV between the Γ valence maximum and a conduction minimum offset from Γ. The distinction profoundly affects optical processes: direct band gaps permit efficient radiative recombination and absorption via vertical transitions, as photons carry negligible momentum, whereas indirect gaps require phonon-assisted processes, which are less probable and suppress light emission and absorption. The transition rate for interband optical processes follows from . For direct transitions, momentum is approximately conserved since the photon momentum is negligible, allowing k = k'. For indirect transitions, a supplies the required momentum difference Δk = k - k', typically via a second-order process involving virtual intermediate states. This leads to practical implications, such as GaAs enabling high-efficiency light-emitting diodes (LEDs) through direct recombination, while Si's indirect gap results in poor luminescence efficiency despite its widespread use in electronics. The concepts of direct and indirect band gaps were theoretically and experimentally clarified in the 1950s through absorption spectroscopy studies, with Roger J. Elliott's 1957 work providing a foundational analysis of exciton-mediated direct and indirect transitions near the band edge.

Theoretical Models

Nearly free electron approximation

The nearly free electron approximation models the behavior of conduction electrons in a crystal lattice by treating them as weakly interacting with the periodic potential of the ion cores, starting from the free electron gas as the unperturbed system. The unperturbed energy dispersion for free electrons is given by the parabolic relation E(\mathbf{k}) = \frac{\hbar^2 k^2}{2m}, where \mathbf{k} is the electron wavevector, m is the effective electron mass, and \hbar is the reduced Planck's constant. The periodic lattice potential V(\mathbf{r}) is incorporated as a perturbation, expressed in its Fourier series form V(\mathbf{r}) = \sum_{\mathbf{G}} V_{\mathbf{G}} e^{i \mathbf{G} \cdot \mathbf{r}}, where the sum runs over reciprocal lattice vectors \mathbf{G} and V_{\mathbf{G}} quantifies the scattering strength of the potential for each \mathbf{G}. This approach captures how the lattice periodicity modifies the free electron states into Bloch waves without requiring strong binding assumptions. A crucial aspect of the model emerges at the boundaries, where free electron states with wavevectors \mathbf{k} and \mathbf{k} - \mathbf{G} are degenerate, having identical energies \frac{\hbar^2 k^2}{2m} = \frac{\hbar^2 |\mathbf{k} - \mathbf{G}|^2}{2m}. The off-diagonal matrix element V_{\mathbf{G}} of the lifts this degeneracy by coupling the states, resulting in an and the formation of an energy gap. At the zone boundary point \mathbf{k} = \mathbf{G}/2, the gap size is \Delta E = 2 |V_{\mathbf{G}}|, which represents the splitting between the upper and lower branches of the perturbed dispersion. This gap opening is derived using degenerate through the secular equation for the two relevant plane-wave basis states. The eigenvalue problem yields the condition: \det \begin{pmatrix} \frac{\hbar^2 k^2}{2m} - E & V_{\mathbf{G}} \\ V_{\mathbf{G}}^* & \frac{\hbar^2 |\mathbf{k} - \mathbf{G}|^2}{2m} - E \end{pmatrix} = 0, which, under the degeneracy condition, simplifies to the split energies E_{\pm}(\mathbf{k}) = \frac{\hbar^2 k^2}{2m} \pm |V_{\mathbf{G}}|. Away from the boundary, non-degenerate applies, with energy corrections of second order in V_{\mathbf{G}}. The approximation holds well for systems with weak potentials, such as the alkali metals, where the conduction electrons are loosely bound and experience minimal scattering from the . In , for instance, the model accurately describes the first at the boundary, arising from the smallest vector, with |V_{\mathbf{G}}| on the order of a few electronvolts consistent with experimental distortions. For higher-lying bands, the effects of multiple \mathbf{G} vectors can be incorporated through repeated applications of , though convergence requires including more terms as the potential strength increases.

Tight binding model

The tight binding model approximates the electronic wavefunctions in crystalline solids by assuming that the periodic potential localizes electrons strongly on sites, with interactions primarily through overlap of neighboring s. This approach, suitable for insulators and semiconductors where electrons are bound more tightly than in metals, constructs Bloch states as a (LCAO): \psi_{\mathbf{k}}(\mathbf{r}) = \sum_{\mathbf{R}} e^{i \mathbf{k} \cdot \mathbf{R}} \phi(\mathbf{r} - \mathbf{R}), where \phi(\mathbf{r}) is an centered at lattice site \mathbf{R}, and \mathbf{k} is the Bloch wavevector within the first . This form adheres to on the periodicity of crystal wavefunctions. The model's Hamiltonian is represented in the LCAO basis, featuring diagonal on-site energies \varepsilon_0 = \langle \phi | H | \phi \rangle that reflect the atomic potential at each site, and off-diagonal hopping integrals t_{ij} = \langle \phi_i | H | \phi_j \rangle that capture interactions between orbitals on sites i and j. The resulting band structure is determined by solving the secular equation for the eigenvalues E_n(\mathbf{k}) of the \mathbf{k}-dependent Hamiltonian matrix: H_{mn}(\mathbf{k}) = \sum_{\delta} e^{i \mathbf{k} \cdot \delta} \langle \phi_m(0) | H | \phi_n(\delta) \rangle, where \delta = \mathbf{R}_n - \mathbf{R}_m denotes lattice vectors connecting basis orbitals, and the sum is typically truncated to nearest or next-nearest neighbors for computational efficiency. This matrix diagonalization yields the energy bands E_n(\mathbf{k}) as a function of \mathbf{k}, providing a semi-empirical framework that interpolates between atomic levels and delocalized states. In the simplest case of a one-dimensional chain of atoms with a single s-orbital per site and nearest-neighbor hopping, the model reduces to a , producing the E(k) = \varepsilon_0 - 2t \cos(ka), where a is the lattice spacing and t < 0 is the hopping amplitude; this cosine form spans a bandwidth of $4|t| from the bottom of the band at k = 0 to the top at the zone boundary k = \pi/a. The model's versatility allows extensions to multi-orbital bases, incorporating s, p, and d orbitals with angular momentum-dependent hopping terms to describe directional bonding and splitting of degenerate levels into bands with varying symmetries. Multi-orbital hopping further refines the band topology, enabling accurate reproduction of features like band gaps and effective masses in covalent solids. A prominent example is the application to graphene, where the tight binding model using \pi (p_z) orbitals on the honeycomb lattice captures the characteristic Dirac cones at the K points, yielding a linear dispersion E(\mathbf{k}) \propto |\mathbf{k}| near these points and mimicking relativistic massless fermions with zero bandgap.

