Fact-checked by Grok 2 weeks ago

Free electron model

The free electron model is a foundational quantum mechanical approximation in that describes the conduction electrons in metals as a non-interacting gas of fermions confined within a periodic potential, obeying the and Fermi-Dirac statistics. This model simplifies the complex interactions in metals by treating electrons as plane waves in a three-dimensional box with , enabling calculations of electronic properties at temperature where states are filled up to the . Developed by Arnold Sommerfeld in the 1920s, the model built upon Paul Drude's classical 1900 theory by incorporating quantum mechanics to resolve discrepancies, such as the incorrect prediction of classical heat capacity. Key assumptions include neglecting electron-electron Coulomb interactions and electron-ion scattering in the basic formulation, assuming a uniform positive background charge from ion cores to maintain neutrality, and applying an infinite square well potential or periodic boundaries to model confinement. These simplifications yield the energy dispersion relation E = \frac{\hbar^2 k^2}{2m}, where k is the wavevector, leading to a spherical Fermi surface in k-space that separates occupied and unoccupied states. Central to the model is the E_F = \frac{\hbar^2}{2m} (3\pi^2 n)^{2/3}, which represents the maximum of electrons at T=0 K and depends on the n; for example, it is approximately 3.23 eV for sodium and 7.00 eV for . The g(E) = \frac{1}{2\pi^2} \left( \frac{2m}{\hbar^2} \right)^{3/2} E^{1/2} quantifies available quantum states per unit energy, scaling as \sqrt{E} in three dimensions and facilitating derivations of thermodynamic properties. The model successfully explains key metallic behaviors, including linear low-temperature electronic heat capacity C_{el} = \frac{\pi^2}{3} g(E_F) k_B^2 T, electrical and thermal conductivity via electron drift, and Pauli paramagnetism with susceptibility \chi = \frac{3n \mu_B^2}{2 E_F}. It is particularly accurate for simple metals like alkali elements (e.g., sodium, potassium) with nearly free electrons, though limitations arise in transition metals due to d-band effects, prompting extensions like the or tight-binding approximations.

Historical Development

Classical Drude Model

The classical , proposed by Paul Drude in 1900, represents the earliest systematic attempt to explain electrical conduction in metals using a kinetic theory approach, shortly after J.J. Thomson's discovery of the . In this framework, metals are viewed as a of fixed positive ions permeated by a gas of electrons that behave like classical particles, accelerating under an applied but frequently interrupted by collisions with the ions. Drude drew inspiration from the , adapting it to account for the one-dimensional motion of electrons along the field direction while assuming isotropic scattering. The model's core assumptions treat electrons as non-interacting classical particles free to move throughout the metal volume, with no forces between them and only occasional, random collisions with ions modeled as instantaneous and isotropic. These collisions are characterized by a relaxation time \tau, the average time between collisions, which determines how quickly electrons lose gained from . Under these premises, the steady-state v_d of electrons is given by v_d = -\frac{e \mathcal{E} \tau}{m}, where e is the electron charge, \mathcal{E} the electric field, and m the electron mass, leading to the current density \vec{J} = -n e \vec{v}_d. The resulting electrical is \sigma = \frac{n e^2 \tau}{m}, where n is the , providing a direct link between microscopic parameters and macroscopic transport. This formulation successfully explains through the linear relation \vec{J} = \sigma \vec{\mathcal{E}}, as the drift velocity remains proportional to the field for small \mathcal{E}. It also qualitatively accounts for the by predicting a transverse voltage due to the on drifting electrons in a , yielding the Hall R_H = -\frac{1}{n e}. Additionally, the model captures the positive temperature of resistivity, as increasing temperature enhances vibrations, shortening \tau and thus raising , consistent with observations in many metals. Despite these achievements, the fails to predict the electronic correctly, estimating it as C_V = \frac{3}{2} n k_B per unit volume—arising from the applied to the three-dimensional of the gas—which yields a value orders of magnitude larger than experimental measurements at . This discrepancy, among others, underscored the limitations of classical statistics and motivated the incorporation of in subsequent theories.

