Fact-checked by Grok 2 weeks ago

Local-density approximation

The local-density approximation (LDA) is a seminal approximation scheme within , a computational framework for quantum many-body systems, where it estimates the exchange-correlation energy by treating the as locally uniform at each point in space, drawing from properties of a homogeneous gas. This approach enables practical calculations of ground-state properties of atoms, molecules, and solids by mapping the interacting system onto a non-interacting one governed by effective single-particle equations. LDA emerged from foundational work in the , building on the Hohenberg-Kohn theorems of , which established that the ground-state energy of a system is uniquely determined by its . In their landmark 1965 paper, and Lu Jeu Sham introduced the Kohn-Sham equations, which incorporate exchange and correlation effects through an ; LDA provides the simplest realization of this by parameterizing the exchange-correlation functional based on the uniform electron gas model. Early parametrizations for the correlation energy, such as those by Wigner in the 1930s and later refinements like the Perdew-Wang form in 1992, have been integral to its implementation. Mathematically, the LDA exchange-correlation energy is expressed as E_{xc}^{\text{LDA}} = \int n(\mathbf{r}) \varepsilon_{xc}^{\text{unif}}(n(\mathbf{r})) \, d\mathbf{r}, where n(\mathbf{r}) is the and \varepsilon_{xc}^{\text{unif}} is the per-particle exchange-correlation energy of a uniform gas at density n(\mathbf{r}). The exchange part is derived exactly from the Dirac exchange for the uniform gas, \varepsilon_x^{\text{unif}}(n) = -\frac{3}{4} \left( \frac{3}{\pi} n \right)^{1/3}, while the correlation is approximated using numerical fits to data or analytic forms. This local treatment simplifies self-consistent field calculations, making LDA computationally efficient for large systems. LDA has proven highly effective for predicting structural properties, such as lattice constants accurate to within 4% and bond lengths to 0.1 in many materials, particularly those with slowly varying densities like wide-band semiconductors and sp-bonded systems. It excels in uniform or nearly uniform electron distributions, where it recovers exact results, and has been widely applied in for band structures and in for molecular geometries. However, LDA tends to overestimate energies by about 10% and underestimate band gaps, leading to metallic predictions for some insulators. To address LDA's shortcomings, such as its neglect of density gradients, subsequent developments include the (GGA), which incorporates spatial variations in the density for improved accuracy in transition metals and molecular energies. Despite these advances, LDA remains a cornerstone of DFT due to its simplicity, interpretability, and role as a benchmark for more sophisticated functionals.

Fundamentals

Definition and Principles

The local-density approximation (LDA) serves as a approximation within the framework of (DFT), specifically in the Kohn-Sham formulation, where the challenging task of accounting for electron-electron interactions is addressed through an effective single-particle potential. In this approach, the exchange-correlation energy functional E_{xc}[\rho], which captures the quantum mechanical effects beyond the classical repulsion, is approximated by assuming that these effects can be treated locally based on the \rho(\mathbf{r}) at each point in space. This simplification is essential for practical computations, as the exact E_{xc}[\rho] remains unknown and inherently nonlocal. The core mathematical expression of LDA is given by E_{xc}[\rho] \approx \int \varepsilon_{xc}(\rho(\mathbf{r})) \, \rho(\mathbf{r}) \, d^3\mathbf{r}, where \varepsilon_{xc}(\rho) represents the exchange-correlation energy per particle for a homogeneous gas (HEG) evaluated at the local \rho(\mathbf{r}). This formulation embodies the fundamental assumption that exchange-correlation interactions in an inhomogeneous system, such as atoms or molecules, can be modeled by dividing the system into regions, each behaving like a uniform HEG with \rho(\mathbf{r}). By doing so, LDA reduces the to solving a set of self-consistent Kohn-Sham equations for non-interacting electrons moving in an that includes the local exchange-correlation contribution. A distinguishing feature of LDA is its status as the simplest semilocal approximation in the hierarchy of DFT functionals, as it depends only on the itself and disregards spatial variations or gradients in \rho(\mathbf{r}). This local treatment makes LDA computationally efficient and widely applicable, enabling accurate predictions for a broad range of materials properties despite its approximations. For systems with magnetic ordering, LDA is naturally extended to the local spin-density approximation (LSDA), which incorporates polarization through separate densities for spin-up and spin-down electrons.