KKR model

The Korringa-Kohn-Rostoker (KKR) method is a multiple-scattering approach to solving the Schrödinger equation for the electronic structure of periodic solids, particularly suited to crystals with muffin-tin potentials. Developed in the late 1940s and early 1950s, it was first introduced by Jan Korringa in 1947, who formulated the problem in terms of Green's functions for Bloch waves in metals, and further refined by Walter Kohn and Norman Rostoker in 1954, who applied it to calculate the band structure of metallic lithium. This method laid the groundwork for modern all-electron electronic structure codes by providing an exact solution within the muffin-tin approximation, enabling accurate computations without adjustable parameters. In the KKR formalism, the crystal is divided into non-overlapping muffin-tin spheres centered at atomic sites, where the potential is treated as spherically symmetric and confined to these regions, and an interstitial region where free-electron propagation occurs. The Schrödinger equation is solved using scattering theory, where electron waves are expanded in terms of incoming and outgoing spherical waves at each atomic site, accounting for multiple scatterings between sites. The single-site t-matrix encapsulates the scattering properties of an isolated atom in its muffin-tin potential, relating the incoming and outgoing wave amplitudes for a given energy E. The interstitial propagation is described by structure constants, which are coefficients in the plane-wave expansion of the free-electron Green's function between different lattice sites. The band structure is determined by solving the secular equation, which sets the determinant of the difference between the inverse single-site t-matrix and the lattice Green's function matrix to zero: \det \left| t^{-1}(E) - G(\mathbf{k}) \right| = 0, where t(E) is the single-site t-matrix and G(\mathbf{k}) incorporates the structure constants summed over the reciprocal lattice for wavevector \mathbf{k}. This equation yields the allowed energies E(\mathbf{k}) for . The muffin-tin approximation assumes a constant potential in the interstitial region and zero overlap between spheres, making it particularly valid for close-packed metals where atomic densities are high and interstitial volumes are small. The KKR method excels in applications requiring precise density of states (DOS) calculations, such as for transition metals, where d-electron correlations demand all-electron treatments. For instance, full-potential implementations have demonstrated high accuracy in computing DOS for elements like vanadium and chromium, capturing fine structure in the d-bands that influences magnetic and transport properties. These capabilities stem from the method's Green's function foundation, which naturally integrates over the Brillouin zone to yield the DOS without explicit k-point sampling.

Density-functional theory

Density-functional theory (DFT) provides a foundational framework for computing electronic band structures in solids by mapping the complex many-electron problem onto an effective single-particle description. Developed as an ab initio method, DFT leverages the electron density as the central variable to determine ground-state properties, enabling efficient calculations of band energies and wavefunctions across the . This approach has become indispensable for predicting material properties, from semiconductors to metals, by solving self-consistent equations that approximate the effects of electron-electron interactions. The theoretical basis of DFT rests on the Hohenberg-Kohn theorems, which demonstrate that the ground-state electron density uniquely determines the external potential and, consequently, all ground-state properties of the system. The first theorem asserts that for a non-degenerate ground state, the external potential is a functional of the density, up to a constant shift, ensuring a one-to-one mapping between density and Hamiltonian. The second theorem introduces a variational principle: the true ground-state energy is the minimum of the energy functional over all possible densities, providing a practical route to optimization. These theorems, formulated for interacting electron systems in an external potential, underpin the density-based formulation of quantum many-body theory. To make DFT computationally tractable, Kohn and Sham introduced a reformulation that transforms the interacting system into a fictitious non-interacting one with the same density. The Kohn-Sham equations are a set of single-particle Schrödinger-like equations: \left[ -\frac{\hbar^2}{2m} \nabla^2 + V_{\text{eff}}(\mathbf{r}) \right] \psi_i(\mathbf{r}) = \varepsilon_i \psi_i(\mathbf{r}), where the orbitals \psi_i yield the density via \rho(\mathbf{r}) = \sum_i |\psi_i(\mathbf{r})|^2, occupied up to the number of electrons. The effective potential is V_{\text{eff}}(\mathbf{r}) = V_{\text{ext}}(\mathbf{r}) + V_{\text{H}}(\mathbf{r}) + V_{\text{xc}}(\mathbf{r}), comprising the external potential V_{\text{ext}} (e.g., from ions), the Hartree potential V_{\text{H}} accounting for classical Coulomb repulsion, and the exchange-correlation potential V_{\text{xc}} = \delta E_{\text{xc}} / \delta \rho. Solving these equations iteratively converges to the self-consistent density and eigenvalues \varepsilon_i, which approximate the single-particle energies forming the band structure. This mean-field approach assumes a single-particle picture, consistent with the underlying approximations in band theory. The exchange-correlation functional E_{\text{xc}}[\rho] captures all quantum many-body effects beyond the classical Hartree term, but its exact form is unknown, necessitating approximations. The local density approximation (LDA) assumes E_{\text{xc}}[\rho(\mathbf{r})] = \int \rho(\mathbf{r}) \epsilon_{\text{xc}}(\rho(\mathbf{r})) \, d\mathbf{r}, where \epsilon_{\text{xc}} is the uniform electron gas exchange-correlation energy per particle, parameterized from quantum Monte Carlo simulations. While LDA yields reasonable structural properties and densities, it systematically underestimates band gaps in semiconductors and insulators by 30-50%, as the resulting Kohn-Sham eigenvalues do not directly correspond to excitation energies. This band gap underestimation arises from the band gap problem in standard DFT approximations like and generalized gradient approximations (). The true fundamental gap equals the difference in Kohn-Sham eigenvalues plus a derivative discontinuity \Delta_{\text{xc}} in V_{\text{xc}} at integer electron numbers, reflecting the change in potential upon adding an electron. However, and lack this discontinuity due to their continuous dependence on density, leading to gaps approximated solely by \varepsilon_{LUMO} - \varepsilon_{HOMO}, which misses crucial many-body corrections. Post-2000 advancements have addressed these limitations through improved functionals. Hybrid functionals incorporate a fraction of exact Hartree-Fock exchange to better approximate E_{\text{xc}}, reducing self-interaction errors and improving gap predictions. The Heyd-Scuseria-Ernzerhof (HSE) screened hybrid functional, mixing 25% exact exchange with screened Coulomb interactions for efficiency in solids, yields band gaps within 0.2-0.3 eV of experiment for many semiconductors, outperforming LDA without excessive computational cost. Additionally, the scissor shift operator empirically adjusts LDA/GGA gaps by a constant offset derived from higher-level calculations or experiments, restoring accuracy for band alignments while preserving the overall structure. These developments have made DFT a reliable tool for band structure engineering in materials design.