Transition to the Quantum Free Electron Model

The classical Drude model, while successful in describing electrical conductivity through a simple picture of drifting electrons, failed to account for several experimental observations in metals, such as the small magnitude of the electronic specific heat and the absence of classical equipartition. In 1927, Arnold Sommerfeld addressed these limitations by applying Fermi-Dirac quantum statistics to the free electron gas, thereby founding the quantum free electron model. Fermi-Dirac statistics, which had been independently derived by Enrico Fermi and Paul Dirac in 1926, This seminal contribution retained the Drude model's assumptions of non-interacting electrons moving freely in a constant potential but introduced quantum degeneracy effects, fundamentally altering the statistical treatment of electron occupation. Sommerfeld's approach, detailed in his paper "Zur Elektronentheorie der Metalle," provided a more accurate framework for metallic properties by incorporating wave mechanics and exclusion principles. A key conceptual shift in this transition was the recognition that electrons, as fermions, obey the , which prohibits two identical fermions from occupying the same simultaneously. Formulated by in 1925 to explain atomic spectral anomalies, this principle implies that electron states in momentum space are filled sequentially from the lowest energy up to a maximum , the E_F, at . At T = [0](/page/0), all states below E_F are fully occupied, and those above are empty, creating a sharp that separates occupied and unoccupied states. This degeneracy pressure arises purely from , contrasting with the classical model's reliance on thermal motion alone, and it resolves paradoxes like the stability of matter against collapse. The derivation of the quantum model builds directly on the framework by substituting the classical Maxwell-Boltzmann distribution for state occupation with the appropriate quantum distribution. In the classical case, the probability of an occupying a state of energy \epsilon is f(\epsilon) = e^{-(\epsilon - \mu)/kT}, assuming rare occupations. Sommerfeld replaced this with the Fermi-Dirac distribution, which accounts for the exclusion principle: f(\epsilon) = \frac{1}{\exp\left( \frac{\epsilon - \mu}{kT} \right) + 1} Here, \mu is the chemical potential, which at low temperatures (T \ll T_F = E_F / k) approximates E_F, ensuring the total number of electrons is conserved. This change modifies averages over the electron gas, such as energy and velocity distributions, while preserving the Drude-like relaxation time approximation for transport. A striking validation of this quantum upgrade is its prediction of the electronic heat capacity, which emerges as finite and linear in temperature (C_e = \gamma T), where \gamma is the Sommerfeld coefficient proportional to the density of states at E_F. In contrast, the classical Drude model erroneously predicts a temperature-independent heat capacity of (3/2) N k_B, vastly overestimating the electronic contribution compared to lattice vibrations observed in experiments on metals like copper. This linear behavior, derived from the smearing of occupations near the Fermi surface, aligns closely with low-temperature measurements and underscores the necessity of quantum statistics.

Fundamental Concepts and Assumptions

Core Ideas

The free electron model conceptualizes the electrons in metals as a gas of delocalized particles that move freely within the material, subject to a constant potential of zero (V=0) throughout the interior volume, while completely neglecting the periodic potential arising from the ion lattice and any direct electron-electron interactions. This simplification treats the metal as a uniform box filled with non-interacting electrons, analogous to an ideal quantum gas confined in a , enabling the application of basic to describe collective electronic behavior. Key assumptions underpin this framework: the system is modeled in a large cubic volume V = L^3 with to mimic an infinite, translationally invariant space, ensuring wavefunctions remain continuous across boundaries; the electrons are identical fermions that obey Fermi-Dirac statistics and the , filling available states up to a maximum energy at ; and scattering events between electrons are absent, with interactions limited to responses from external electric or magnetic fields. These conditions allow the electrons to be described as independent particles in a self-consistent neutralizing background of positive charge from the cores. The single-particle eigenstates in this model are plane waves of the form \psi_{\mathbf{k}}(\mathbf{r}) = \frac{1}{\sqrt{V}} \exp(i \mathbf{k} \cdot \mathbf{r}), where \mathbf{k} is the wavevector discretized by the boundary conditions, providing a simple basis for expanding the many-body wavefunction and computing properties like and distribution. This approach finds particular success in simple metals, such as the alkali metals (e.g., sodium, ), where conduction is primarily dominated by loosely bound s-electrons that exhibit behavior close to free particles due to weak binding and minimal overlap with core states. The omission of long-range interactions between electrons is further justified by screening effects within the high-density electron gas, where the uniform positive background and collective rearrangements of surrounding electrons effectively dampen the bare potential over distances beyond the Thomas-Fermi screening length, rendering the residual interactions perturbative rather than dominant.