Historical Development

The local-density approximation (LDA) in (DFT) traces its origins to early semiclassical models of electron behavior in atoms and solids. In the , the Thomas-Fermi model provided the first density-based approach to , treating electrons as a non-interacting in a local potential to estimate and total energy. This model, independently developed by Llewellyn Thomas and , laid the groundwork for approximating properties of many-electron systems using rather than wavefunctions, motivated by the need to simplify calculations for heavy atoms beyond Hartree-Fock methods. Building on this, approximations for exchange effects in uniform electron systems emerged in the 1930s, with deriving an exact expression for the exchange energy of the homogeneous electron gas (HEG), which became central to later local approximations. Eugene Wigner extended this by introducing an estimate for the correlation energy in metals, accounting for electron-electron interactions beyond mean-field treatments through a perturbation expansion. These HEG-based insights influenced John C. Slater's Xα method in 1951, which approximated the nonlocal Hartree-Fock with a local potential parameterized by α, enabling practical computations for atoms, molecules, and solids while bridging statistical and wavefunction approaches. The formal foundation of modern DFT arrived with the Hohenberg-Kohn theorems in 1964, proving that the ground-state energy is a functional of the , followed by and Lu Jeu Sham's 1965 equations, which reformulated the into single-particle equations resembling Hartree-Fock but with an exchange-correlation potential. LDA emerged naturally as the inaugural approximation for this potential, interpolating HEG exchange-correlation energies locally based on density, as suggested in the Kohn-Sham to make computations tractable for inhomogeneous systems. In the 1970s, Ola Gunnarsson and Bertil I. Lundqvist advanced LDA's application through spin-density functional calculations for atoms, providing interpolation formulas for exchange-correlation energies and demonstrating its utility in predicting ionization potentials comparable to experiment. LDA gained widespread adoption in the 1980s for solid-state and molecular simulations, bolstered by accurate parametrizations of the HEG correlation functional from data. David M. Ceperley and Berni J. Alder computed precise ground-state energies for the unpolarized and polarized HEG using variational and diffusion methods, offering benchmark results that refined LDA's correlation component and improved its reliability for practical use. This evolution from semiclassical roots to a cornerstone of computational and physics highlighted LDA's role in enabling efficient, density-only treatments of complex systems.

Theoretical Basis

Homogeneous Electron Gas

The homogeneous electron gas (HEG), also known as the jellium model, represents an idealized infinite system of interacting electrons embedded in a uniform positive background charge density to ensure overall neutrality. This model simulates the behavior of electrons in metals where the density varies slowly, providing the physical basis for exchange-correlation effects in uniform-density systems. A central parameter of the HEG is the Wigner-Seitz radius r_s = \left( \frac{3}{4\pi \rho} \right)^{1/3}, where \rho is the uniform electron density, defining the effective average inter-electron spacing in . In the high-density regime (r_s \to 0), interactions are weak, and the system approximates a non-interacting amenable to perturbative treatments. Conversely, in the low-density regime (r_s \to \infty), strong correlations dominate, potentially leading to ordered phases such as the Wigner crystal. The exchange energy per particle in the HEG is exactly solvable within the Hartree-Fock approximation for uniform , yielding \varepsilon_x(\rho) = -\frac{3}{4} \left( \frac{3}{\pi} \right)^{1/3} \rho^{1/3}. This Dirac exchange expression accounts for the antisymmetry of the fermionic wave function in the uniform system. In contrast, the correlation energy per particle \varepsilon_c(\rho) lacks an exact analytic form due to the complexity of many-body correlations beyond mean-field treatments. Accurate numerical benchmarks were obtained by Ceperley and Alder using diffusion simulations of the ground-state , providing correlation energies for representative densities spanning r_s from near 0 to high values like 100. These data enabled practical parametrizations, such as the widely used Perdew-Zunger fit, which interpolates \varepsilon_c(\rho) analytically for application in density functional approximations across all densities. The total energy per particle of the HEG is given by \varepsilon(\rho) = \varepsilon_\mathrm{kin}(\rho) + \varepsilon_H(\rho) + \varepsilon_{xc}(\rho), where \varepsilon_{xc}(\rho) = \varepsilon_x(\rho) + \varepsilon_c(\rho). The term \varepsilon_H(\rho) vanishes identically due to the uniform positive background exactly canceling the classical electron-electron repulsion. The per particle \varepsilon_\mathrm{kin}(\rho) is typically approximated via the Thomas-Fermi model for the non-interacting uniform , \varepsilon_\mathrm{kin}(\rho) = \frac{3}{10} (3\pi^2 \rho)^{2/3}, which captures the leading semiclassical behavior of the Fermi sea.