GW approximation and Green's function methods

The Green's function approach provides a framework for calculating quasiparticle energies in electronic systems, going beyond ground-state methods like to address excited-state properties such as band gaps. The one-particle Green's function G(1,2;\omega), where 1 and 2 denote space-time coordinates and \omega is frequency, acts as a propagator describing the probability amplitude for adding an electron to the system at position 1 and detecting it at 2, or vice versa for removal. This function satisfies Dyson's equation: G = G_0 + G_0 \Sigma G, where G_0 is the non-interacting Green's function and \Sigma is the self-energy operator capturing many-body interactions. In the GW approximation, developed by Hedin in 1965, the self-energy is approximated as \Sigma(1,2;\omega) = i G(1,2;\omega) W(1^+,2;\omega), where W is the screened Coulomb interaction and the superscript + indicates a slight time shift to ensure causality. The screened interaction W is typically computed within the random phase approximation (RPA), which accounts for dielectric screening by treating electron-hole interactions perturbatively. This GW form represents a first-order expansion in W (the screened interaction beyond the bare Coulomb v) while using the full G or an approximate G_0, making it a many-body perturbation theory suitable for weakly correlated systems. Quasiparticle energies, which approximate the band structure for low-energy excitations, are obtained by solving the Dyson-like quasiparticle equation: [T + V_\text{ext} + V_H] \psi + \langle \psi | \Sigma(E_\text{qp}) - V_\text{xc} | \psi \rangle \chi = E_\text{qp} \psi, where T is the kinetic energy operator, V_\text{ext} the external potential, V_H the Hartree potential, V_\text{xc} the exchange-correlation potential from , and \chi a normalization factor. This equation corrects band structures by replacing the approximate V_\text{xc} with the energy-dependent \Sigma(E_\text{qp}), particularly improving underestimated band gaps in semiconductors. For example, in silicon, the local-density approximation (LDA) within yields an indirect band gap of approximately 0.7 eV, while ab initio increase it to about 1.2 eV, closely matching the experimental value of 1.17 eV. Despite its accuracy for quasiparticle spectra in many materials, the GW approximation has significant limitations due to its perturbative nature and high computational cost. The standard implementation scales as O(N^4) with system size N (from the evaluation of W), restricting applications to small systems or requiring approximations like G_0W_0, where G_0 is taken from a DFT starting point without self-consistency in G or \Sigma. Self-consistent GW variants improve convergence but exacerbate the cost, often making full-frequency treatments challenging for complex solids.

Dynamical mean-field theory

Standard band theory, such as density-functional theory in its local approximations, often fails to describe strongly correlated electron systems like Mott-Hubbard insulators, where local electron-electron interactions lead to phenomena such as metal-insulator transitions that cannot be captured by single-particle pictures. Dynamical mean-field theory (DMFT) addresses this limitation by providing a non-perturbative treatment of local correlations in lattice models of interacting electrons. Developed in the early 1990s, DMFT becomes exact in the limit of infinite spatial dimensions, where nonlocal correlations are suppressed, allowing the self-energy to be approximated as momentum-independent and purely local, Σ(k, ω) ≈ Σ(ω). In DMFT, the interacting lattice problem is mapped onto an effective single-site quantum impurity model, such as the , where the impurity represents a lattice site coupled to a self-consistent bath that mimics the rest of the lattice. The local self-energy is obtained by solving this impurity problem using numerical methods, including (QMC) techniques like the or continuous-time QMC, exact diagonalization, or the non-crossing approximation. Self-consistency is enforced by requiring that the local lattice matches the impurity Green's function, leading to the lattice Green's function expressed as G(\mathbf{k}, \omega) = \frac{1}{\omega + \mu - \varepsilon_{\mathbf{k}} - \Sigma(\omega)}, where μ is the chemical potential, ε_k is the non-interacting dispersion, and ω is the frequency. This formalism captures dynamical fluctuations in the local charge and spin degrees of freedom, enabling the study of spectral properties and thermodynamics beyond static mean-field approximations. A primary application of DMFT is to the single-band Hubbard model, which includes nearest-neighbor hopping t and on-site Coulomb repulsion U, H = -t ∑{} (c{iσ}^† c_{jσ} + h.c.) + U ∑i n{i↑} n_{i↓}. At half-filling and low temperatures, DMFT predicts a Mott metal-insulator as U increases, with a critical value U_{c2} ≈ 2.9W (where W is the ) marking the onset of the insulating phase at T=0 via a second-order , while at finite temperatures the becomes with a coexistence region between U_{c1} ≈ 2.2W and U_{c2}. In the metallic phase, the spectral weight Z ∝ (1 - (U/U_{c2})^2) vanishes at the , reflecting the breakdown of Fermi liquid behavior due to strong local correlations. DMFT was formalized in the early through foundational works establishing its mapping and solvability in infinite dimensions. To apply it to realistic materials, DMFT is combined with density-functional theory (DFT) in the DFT+DMFT framework, where correlated orbitals (e.g., d or f electrons) are treated with DMFT while others use DFT, often via (LDA+DMFT). This hybrid approach has successfully described the electronic structure of δ-plutonium, reproducing its large specific heat and volume under strong correlations in the 5f shell without . Similarly, it elucidates the physics of high-temperature superconductors like cuprates, capturing the pseudogap and doping-dependent Mott features in materials such as La_{2-x}Sr_xCuO_4. Recent advances in DMFT focus on extensions beyond the infinite-dimensional limit to better handle finite-dimensional systems and incorporate nonlocal correlations, which are crucial for low-dimensional or frustrated lattices. Methods such as cluster dynamical mean-field theory (CDMFT) treat short-range spatial correlations by embedding finite clusters, while diagrammatic extensions like the dual fermion or TRILEX approaches systematically include nonlocal effects starting from the local DMFT solution. These developments enable more accurate predictions for two-dimensional systems, such as the on square lattices, where nonlocal fluctuations modify the Mott transition and superconducting instabilities.