Mathematical Formulation

The free electron model treats conduction electrons in a metal as a non-interacting confined within a potential-free region, governed by the H = \frac{p^2}{2m}, where p is the and m is the . In position space, this corresponds to the -\frac{\hbar^2}{2m} \nabla^2 \psi = \varepsilon \psi, assuming a constant potential V = 0 inside the sample. To model a large , periodic boundary conditions are imposed on a cubic of side L and volume V = L^3, ensuring the wavefunction satisfies \psi(\mathbf{r} + L \hat{x}) = \psi(\mathbf{r}) and similarly for y and z directions. Under these conditions, the eigenfunctions are plane waves \psi_{\mathbf{k}}(\mathbf{r}) = \frac{1}{\sqrt{V}} \exp(i \mathbf{k} \cdot \mathbf{r}), where \mathbf{k} is the wavevector. The allowed wavevectors are quantized as discrete points in : k_x = \frac{2\pi n_x}{L}, k_y = \frac{2\pi n_y}{L}, k_z = \frac{2\pi n_z}{L}, with integers n_x, n_y, n_z, leading to a uniform grid spacing \Delta k = \frac{2\pi}{L} along each axis. For large L, the sum over states becomes a , with the of allowed k-states per unit volume in given by \frac{1}{(2\pi)^3}. Accounting for the degeneracy of electrons (g = 2, for spin-up and spin-down), the number of states per unit volume in is \frac{2}{(2\pi)^3}. The single-particle energy spectrum is parabolic, starting from zero at k = 0: \varepsilon(\mathbf{k}) = \frac{\hbar^2 k^2}{2m}, where k = |\mathbf{k}|. This reflects the free-particle nature, with no band gaps or lattice effects. The total number of electrons N is determined by filling these states according to the Fermi-Dirac distribution f(\varepsilon_{\mathbf{k}}) = \frac{1}{\exp[(\varepsilon_{\mathbf{k}} - \mu)/k_B T] + 1}, where \mu is the , k_B is Boltzmann's constant, and T is : N = 2 \sum_{\mathbf{k}} f(\varepsilon_{\mathbf{k}}). For large systems, the sum is approximated as an integral over : N = \frac{V}{(2\pi)^3} \int 2 f(\varepsilon(\mathbf{k})) \, d^3\mathbf{k}.

Equilibrium Properties

Density of States

In the free electron model, the density of states D(\varepsilon) quantifies the number of available electron states per unit energy interval \mathrm{d}\varepsilon for a system of volume V, providing a crucial tool for understanding the distribution of electrons across energy levels in a metal.[Ashcroft and Mermin (1976) derive this quantity within the quantum mechanical framework of non-interacting electrons confined to a three-dimensional box with periodic boundary conditions, leading to plane-wave states labeled by wavevector \mathbf{k}.] The derivation begins by counting the number of states in \mathbf{k}-space. Each state occupies a volume (2\pi)^3 / V in \mathbf{k}-space, accounting for the large-system limit where states are densely packed. Including the two possible spin orientations for electrons, the number of states in a differential volume element \mathrm{d}^3\mathbf{k} is $2 \times (V / (2\pi)^3) \mathrm{d}^3\mathbf{k}. To relate this to energy, consider states between energies \varepsilon and \varepsilon + \mathrm{d}\varepsilon, which form a spherical shell in \mathbf{k}-space with volume $4\pi k^2 \mathrm{d}k, where the magnitude k = |\mathbf{k}| satisfies the parabolic dispersion \varepsilon = \frac{\hbar^2 k^2}{2m}. Solving for k gives k = \sqrt{2m\varepsilon}/\hbar, and differentiating yields \mathrm{d}k = \frac{\sqrt{2m}}{2\hbar} \varepsilon^{-1/2} \mathrm{d}\varepsilon. Substituting these into the state count produces the density of states: D(\varepsilon) = \frac{V}{2\pi^2} \left( \frac{2m}{\hbar^2} \right)^{3/2} \sqrt{\varepsilon}. This expression includes the factor of 2 for spin degeneracy and is valid for \varepsilon > 0. A key feature of this result is that D(\varepsilon) \propto \sqrt{\varepsilon}, reflecting the three-dimensional nature of the free electron gas: the increasing volume of the energy shell with energy leads to more states at higher energies, in contrast to lower-dimensional systems where the dependence differs (e.g., constant in 2D or inversely proportional to \sqrt{\varepsilon} in 1D). The enables the computation of the total number of electrons N in the system via integration over all energies, weighted by the occupation probability given by the Fermi-Dirac distribution f(\varepsilon): N = \int_0^\infty D(\varepsilon) f(\varepsilon) \, \mathrm{d}\varepsilon, where f(\varepsilon) = [ \exp((\varepsilon - \mu)/k_B T) + 1 ]^{-1}, with \mu the , k_B Boltzmann's constant, and T the . At (T = 0), f(\varepsilon) is a , filling all states up to the E_F and leaving higher states empty; this allows direct evaluation of the ground-state total energy as U_0 = \int_0^{E_F} \varepsilon \, D(\varepsilon) \, \mathrm{d}\varepsilon.

Fermi Energy and Surface

In the free electron model, the E_F is defined as the highest occupied at temperature (T = [0](/page/0)), below which all quantum states are filled according to the . This corresponds to the \mu at T = [0](/page/0), where \mu(T=[0](/page/0)) = E_F, determining the of the gas. The value of E_F is derived by ensuring the total number of electrons N fills the available states up to this energy, using the density of states D(\varepsilon) for a three-dimensional free electron gas. Specifically, N = \int_0^{E_F} D(\varepsilon) \, d\varepsilon, where the integral accounts for the volume in k-space occupied by electrons. Solving this yields the expression E_F = \frac{\hbar^2}{2m} (3\pi^2 n)^{2/3}, with n = N/V the electron number density, m the electron mass, and \hbar the reduced Planck's constant. The corresponding Fermi wave number is k_F = (3\pi^2 n)^{1/3}, defining the boundary in reciprocal space. The in the free electron model is a in centered at the origin with radius k_F, enclosing all filled states at T = 0. This geometric representation highlights the sharp cutoff between occupied and unoccupied states, a key feature of the degenerate . For typical metals, E_F ranges from 2 to 10 eV, while k_F \approx 10^8 cm^{-1}. At finite but low temperatures, the chemical potential \mu(T) exhibits a weak dependence on temperature, approximated by the Sommerfeld expansion as \mu(T) \approx E_F \left[1 - \frac{\pi^2}{12} \left(\frac{k_B T}{E_F}\right)^2 \right], where k_B is Boltzmann's constant; this correction arises from the smearing of the occupancy near E_F due to thermal excitations.

Degeneracy Pressure and Compressibility

In the free electron model, degeneracy pressure emerges as a fundamental quantum mechanical consequence of the Pauli exclusion principle, which requires electrons to occupy distinct quantum states, thereby filling the Fermi sea up to the Fermi energy even at absolute zero temperature. This filling implies that electrons possess a minimum kinetic energy distribution, leading to a pressure that originates from the Heisenberg uncertainty principle: confining electrons within the atomic lattice of a metal reduces their position uncertainty, necessitating a corresponding increase in momentum uncertainty and thus higher average kinetic energies that manifest as pressure. A distinctive feature of this degeneracy pressure is its independence from at low temperatures, in stark contrast to the classical where scales with . At T=0, the P relates directly to the total energy U and volume V of the system through the expression P = \frac{2}{3} \frac{U}{V}, where the U/V = (3/5) n E_F for a fully degenerate gas, with n denoting the and E_F the . Substituting yields P = \frac{2}{5} n E_F. This temperature-independent quantum pressure provides essential stability in dense electron systems, such as preventing gravitational collapse in white dwarfs, where it balances immense self-gravitational forces in a manner analogous to its role in maintaining the structural integrity of metals. The implications for compressibility are captured by the B = -V (\partial P / \partial V)_T, which for the degenerate free electron gas approximates to B \approx \frac{2}{3} n E_F. This expression highlights the model's success in predicting the resistance to compression in simple metals; for example, in magnesium with n \approx 8.61 \times 10^{28} m^{-3} , the calculated B is on the order of $10^{11} Pa, aligning reasonably with experimental measurements around $4.5 \times 10^{10} Pa and underscoring the dominance of electron degeneracy in metallic incompressibility.

Magnetic Susceptibility

In the free electron model, the magnetic susceptibility of the electron gas is dominated by Pauli paramagnetism, which arises from the alignment of spins near the in response to an applied B. This mechanism exploits the spin degree of freedom of electrons, where the external field preferentially populates spin-up states over spin-down states among the degenerate fermions at low temperatures. The derivation begins with the Zeeman splitting of the energy bands for spin-up and spin-down electrons, introducing an energy shift \Delta \epsilon = \pm \mu_B B, where \mu_B is the . This shift displaces the Fermi levels for the two spin populations, creating a net excess of electrons with magnetic moments aligned parallel to the field and inducing a M. The resulting Pauli paramagnetic is expressed as \chi_P = \mu_0 \mu_B^2 \frac{D(E_F)}{V}, where \frac{D(E_F)}{V} is the density of states per unit volume at the Fermi energy E_F, and \mu_0 is the vacuum permeability. This formula highlights the susceptibility's dependence on the availability of states near E_F for spin reorientation. In simple metals, Pauli paramagnetism yields a weak positive susceptibility of order $10^{-5} (in SI units, emu/mol), reflecting the small fraction of electrons responsive to typical fields. However, this is partially offset by the orbital Landau diamagnetism, which contributes a negative term approximately one-third the magnitude of \chi_P, resulting in a net susceptibility roughly (2/3) \chi_P. At sufficiently high magnetic fields, where \mu_B B becomes comparable to E_F (typically requiring B \sim 10^4 T for typical metallic Fermi energies), the electron gas achieves full spin polarization, with all spins aligned along the field direction.