Other advanced methods

Hybrid functionals address some limitations of standard (DFT) by incorporating a fraction of exact Hartree-Fock exchange, improving the accuracy of predictions without the high computational cost of many-body methods like . Popular examples include B3LYP, which mixes 20% exact exchange with Becke's gradient-corrected exchange and Lee-Yang-Parr correlation, and PBE0, which uses 25% exact exchange with the Perdew-Burke-Ernzerhof functional; these have been shown to yield s for semiconductors and insulators that are closer to experimental values, often within 0.2-0.5 error for diverse materials. Such approaches balance accuracy and efficiency, making them suitable for routine electronic structure calculations in solids. The Bethe-Salpeter equation (BSE) extends approximations by explicitly treating electron-hole interactions, enabling the computation of excitonic effects that are crucial for interpreting optical absorption spectra in semiconductors. In BSE, the optical response is obtained by diagonalizing a that includes screened attraction and exchange terms between states, often resulting in bound excitons with binding energies of 0.1-1 eV in materials like or transition metal dichalcogenides. This method has become standard for predicting redshifted optical peaks and in low-dimensional systems. Time-dependent density functional theory (TDDFT), particularly in its linear response formulation via the Casida equations, provides an efficient framework for calculating vertical excitation energies and oscillator strengths from ground-state band structures. However, TDDFT with common semilocal or often underestimates charge-transfer excitations due to the lack of exact exchange and long-range correction, leading to errors exceeding 1 eV in donor-acceptor systems or wide-gap insulators. Despite these limitations, it remains widely used for rapid screening of in molecular crystals and nanostructures. Since the , potentials have accelerated band structure calculations by surrogate modeling of DFT energies and forces, enabling high-throughput exploration of materials with reduced computational overhead by orders of magnitude. For instance, neural network-based approximate the to predict electronic properties directly, while approaches like Boltzmann generators use normalizing flows to sample Boltzmann distributions of atomic configurations, facilitating efficient computation of thermodynamic averages relevant to band alignments in alloys. Topological invariants, such as the Chern number, further characterize band topology by integrating the Berry curvature over the , quantifying nontrivial phases like the quantum anomalous Hall state in ferromagnetic insulators with nonzero integer values dictating edge conductance. Relativistic effects, primarily through spin-orbit coupling, modify band structures by splitting degenerate states and inducing band inversions that form Dirac-like cones at high-symmetry points, as seen in topological insulators like Bi_2Se_3 where strong (on the order of 0.3-1 eV) protects helical surface states. In 2025, emerging methods for localization from DFT data have streamlined the generation of tight-binding Hamiltonians, allowing scalable of band structures across large k-point grids with minimal loss in accuracy for complex materials.

Extensions Beyond Ideal Crystals

Band structures in non-crystalline solids

In non-crystalline solids, such as amorphous materials, , and polymers, the absence of long-range periodic order eliminates the applicability of the Bloch theorem, which underpins the sharp wavevector labeling of electronic states in crystals. Consequently, electronic states cannot be precisely characterized by a well-defined crystal k, leading to a where states are instead classified as extended or localized based on their spatial extent and transport properties. This disorder-induced modification fundamentally alters the band structure concept, shifting focus from periodic bands to a continuous influenced by short-range atomic correlations. A key feature in these systems is the distinction between extended states, which enable metallic-like conduction, and localized states, which contribute to insulating behavior. Anderson localization arises when sufficient disorder confines electron wavefunctions to finite regions, preventing diffusion and transport; this phenomenon was first theoretically demonstrated in random lattices where disorder strength exceeds a critical threshold, resulting in exponentially decaying wavefunctions. In amorphous solids, this localization typically affects states near the band edges or within the gap, while higher-energy states may remain extended, preserving some band-like conductivity at elevated temperatures or carrier densities. Mobility edges demarcate the transition between these regimes, serving as energy thresholds in the that separate extended states (above the edge, allowing delocalized transport) from localized ones (below, impeding conduction). Proposed in the context of disordered semiconductors, these edges explain the energy-dependent metal-insulator transition observed in materials with moderate disorder, where the mobility edge position shifts with increasing randomness, narrowing the effective conducting region. The in non-crystalline solids features broadened bands compared to crystals, with exponential tails extending into the band gap due to structural . These Urbach tails, characterized by an exponential form in the , arise from fluctuations in local bonding environments and have a slope parameter () typically around 50 meV in amorphous semiconductors, reflecting the degree of . Unlike sharp band edges in periodic systems, these tails introduce sub-gap states that enhance recombination but can be partially passivated in hydrogenated variants. A representative example is (a-Si), where the optical is approximately 1.7-1.8 eV—larger than the 1.12 eV indirect gap in —due to increased sp³ bonding strain and reduced overlap from distorted lattices, though the overall valence-conduction band separation remains conceptually similar. However, a-Si exhibits a higher density of defect states within the gap (around 10^{16}-10^{17} cm^{-3} eV^{-1}), arising from dangling bonds and contributing to broader Urbach tails with energies of 40-60 meV, which limit carrier mobility compared to the crystalline form. Theoretical modeling of these structures often employs the coherent potential approximation (), which treats as an effective average potential on lattice sites to compute the averaged and . Originally developed for substitutional alloys, CPA extends naturally to amorphous systems with alloy-like positional or on-site , providing a single-site mean-field description that captures broadening and tail formation without requiring full structural simulation; for instance, it predicts mobility edges in moderately disordered semiconductors by solving self-consistent equations for the coherent scattering potential. Such band concepts are particularly relevant for applications in and polymers, where the lack of a precludes k-space analysis, yet short-range order—such as chain conjugation in polymers or tetrahedral coordination in glasses—maintains pseudo-band features like a finite near the and tailing into gaps. In chalcogenide , for example, this leads to tunable gaps of 1-3 eV with localized lone-pair states dominating the valence band tail, enabling switchable . In conjugated polymers, π-orbital overlap yields effective bandwidths of several eV, supporting charge despite amorphous chain packing.