Transport Properties

Electrical Conductivity and Mean Free Path

In the free electron model, the DC electrical conductivity is derived by extending the classical Drude formula using quantum statistics and the Boltzmann transport equation, treating electrons as a degenerate Fermi gas subject to scattering. The conductivity \sigma retains the form \sigma = \frac{n e^2 \tau}{m}, where n is the electron density, e is the electron charge, m is the electron mass, and \tau is the relaxation time obtained from the quantum relaxation time approximation, which accounts for Pauli exclusion and the Fermi-Dirac distribution. This quantum treatment resolves classical shortcomings by showing that only electrons within \sim kT of the Fermi energy E_F contribute significantly to conduction, unlike the classical case where all electrons participate equally. In the relaxation time approximation, the deviation from f - f_0 = -\tau(\mathbf{k}) \mathbf{v} \cdot \nabla_{\mathbf{k}} f_0 leads to \tau \approx constant near E_F, yielding \mathbf{j} \approx \frac{n e^2 \tau}{m} \mathbf{E} and thus \sigma = \frac{n e^2 \tau}{m}. The sharp peak of -\partial f / \partial \varepsilon at E_F ensures that the effective number of contributing electrons is n_{\text{eff}} \sim n (kT / E_F), explaining the temperature-independent at low T in pure metals. Scattering processes introduce the relaxation time \tau, related to the mean free path \lambda = v_F \tau, where v_F = \hbar k_F / m is the Fermi velocity and k_F = (3\pi^2 n)^{1/3} is the Fermi wavevector. In typical metals at room temperature, \lambda ranges from 10 to 100 nm, as calculated for elements like copper (\lambda \approx 40 nm) and silver (\lambda \approx 53 nm) using Fermi surface integrations. At low temperatures, conductivity is limited by impurity scattering, resulting in residual resistivity \rho_0 = m / (n e^2 \tau_{\text{imp}}), where \tau_{\text{imp}} is impurity-dominated./09%3A_Electronic_Properties_of_Materials_-_Superconductors_and_Semiconductors/9.05%3A_Resistivity) In high-purity metals, reduced impurity scattering allows \lambda to extend up to several microns, as observed in materials like PdCoO_2 with mean free paths reaching 20 \mum below 20 K.

Specific Heat Capacity

In the classical Drude model of metals, the electronic contribution to the specific heat capacity at constant volume is predicted to be C_V = \frac{3}{2} N k_B, where N is the number of conduction electrons and k_B is Boltzmann's constant; this value is independent of temperature but exceeds experimental observations by two to three orders of magnitude at low temperatures. The quantum free electron model, developed by Sommerfeld, resolves this discrepancy by incorporating Fermi-Dirac statistics, which restricts thermal excitations to electrons near the Fermi energy E_F due to the Pauli exclusion principle. At low temperatures, this yields a linear temperature dependence for the electronic specific heat: C_V = \gamma T, where the coefficient \gamma = \frac{\pi^2}{3} k_B^2 D(E_F) and D(E_F) is the density of states at the Fermi level. This result follows from the internal energy U = \int_0^\infty \epsilon \, D(\epsilon) \, f(\epsilon) \, d\epsilon, where f(\epsilon) is the Fermi-Dirac distribution function. For low temperatures T \ll T_F (with Fermi temperature T_F = E_F / k_B), the Sommerfeld expansion approximates the excited energy as U \approx U_0 + \frac{\pi^2}{6} D(E_F) (k_B T)^2, leading to C_V = \left( \frac{\partial U}{\partial T} \right)_V = \frac{\pi^2}{3} k_B^2 T D(E_F). Experimental measurements confirm the model's predictions for simple metals; for copper, the observed \gamma \approx 0.69 mJ/mol K² aligns closely with the theoretical value derived from the density of states. While lattice vibrations (phonons) dominate the total specific heat at , the linear electronic term is discernible in low-temperature data below approximately 10 , as verified by heat capacity plots for metals like .

Thermal Conductivity

In the free electron model, thermal conductivity arises from the transport of by conduction electrons in response to a ∇T. The mechanism involves an asymmetry in the Fermi distributions between hotter and cooler regions: electrons from the hotter side have a slightly higher due to the local increase in , leading to a net flow of toward the colder side when electrons diffuse across the gradient. The derivation of the thermal conductivity κ follows an approach analogous to that for electrical conductivity, using the Boltzmann transport equation in the relaxation time approximation. The heat current density j_Q is given by \mathbf{j}_Q = -\frac{\pi^2}{9} k_B^2 T D(E_F) v_F^2 \tau \nabla T, where D(E_F) is the at the , v_F is the Fermi velocity, k_B is Boltzmann's constant, T is , and τ is the relaxation time. By definition, j_Q = -κ ∇T, yielding \kappa = \frac{\pi^2}{9} k_B^2 T D(E_F) v_F^2 \tau. This expression links thermal transport directly to the electronic parameters at the Fermi level, emphasizing the role of electrons near E_F in carrying heat. A key result from this framework is the Wiedemann-Franz law, which relates thermal and electrical conductivities through the Lorenz number L = κ / (σ T) = (\pi^2 / 3) (k_B / e)^2 ≈ 2.45 \times 10^{-8} , \mathrm{W \Omega K^{-2}}, where σ is the electrical conductivity and e is the electron charge. This law emerges because both charge and heat currents are carried by the same electrons with the same relaxation time τ, assuming energy-independent scattering. Equivalently, κ = (\pi^2 / 3) (k_B^2 T / e^2) σ. The theoretical value of L matches experimental measurements for many metals at room temperature, providing strong validation for the free electron description of transport processes. The temperature dependence of κ varies with the dominant scattering mechanism. At low temperatures, where impurity scattering prevails and τ is approximately constant, κ ∝ T. At higher temperatures, dominates with τ ∝ T^{-1}, leading to κ ∝ T^{-1}. These behaviors highlight how influences the and thus the efficiency of heat transport by the electron gas.