Effects of defects and doping

Defects in crystalline semiconductors disrupt the periodic potential, introducing localized energy states within the band gap that alter the overall electronic band structure. Point defects, such as vacancies or atoms, typically create levels positioned near the middle of the band gap, acting as recombination centers or traps for charge carriers. These defect states arise from strong distortions and dangling bonds, which localize electrons or holes without significant coupling to the extended band states. For instance, in III-V semiconductors like GaAs, aluminum vacancies (V_Al) introduce multiple levels across the band gap, with transition energies around mid-gap, influencing carrier lifetimes and device performance. In contrast, controlled introduction of impurities through doping generates shallow defect levels close to the band edges, enabling tunable conductivity. In n-type doping, donor impurities (e.g., group V elements in group IV hosts) contribute an extra valence electron, forming a shallow donor level just below the conduction band minimum; this electron can be thermally excited into the conduction band, increasing free electron density. Similarly, p-type doping with acceptor impurities (e.g., group III elements) creates a shallow acceptor level above the valence band maximum, accepting an electron from the valence band and generating holes. The energy of these shallow levels is small compared to the band gap, typically on the order of 0.01–0.1 eV at room temperature, allowing efficient ionization. A representative example is phosphorus doping in silicon, where the donor level lies 0.045 eV below the conduction band minimum, facilitating n-type conduction with the Fermi level shifting toward the conduction band. The of shallow donors can be approximated using the hydrogenic model within the effective approximation, treating the as a screened potential analogous to the but scaled by the host material's properties: E_d = -\frac{13.6 \, \mathrm{eV} \left( \frac{m^*}{m_0} \right)}{\epsilon_r^2} Here, E_d is the donor relative to the conduction band, m^* is the effective , m_0 is the , and \epsilon_r is the relative constant of the . This model provides a reasonable estimate for many materials; for example, it yields approximately 0.026 eV for in using m^* \approx 0.26 m_0 and \epsilon_r \approx 12, while the measured is 0.045 eV due to multi-valley effects and central-cell corrections. At higher doping concentrations (typically >10^{18} cm^{-3}), the discrete impurity levels broaden into impurity bands due to wavefunction overlap, leading to band tailing where exponential tails extend into the band gap. These tails arise from potential fluctuations caused by the random distribution of impurities, merging the impurity band with the host conduction or valence band and reducing the effective band gap. In heavily n-doped semiconductors, this results in degenerate conditions where the enters the conduction band, enhancing metallic-like behavior. A notable consequence of heavy n-type doping is the Burstein-Moss effect, where the blocks low-energy transitions near the conduction band minimum due to filled states, effectively blue-shifting the optical edge. This apparent increase in is proportional to the carrier density and has been observed in materials like InSb, where high concentrations raise the onset by several hundred meV. The effect is particularly relevant in degenerate semiconductors, linking directly to band filling and positioning.

Visualization and Interpretation

Band diagrams

Band diagrams are graphical representations of the electronic band structure, obtained by plotting the energy eigenvalues E_n(\mathbf{k}) for different bands n as a function of the wavevector \mathbf{k} along predefined paths in the . These paths typically connect high-symmetry points, such as \Gamma, X, L, and back to \Gamma for face-centered cubic (FCC) crystals, to capture the essential features of the band dispersion while exploiting the symmetry of the . The choice of path ensures that the diagram reveals critical points like band extrema and crossings, providing insight into the material's electronic properties without needing to plot the full three-dimensional . Key features visible in band diagrams include the dispersion of bands, which describes how varies with , and gaps between bands. Band dispersion is quantified by the , which relates to the effective of charge via m^* = \frac{\hbar^2}{\frac{d^2 E}{dk^2}}, where a larger (steeper ) corresponds to a smaller effective and higher . In standard plots, the axis is conventionally set to zero at the E_F, allowing clear distinction between occupied (below E_F) and unoccupied (above E_F) states, with multiple bands shown to illustrate overlaps or separations. For a simple metal, the band diagram exhibits free-electron-like parabolic dispersions folded back into the first due to the periodic lattice potential, resulting in overlapping bands that cross E_F. In direct-bandgap semiconductors, the diagram shows a clear at the \Gamma point where both the valence band maximum and conduction band minimum occur, with no states at E_F. In indirect-bandgap semiconductors, the conduction band minimum is offset in from the \Gamma point. Such diagrams are routinely generated from computational outputs of codes like and , which perform non-self-consistent calculations along specified k-paths to produce eigenvalues for plotting tools. Interpretation of band diagrams focuses on dispersion characteristics: flat bands, with minimal variation over k, indicate localized electrons confined by strong potential variations, leading to low ; steep bands, resembling free-electron behavior, signify delocalized carriers with high .

Examples of band structures in materials

In metals such as (Cu), the electronic band structure closely resembles that of a nearly gas, characterized by a nearly spherical that touches the boundary at the L points along the ⟨111⟩ directions. This distortion from the ideal free-electron sphere arises due to the weak periodic potential of the crystal lattice, leading to energy gaps at the zone boundaries while maintaining high from overlapping bands at the . Semiconductors exhibit diverse band structures depending on their gap type. In (GaAs), a III-V compound, the fundamental bandgap is direct, with both the valence band maximum and conduction band minimum located at the Γ point in the . The conduction band edge at Γ is predominantly s-like, contributing to a low effective mass for electrons (~0.067 m₀) and efficient optical transitions for light emission. In contrast, silicon (Si), an elemental group-IV semiconductor, features an indirect bandgap of approximately 1.12 eV at 300 K, where the conduction band minimum occurs near the X point along the Δ line (about 85% from Γ to X), while the valence band maximum remains at Γ. This misalignment in k-space hinders direct radiative recombination, influencing Si's primary use in over . Insulators display wide bandgaps that prevent conduction under typical conditions. Diamond (C), a group-IV elemental solid, has a large indirect bandgap of 5.47 , with the conduction band minimum along the Δ direction near the X point and a deep valence band maximum at Γ. The substantial gap and strong covalent bonding result in exceptional electrical alongside high thermal conductivity, making diamond a for wide-bandgap materials. Topological insulators represent an emerging class with insulating bulk states but conducting surface states. In bismuth selenide (Bi₂Se₃), the bulk exhibits a nontrivial bandgap of ~0.3 eV, insulating the interior, while the surface hosts gapless Dirac cones with linear dispersion, enabling spin-momentum-locked helical states. Two-dimensional materials offer unique band topologies. Graphene, a single layer of carbon atoms in a honeycomb lattice, features a zero bandgap with linear dispersion relations E = ±ℏv_F |k| near the K and K' points at the Brillouin zone corners, forming Dirac cones that impart massless Dirac fermion behavior to charge carriers. Computed band structures from methods like are routinely validated against experimental measurements, particularly (ARPES), which directly maps momentum-resolved occupied states and confirms key features such as gap locations and dispersions in materials from to .

References

  1. [1]
    [PDF] Chapter 7: The Electronic Band Structure of Solids
    Mar 2, 2017 · As we bring more atoms together (left) or bring the atoms in the lattice closer together (right), bands form from mixing of the orbital states. ...
  2. [2]
    [PDF] Band Structures and the Meaning of the Wave Vector k - UCSB MRL
    Band structures are a representation of the allowed electronic energy levels of solid materials and are used to better inform their electrical properties.
  3. [3]
    [PDF] Energy Bands in Solids - Physics Courses
    ELECTRONIC BAND STRUCTURE OF CRYSTALS called the band index. Thus ... Then we can define the electron filling factor ν = eNe/N, where eNe is the total ...
  4. [4]
    Band Theory for Solids - HyperPhysics
    Band theory visualizes electron energies as bands, not discrete, and the Fermi level is key. Insulators have large gaps, conductors overlap bands, and ...Missing: definition | Show results with:definition
  5. [5]
  6. [6]
  7. [7]
    Band structure
    The gaps occur at Brillouin zone boundaries (BZBs). This is because electron waves with wavevectors at BZBs satisfy Bragg's law for reflection. Then the ...Missing: destructive | Show results with:destructive
  8. [8]
    [PDF] 2. Band Structure - DAMTP
    Solids are collections of tightly bound atoms. For most solids, these atoms arrange themselves in regular patterns on an underlying crystalline lattice.
  9. [9]
    Quantum mechanics of electrons in crystal lattices - Journals
    With this model an interpretation of the specific heat, the electrical and thermal conductivity, the magnetic susceptibility, the Hall effect, and the optical ...
  10. [10]
    [PDF] Condensed matter physics
    Explain the Bloch theorem and calculate the electronic band structure ... In Born-Oppenheimer approximation, electron and ion dynamics is separated ...
  11. [11]
    [PDF] Electronic Band Structure Calculations - Exstrom Laboratories LLC
    Dec 1, 1998 · other electrons in the system as well as the potential due to the nuclei. This is known as the independent electron approximation. For a ...
  12. [12]
    [PDF] 3. Electron Dynamics in Solids kz k k
    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 ...
  13. [13]
    Why is there a band structure for strongly correlated systems?
    Aug 19, 2015 · The existence of band structure of a crystalline solid comes from the Bloch theorem, which relies on the independent-electron approximation.
  14. [14]
    Strongly Correlated Materials: Insights From Dynamical Mean-Field ...
    Mar 1, 2004 · The failure of band theory was first noticed in insulators such as ... theory of the Mott transition with experimental observations.
  15. [15]
    Electronic structure calculations with dynamical mean-field theory
    Aug 14, 2006 · A review of the basic ideas and techniques of the spectral density-functional theory is presented. This method is currently used for ...
  16. [16]
    [PDF] Efficient Band Gap Prediction for Solids - Ceder Group
    Nov 5, 2010 · Typically, EKS underestimates band gaps (EExp) of solids by 30%–100 ... Martin, Electronic Structure: Basic Theory and. Practical ...
  17. [17]
    Sir Nevill F. Mott – Facts - NobelPrize.org
    By observing the interaction between electrons, Nevill Mott explained in 1949 how certain crystals can alternate between being electrical conductors and ...
  18. [18]
    [PDF] DAMTP - 2. Band Structure
    The statement of Bloch's theorem is that uk(x) has the periodicity of Λ which is indeed true, since uk(x + r) = e ik·xe ik·rψk(x + r) = e ik·xψk(x) = uk(x).
  19. [19]
    [PDF] ECE606: Solid State Devices Lecture 4 - Purdue Engineering
    For 2D/3D systems, energy-bands are often difficult to visualize. E-k diagrams along specific direction and constant energy surfaces for specific bands ...
  20. [20]
    Brillouin zone - Online Dictionary of Crystallography
    Nov 9, 2017 · A Brillouin zone is a particular choice of the unit cell of the reciprocal lattice. It is defined as the Wigner-Seitz cell (also called ...
  21. [21]
    [PDF] Energy bands - Rutgers Physics
    Equation (14) is known as Bloch theorem, which plays an important role in electronic band structure theory. Now we discuss a number of important conclusions ...
  22. [22]
    [PDF] Energy Bands in Solids - Physics Courses
    4.8 Atomic energy levels and crystalline energy bands . ... Each energy level is doubly degenerate, corresponding to two electron spin polarizations.
  23. [23]
    [PDF] Density of States Derivation
    The density of states gives the number of allowed electron (or hole) states per volume at a given energy. It can be derived from basic quantum mechanics.
  24. [24]
    [PDF] 6.730 Physics for Solid State Applications
    Specific heat is independent of temperature…NOT TRUE ! To get this correct we will need to (a) quantize electron energy levels, (b) introduce discreteness ...
  25. [25]
    [PDF] Density of States: - Georgia Institute of Technology
    Dec 17, 2005 · Derivation of Density of States (2D). Recalling from the density of states 3D derivation… k-space volume of single state cube in k-space: k ...
  26. [26]
    The Occurrence of Singularities in the Elastic Frequency Distribution ...
    The Occurrence of Singularities in the Elastic Frequency Distribution of a Crystal. Léon Van Hove ... Phys. Rev. 89, 1189 – Published 15 March, 1953. DOI ...
  27. [27]
    [PDF] Electrons and Holes in Semiconductors
    Naturally, the electrons tend to fill up the low energy bands first. The lower the energy, the more completely a band is filled. In a semiconductor, most of ...
  28. [28]
    [PDF] Lecture 13: Metals
    Fermions satisfy the Pauli-exclusion principle: no two fermions can occupy the same state. This makes fermionic systems act very di erently from bosonic ...
  29. [29]
    [PDF] Section 10 Metals: Electron Dynamics and Fermi Surfaces
    The set of all such surfaces is known as the Fermi surface, and is the generalization to Bloch electrons of the free electron Fermi sphere.
  30. [30]
    [PDF] Band Structure and Electrical Conductivity in Semiconductors
    Figure 1: Simplified diagram of the electronic band structure of insulators, semiconductors and metals. The position of the Fermi level is when the sample is at ...
  31. [31]
    [PDF] Carrier concentration Contents 1 Density of States - GitHub Pages
    Nov 14, 2020 · The density of states for electrons in a band yields the number of states in a certain energy range. This function is important in ...
  32. [32]
    Band Theory of Solids – University Physics Volume 3