Thermoelectric Power

The thermoelectric power, also known as the S, quantifies the voltage difference \Delta V generated across a due to a \Delta T, defined as S = -\frac{\Delta V}{\Delta T}. In the free electron model, this effect arises primarily from the diffusion of charge carriers in response to the temperature gradient, where hotter regions have higher velocities and a slight imbalance in the Fermi distribution drives a net current until balanced by an . Within the Boltzmann transport equation framework using the relaxation time approximation, the Seebeck coefficient for a degenerate electron gas is approximated using the Mott formula: S \approx -\frac{\pi^2}{3} \frac{k_B}{e} \frac{k_B T}{E_F} \left[ \frac{d \ln \sigma(\varepsilon)}{d \ln \varepsilon} \right]_{\varepsilon = E_F}, where k_B is the Boltzmann constant, e is the elementary charge, T is the temperature, E_F is the Fermi energy, and \sigma(\varepsilon) is the energy-dependent conductivity. For the free electron model assuming a constant relaxation time \tau, the conductivity \sigma(\varepsilon) \propto \varepsilon^{3/2}, yielding \left[ \frac{d \ln \sigma(\varepsilon)}{d \ln \varepsilon} \right]_{\varepsilon = E_F} = \frac{3}{2}, so the expression simplifies to S = -\frac{\pi^2 k_B^2 T}{2 e E_F}. This results in a negative value for electrons, reflecting their charge, and the magnitude is small due to the high degeneracy of the electron gas in metals, typically on the order of a few \mu \mathrm{V/K} at room temperature, dominated by the diffusion mechanism. The negative sign of S in the free electron model indicates n-type carriers (electrons), a feature that distinguishes carrier types and is particularly diagnostic in semiconductors, though the effect remains weak in metals where E_F is large (\sim 1-10 \, \mathrm{eV}). The basic model neglects the phonon-drag contribution, which involves momentum transfer from phonons to electrons and can enhance S at higher temperatures but is secondary in the low-temperature, diffusion-limited regime.

Limitations and Extensions

Principal Inaccuracies

The free electron model treats conduction electrons as non-interacting particles moving in a uniform potential, completely ignoring the periodic lattice potential of the crystal lattice. This oversight prevents the model from accounting for the formation of band gaps at boundaries, a key feature arising from Bragg reflection of electron waves. As a result, the model cannot explain why some materials are insulators or semiconductors, where the band gap separates filled valence bands from empty conduction bands, prohibiting electrical conduction at low temperatures; it predicts all materials with partially filled bands would conduct like metals. Additionally, the model neglects electron-electron interactions, particularly the long-range repulsion between electrons, which plays a critical role in determining the cohesive energy of metals. By assuming independent electrons in a uniform positive background (as in the approximation), it overestimates the because it fails to properly incorporate effects that reduce the effective repulsion; corrections via the Hartree-Fock method, which includes terms, are necessary to better approximate these interactions and yield more realistic values. The model's assumption of a parabolic energy dispersion leads to a spherical , implying isotropic independent of direction in . In contrast, real metals exhibit anisotropic distorted by the potential, as observed in noble metals like , silver, and , where deviations from sphericity affect properties such as magnetotransport. Furthermore, the free electron , which varies as the of , overestimates the availability of states at high energies because it disregards band gaps that suppress states in real materials. The absence of a periodic also eliminates processes, which involve vectors and are essential for explaining finite electrical and thermal resistivities at low temperatures in pure metals. The free electron model is especially inadequate for transition metals, where d-electrons exhibit localized character due to strong on-site correlations and incomplete d-shell filling, rather than delocalizing as free carriers; this localization leads to phenomena like and poor agreement with observed electronic properties, which the model cannot capture.

Advanced Models and Applications

The extends the basic free electron model by incorporating a weak periodic potential from the ionic lattice, treating it as a perturbation that mixes plane-wave states and opens energy gaps at the boundaries. This approach, developed through , accounts for the onset of structure in metals where the potential is not negligible but still weak compared to . The model predicts avoided crossings in the , explaining phenomena like the distinction between metals and insulators near zone edges, and serves as a bridge to more complex theory. Further advancements incorporate many-body interactions absent in the original model. Electron-phonon interactions, mediated by lattice vibrations, lead to phenomena such as , as described in the , where free electrons form pairs via exchange, resulting in a condensate with zero resistivity below a critical . Similarly, electron-electron interactions are captured by the , a method that computes the as the product of the one-particle and the screened Coulomb interaction, improving energies and band gaps in solids beyond mean-field approximations. A modern computational extension is orbital-free density functional theory (OFDFT), which approximates the non-interacting functional directly from the , bypassing the need for orbital solutions and inheriting the free electron model's Thomas-Fermi expression as a starting point. This makes OFDFT efficient for large-scale simulations of metallic systems, where the exact is parameterized via generalized gradient approximations or , achieving accuracy comparable to Kohn-Sham DFT for bulk properties while scaling linearly with system size. The free electron model underpins applications in low-dimensional systems, providing the basis for describing two-dimensional electron gases (2DEGs) in quantum wells and , where confinement quantizes the out-of-plane motion, yielding a constant in standard 2DEGs and a of states in , enabling phenomena like the integer . In quantum wells, such as GaAs heterostructures, the model predicts Shubnikov-de Haas oscillations in magnetotransport, while in , it approximates the low-energy Dirac spectrum for ballistic transport.