    The free electron model explains many important properties of conductors but is weak in at least two areas. First, it assumes a constant potential energy ...<|control11|><|separator|>
  33. [33]
    [PDF] Lecture 14: Semiconductors
    When N orbitals combine, the many bonding and antibonding orbitals merge into bands, ... above it the conduction band and the band below it the valence band.
  34. [34]
    [PDF] BANDS AND BONDS
    The band of empty or antibonding states is called the conduction band. The highest energy occupied states are separated from the lowest energy unoccupied states ...
  35. [35]
    [PDF] Compilation of Energy Band Gaps in Elemental and Binary ...
    Oct 29, 2009 · in the valence band and the minimum energy in the conduction band. If Eg is small (0-3 or 4 eV) a material is considered to be a ...
  36. [36]
    [PDF] Lecture 12 - MIT OpenCourseWare
    “Solid State Physics”, Ashcroft and Mermin ... For some materials it turns out that the valence band maximum does not coincide with the conduction band minimum in ...
  37. [37]
    Intrinsic Carrier Concentration - PVEducation
    A large band gap will make it more difficult for a carrier to be thermally excited across the band gap, and therefore the intrinsic carrier concentration is ...
  38. [38]
    Direct and Indirect Band Gap Semiconductors - DoITPoMS
    In a direct band gap semiconductor, the top of the valence band and the bottom of the conduction band occur at the same value of momentum.
  39. [39]
    [PDF] I. GaAs Material Properties
    GaAs is a III-V semiconductor with a zinc blende structure, a 1.42 eV band gap, and is a direct band gap semiconductor. It has a density of 5.32 g/cm3.
  40. [40]
    [PDF] Band structure in semiconductors: – Direct/indirect bandgaps
    The conduction band electron effective mass has a strong dependence on the bandgap, the smaller the bandgap the smaller the effective mass.<|separator|>
  41. [41]
    Band Gap – dielectrics, semiconductors, metals, energy, electronic ...
    Examples of direct band gap semiconductor materials are gallium arsenide (GaAs), indium gallium arsenide (InGaAs), gallium nitride (GaN), aluminum nitride (AlN ...
  42. [42]
    Introduction to Light Emitting Diodes - Evident Scientific
    For a recombination event to occur in indirect gap materials, an electron must change its momentum before combining with a hole, resulting in a significantly ...
  43. [43]
    Intensity of Optical Absorption by Excitons | Phys. Rev.
    If the extrema are at different K values, indirect transitions involving phonons occur, giving absorption proportional to ( 𝛥 ⁢ E ) 1 2 for each exciton band, ...
  44. [44]
    Les électrons libres dans les métaux et le role des réflexions de Bragg
    L'article se termine par quelques tableaux numériques sur les métaux vrais, leurs réseaux, leurs électrons « libres » et le rôle que peuvent encore y jouer les ...
  45. [45]
    [PDF] Über die Quantenmechanik der Elektronen in Kristallgittern
    Von Felix Bloeh in Leipzig. 5lit 2 Abbildungen. (Eingegangen am 10. August 1928.) Die Bewegung eines Elektrons im Gitter wird untersucht, indem wir uns dieses.
  46. [46]
    [PDF] Simplified LCAO Method for the Periodic Potential Problem
    The LCAO, or Bloch, or tight binding, approximation for solids is discussed as an interpolation method, ... his original paper, only computed the diagonal matrix.
  47. [47]
    [PDF] The Tight Binding Method - Rutgers Physics
    May 7, 2015 · The tight binding method is a semi-empirical method used to calculate band structure and single-particle states, expanding Bloch states using ...Missing: seminal paper
  48. [48]
    [PDF] The Band Theory of Graphite
    The band theory of graphite uses 'tight binding' to explain its properties, treating it as a semiconductor with zero activation energy, and one conduction ...
  49. [49]
    On the calculation of the energy of a Bloch wave in a metal
    August 1947, Pages 392-400. Physica. On the calculation of the energy of a ... J. Korringa, R. Kronig, A. Smit. Physica, 11 (1945), p. 209. View PDFView ...
  50. [50]
    Solution of the Schrödinger Equation in Periodic Lattices with an ...
    Solution of the Schrödinger Equation in Periodic Lattices with an Application to Metallic Lithium. W. Kohn and N. ... Phys. Rev. 94, 1111 – Published 1 June ...
  51. [51]
    Screened full-potential KKR calculations for transition metals, based ...
    We show the accuracy of the SFPKKR method combined with the generalized-gradient approximation of density-functional theory. The present calculations ...
  52. [52]
    Inhomogeneous Electron Gas | Phys. Rev.
    This paper deals with the ground state of an interacting electron gas in an external potential v ⁡ ( r ) . It is proved that there exists a universal ...
  53. [53]
    Self-Consistent Equations Including Exchange and Correlation Effects
    Self-Consistent Equations Including Exchange and Correlation Effects. W. Kohn and L. J. Sham. University of California, San Diego, La Jolla, California. PDF ...
  54. [54]
    Exchange-correlation functionals for band gaps of solids - Nature
    Jul 10, 2020 · We conducted a large-scale density-functional theory study on the influence of the exchange-correlation functional in the calculation of electronic band gaps ...
  55. [55]
    Density functional theory and the band gap problem - Perdew - 1985
    Mar 23, 1985 · This functional leads to more accurate ground-state energies and densities than those of the LDA with little change in the calculated band ...
  56. [56]
    New Method for Calculating the One-Particle Green's Function with ...
    The new method derives self-consistent equations for the one-electron Green's function, using a screened potential, and is numerically manageable.
  57. [57]
    The GW Compendium: A Practical Guide to Theoretical ... - Frontiers
    GW is an approximation to an exact set of coupled integro-differential equations called Hedin's equations (Hedin, 1965), the full derivation of which can be ...
  58. [58]
    Electron correlation in semiconductors and insulators: Band gaps ...
    Oct 15, 1986 · In particular, the indirect band gap is calculated as 5.5, 1.29, and 0.75 eV as compared with experimental gaps of 5.48, 1.17, and 0.744 eV for ...
  59. [59]
    Ab initio Green's function formalism for band structures | Phys. Rev. B
    Nov 17, 2005 · It is based on the algebraic diagrammatic construction scheme for the self-energy which is formulated for crystal orbitals (CO-ADC).
  60. [60]
    Nonlocal corrections to dynamical mean-field theory from the two ...
    Nov 3, 2022 · Here, we benchmark an approach that combines dynamical mean-field theory (DMFT) observables with the two-particle self-consistent theory (TPSC).Missing: 2020-2025 extensions finite
  61. [61]
    Range-separated hybrid functionals for accurate prediction of band ...
    Jun 21, 2023 · In this work, we systematically evaluate the accuracy in band gap prediction of range-separated hybrid functionals on a large set of semiconducting and ...
  62. [62]
    Non-adiabatic approximations in time-dependent density functional ...
    Jul 13, 2023 · On the other hand, in the vast majority of cases, TDDFT is applied in the linear response regime, where weak perturbations of the ground-state ...
  63. [63]
    Solving the electronic structure problem with machine learning
    Feb 18, 2019 · Here we introduce a machine-learning-based scheme to efficiently assimilate the function of the KS equation, and by-pass it to directly, rapidly, and ...Missing: post- | Show results with:post-
  64. [64]
    Predicting electronic structures at any length scale with machine ...
    Jun 27, 2023 · We have developed a machine learning framework for predicting the electronic structure on any length scale. It shows up to three orders of magnitude speedup.Results · Density Functional Theory · Dft Surrogate Models
  65. [65]
    [2307.16258] Topological electronic bands in crystalline solids - arXiv
    Jul 30, 2023 · This review provides a gentle but firm introduction to topological electronic band structure suitable for new researchers in the field.
  66. [66]
    A machine learning route between band mapping and band structure
    Dec 30, 2022 · The electronic band structure and crystal structure are the two complementary identifiers of solid-state materials.
  67. [67]
    Electronic Structures of Amorphous Solids - SpringerLink
    The nature and behaviour of electron states in amorphous solids are reviewed. The usefulness of the density of states as a description of electron states.
  68. [68]
    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 ...
  69. [69]
    The mobility edge since 1967 - IOPscience
    A mobility edge is defined as the energy separating localised and non-localised states in the conduction or valence bands of a non-crystalline material.
  70. [70]
    Coherent-Potential Model of Substitutional Disordered Alloys
    We introduce a model of a substitutional alloy based on the concept of an effective or coherent potential which, when placed on every site of the alloy lattice,
  71. [71]
    Vacancies and defect levels in III–V semiconductors - AIP Publishing
    Aug 13, 2013 · The V Al q defect level transitions occur at or near the middle of the band gap implying that they are all deep level traps. The point group ...
  72. [72]
    Electrical Conduction in Semiconductors - MKS Instruments
    This dopant electron can easily jump the small energy gap (0.045 eV for P, much less than the 1.1 eV band gap for pure silicon at room temperature; typical ...
  73. [73]
    Theory of Donor States in Silicon | Phys. Rev.
    This error is attributed to failure of the effective mass theory near the donor ... Kohn and J. M. Luttinger, Phys. Rev. 97, 883 (1955). Omitted endnote.Missing: levels | Show results with:levels
  74. [74]
    Theory of band tails in heavily doped semiconductors
    Jul 1, 1992 · Morgan, T. N., 1965, "Broadening of impurity bands in heavily doped semiconductors," Phys. Rev. 139, 343A-348A; Morse, P. M., and H. Feshbach ...
  75. [75]
    Anomalous Optical Absorption Limit in InSb | Phys. Rev.
    Phys. Rev. 93, 632 ( ... Anomalous Optical Absorption Limit in InSb. Elias Burstein. Crystal Branch, Metallurgy Division, Naval Research Laboratory ...
  76. [76]
    [1602.06402] Band structure diagram paths based on crystallography
    Feb 20, 2016 · The electronic band structure is typically sampled along a path within the first Brillouin zone including the surface in reciprocal space.
  77. [77]
    An improved symmetry-based approach to reciprocal space path ...
    Jul 29, 2020 · Stationary points or cusps in a band structure can often be found at and along high-symmetry points and line segments in the Brillouin zone due ...
  78. [78]
    Effective mass and conductivity
    The effective mass varies with k. For conduction, the relevant effective mass values are determined by the band curvature at specific k points: the Fermi level ...
  79. [79]
    Electronic Structure | Materials Project Documentation
    Aug 18, 2025 · The line-mode NSCF calculation is run with k-points chosen along high-symmetry lines within the Brillouin zone of the material. Currently ...Missing: construction | Show results with:construction
  80. [80]
    3.3 Electronic structure calculations - Quantum Espresso
    0.2 Band structure calculation. First perform a SCF calculation as above; then do a non-SCF calculation (at fixed potential, computed in the previous step) ...
  81. [81]
    Visualizing the localized electrons of a kagome flat band
    Dec 20, 2023 · Electrons occupying flat bands in momentum space are expected to be localized in real space. Hence, materials with electronic flat bands present ...Missing: interpretation steep
  82. [82]
    An experimental determination of the Fermi surface in copper
    As determined, the surface, which holds one electron per atom, will not quite fit into the Brillouin zone, overlapping in the (111) directions where the zone ...
  83. [83]
    Band structure and carrier concentration of Gallium Arsenide (GaAs)
    GaAs has an energy gap of 1.424 eV, intrinsic carrier concentration of 2.1·10^6 cm^-3, and energy separations of 0.29 eV (Γ-L) and 0.48 eV (Γ-X).
  84. [84]
    [PDF] Simplified Band-Structure Models and Carrier Dynamics | nanoHUB
    The effective mass of a semiconductor is obtained by fitting the actual E-k diagram around the conduction band minimum or the valence band maximum by a parabola ...
  85. [85]
    NSM Archive - Band structure and carrier concentration of Silicon (Si)
    Effective density of states in the valence band​​ Nv = 3.5·1015·T3/2 (cm-3). The temperature dependence of the intrinsic carrier concentration.
  86. [86]
    NSM Archive - Diamond (C) - Band structure and carrier concentration
    Effective conduction band density of states, ~1020 cm-3. Effective valence band density of states, ~1019 cm-3. Band structure and carrier concentration of ...Missing: large indirect
  87. [87]
    From graphene to graphite: Electronic structure around the $K$ point
    Aug 2, 2006 · It is well known that a single graphene layer is a zero-gap semiconductor with a linear Dirac-like spectrum around the Fermi energy, while ...