References

  1. [1]
    [PDF] Introduction to Solid State Physics
    This book is the eighth edition of an elementary text on solid state/ condensed matter physics for seniors and beginning graduate students of the physical ...
  2. [2]
  3. [3]
    Zur Elektronentheorie der Metalle - Drude - 1900 - Annalen der Physik
    566-613 Annalen der Physik. Article. Free Access. Zur Elektronentheorie der ... Drude, Phys. d. Aethers p. 212, Formel (36). Google Scholar. p610_2) Dies ...
  4. [4]
    The Drude Model - SciELO
    Oct 7, 2024 · In 1900, Paul Drude made a groundbreaking advancement with a model to describe the electrical conductivity of metals. Over 120 years later, ...Missing: paper | Show results with:paper
  5. [5]
    [PDF] Topic 8-1: Free Electron Model Kittel Pages: 133-138 Summary
    Begin with the free electron model developed by Summerfield in 1933. • Assumptions of the free electron model o No interactions between electrons and nuclei ...
  6. [6]
    [PDF] Chapter 6 Free-Electrons in Solids: Fermi Gas
    For this model, we make the following assumptions: (i) The crystal comprises a fixed background of N identical positively charge nuclei and N electrons, which ...
  7. [7]
    Zur Elektronentheorie der Metalle auf Grund der Fermischen Statistik
    Download PDF ... Cite this article. Sommerfeld, A. Zur Elektronentheorie der Metalle auf Grund der Fermischen Statistik. Z. Physik 47, 1–32 (1928).
  8. [8]
    Ashcroft Solid State : Free Download, Borrow, and Streaming
    Mar 14, 2019 · N.W. Ashcroft y N. D. Mermin, Solid State Physics, Saunders Collage Publishing, 1976 ... PDF WITH TEXT download · download 1 file · SINGLE ...
  9. [9]
    [PDF] Lecture 13: Metals
    Here is such a plot for copper: Figure 1. Heat capacity of copper at low temperatures compared to the prediction from the Sommerfeld free electron gas model.
  10. [10]
    None
    ### Summary of Fermi Energy, Typical Values, and Fermi Surface from Free Electron Model PDF
  11. [11]
    6.3: Degenerate Electron Gases - Physics LibreTexts
    Mar 31, 2025 · Indeed, most metals are hard to compress as a direct consequence of the high degeneracy pressure of their conduction electrons. To be more ...
  12. [12]
    [PDF] 193lApJ. THE MAXIMUM MASS OF IDEAL WHITE DWARFS
    extreme degeneracy, we can regard 1.822 X1033 as the maximum mass of an ideal white dwarf. This can be compared with the earlier esti- mate of Stoner2. Afmax ...
  13. [13]
    [PDF] Unit 3-10: Pauli Paramagnetism of the Electron Gas
    Pauli paramagnetism describes the magnetization of a degenerate ideal (non-interacting particles) Fermi gas of elec- trons with an external magnetic field. An ...
  14. [14]
    Tutorial: a beginner's guide to interpreting magnetic susceptibility ...
    Apr 19, 2022 · The susceptibility of non-localized conduction electrons, which by definition is only relevant to metals, is called Pauli paramagnetism. In ...
  15. [15]
    Pauli paramagnetism and Landau-Diamagnetism
    In general Pauli paramagnetism is much smaller than the paramagnetism due to atomic moments. E.g. for a fermi energy of 1 eV 1 e V , a field of B=17000 T B = ...
  16. [16]
    [PDF] 14.1 Boltzmann Transport equation - bingweb
    We discuss the Boltzmann transport equation which is very useful in understanding the transport properties such as electrical conductivity, ...
  17. [17]
    [PDF] Zur Elektronentheorie der Metalle - Gilles Montambaux
    SOMMERFELD: Zur Elektronentheorie der Metalle. 377 weil andernfalls alle Elektronen das Metall ver- lassen und dieses seinen metallischen Charakter.
  18. [18]
    Electron mean free path in elemental metals - AIP Publishing
    Feb 23, 2016 · The electron mean free path λ and carrier relaxation time τ of the twenty most conductive elemental metals are determined by numerical integration over the ...INTRODUCTION · II. COMPUTATIONAL... · III. RESULTS AND DISCUSSION
  19. [19]
    Directional ballistic transport in the two-dimensional metal PdCoO2
    Due to the strikingly high purity of this oxide, it can support electron mean free paths of up to 20 µm at temperatures below 20 K, as evidenced by an in-plane ...
  20. [20]
    [PDF] Sommerfeld-Drude model Ground state of ideal electron gas
    These observations imply that electrons obey Fermi-Dirac (FD) statistics. Sommerfeld Drude model: - Retains almost all aspects of Drude model with the following.
  21. [21]
    [PDF] Material Properties at Low Temperature
    The intensive property is given by the specific heat c , which is the heat capacity or ... where γ is the constant coefficient of the electronic contribution ...
  22. [22]
    [PDF] The Wiedemann-Franz Law for Electrical and Thermal Conduction ...
    Abstract: The Wiedemann-Franz law states that the ratio of the thermal conductivity to the electrical conductivity is the same for all metals at a given.
  23. [23]
    Understanding the thermal conductivity and Lorenz number in ...
    Jan 22, 2019 · The proportionality constant L , known as the Lorenz number, is theoretically equal to the Sommerfeld value L 0 = π 2 k B 2 / 3 e 2 = 2.445 × 10 ...
  24. [24]
    [PDF] Thermal conductivity
    The mean free path Λ of the conduction electron due to impurity scattering is obviously temperature independent. •C el is proportional to the absolute ...
  25. [25]
    Mott's formula for the thermopower and the Wiedemann-Franz law
    May 15, 1980 · The Mott formula for the thermopower, S = ( π 2 3 ) ( k 2 B ⁢ T e ) σ ′ σ , and the Wiedemann-Franz law, K σ ⁢ T = ( k B e )
  26. [26]
    None
    ### Summary of Seebeck Coefficient Derivation in Quasi-Free Electron Model
  27. [27]
    Thermocouple Principles—the Seebeck Effect ... - All About Circuits
    Jul 10, 2022 · The magnitude and sign of the Seebeck coefficient are related to an asymmetry of the electron distribution around the Fermi level. To learn ...
  28. [28]
    [PDF] Energy bands - Rutgers Physics
    Energy bands are discussed in the context of the free electron model, which fails to explain the difference between metals, semiconductors and insulators. ...
  29. [29]
    [PDF] PART IV. Fermi Statistics of Electrons & Electron-Electron Interactions
    Fermi Statistics of Electrons & Electron-Electron Interactions. References: 1 ... • Estimate of the metallic cohesive energy in a nearly-free-electron model.
  30. [30]
    Life on the edge : A beginner's guide to the Fermi surface
    Apr 18, 2016 · Of course, the energy bands of real metals deviate from the free electron parabolæ, and the Fermi surfaces are not spheres. In many cases, the ...
  31. [31]
    [PDF] 58 Chapter 3 Failures of the Free Electron Model
    58 Chapter 3 Failures of the Free Electron Model. Free electron theory successfully accounts for a wide range of metallic properties. In the form originally ...
  32. [32]
    [PDF] 1 The course is taught on basis of lecture notes which are ...
    However, the Drude model appears to fail for noble metals, such as copper (reddish) and gold (yellow), although it works reasonably well for silver. The ...
  33. [33]
    Nearly-Free Electron Approach | Basic Theory of Surface States
    Oct 31, 2023 · Abstract. Although Maue (1935) was the first to use the nearly-free electron approximation (NFE) to study surface states, and their effect ...
  34. [34]
    [PDF] arXiv:1905.12084v2 [cond-mat.mtrl-sci] 10 Dec 2019
    Dec 10, 2019 · This corresponds to the so-called nearly-free-electron model when P is “small”. One sees that band 1 has negative energy, as found for the ...
  35. [35]
    Orbital-free density functional theory: Kinetic potentials and ab initio ...
    May 23, 2007 · Orbital-free methods attempt to calculate this directly from the electron density by approximating the universal but unknown kinetic energy ...
  36. [36]
    Recent advancements and challenges in orbital‐free density ...
    Jun 12, 2024 · In particular, the kinetic energy density functionals (KEDFs) are utilized to approximately calculate the electronic kinetic energy in OFDFT, ...Abstract · INTRODUCTION · BASIC THEORY · ALGORITHMS AND...
  37. [37]
    Engineering artificial graphene in a two-dimensional electron gas
    Graphene is a one-atom-thick two-dimensional (2D) electron system composed of carbon atoms on a honeycomb lattice. · Because the typical values of a 0 are much ...
  38. [38]
  39. [39]
    Realization of a Pre-Sample Photonic-Based Free-Electron ...
    Oct 9, 2025 · These results demonstrate the ability to manipulate the electron wave function before probing the sample, paving the way for photonics-inspired ...