Exchange Functional

The functional in the local-density approximation (LDA) approximates the energy of an inhomogeneous \rho(\mathbf{r}) by integrating the energy per particle of a homogeneous electron gas (HEG) at the local : E_x[\rho] = \int \varepsilon_x(\rho(\mathbf{r})) \rho(\mathbf{r}) \, d\mathbf{r}, where \varepsilon_x(\rho) is the energy per for an HEG of \rho. This form arises from the exact Hartree-Fock treatment of the HEG, where the ground-state wave function is a Slater determinant constructed from plane waves filling up to the Fermi momentum k_F = (3\pi^2 \rho)^{1/3}. The exchange energy is computed by evaluating the nonlocal exchange operator's expectation value over this determinant, which simplifies due to the uniform density and translational invariance, yielding a local contribution proportional to \rho^{4/3} for the total energy density. The resulting per-particle exchange energy is \varepsilon_x(\rho) = -\frac{3}{4} \left( \frac{3}{\pi} \right)^{1/3} \rho^{1/3}, known as the Dirac exchange expression, reflecting the $1/3 power dependence on density from the Fermi surface scaling. In pure LDA implementations, this exact HEG form is used without modification for the energy. However, early variants like the X\alpha method introduce a scaling parameter \alpha to the potential, approximating v_x(\mathbf{r}) = -\alpha (3/\pi \rho(\mathbf{r}))^{1/3}, with \alpha = 2/3 chosen to ensure variational consistency between the energy functional and its for the HEG, bridging the full Dirac energy (\alpha = 1) and a statistically averaged potential. A key limitation of the LDA exchange functional is its partial cancellation of the self-interaction in the term, which represents the classical electron-electron repulsion including unphysical self-repulsion. In exact , the exchange-correlation functional should fully cancel this self- energy for one-electron densities, but the LDA only approximately does so due to its delocalized HEG form, leading to a self-interaction error that delocalizes electrons and underestimates in localized systems. This component forms part of the full exchange-correlation functional when combined with a term.

Correlation Functional

The functional in the local-density approximation (LDA) approximates the energy of an interacting system by integrating the energy per of a homogeneous gas (HEG) at the local : E_c[\rho] = \int \epsilon_c(\rho(\mathbf{r})) \rho(\mathbf{r}) \, d\mathbf{r}. Here, \epsilon_c(\rho) captures dynamic correlations beyond the static hole, including the short-range - cusp and long-range many-body effects that reduce the classical repulsion. A seminal parametrization is the Perdew-Zunger (PZ81) form, developed by fitting to accurate (QMC) calculations of the HEG correlation energy by Ceperley and Alder, which provided data for unpolarized (\zeta = 0) and fully polarized (\zeta = 1) cases at metallic and low densities (r_s = 1 to $100, where r_s = (3/(4\pi\rho))^{1/3} is the Wigner-Seitz radius in ). The PZ81 functional interpolates between high-density (small r_s, weak correlation) and low-density (large r_s, strong correlation) regimes using distinct analytic expressions to ensure smooth matching at r_s = 1 and adherence to known asymptotic behaviors. For intermediate spin polarizations $0 < \zeta < 1, it employs a linear interpolation in \zeta. In the high-density regime (r_s < 1), where perturbation theory dominates, PZ81 uses the logarithmic expansion derived from random-phase approximation (RPA) and beyond: \epsilon_c(r_s) = A \ln r_s + B + C r_s \ln r_s + D r_s with constants A = 0.0310907, B = -0.048, C = 0.0020, and D = -0.0116 for the unpolarized case (similar forms apply for polarized, with adjusted coefficients). This recovers the exact high-density limit \epsilon_c \sim (1 - \ln 2)/\pi^2 \ln r_s from Gell-Mann and Brueckner. In the low-density regime (r_s > 1), PZ81 adopts the Gunnarsson-Lundqvist , motivated by RPA for the correlation hole: \epsilon_c(r_s) = \frac{\gamma}{1 + \beta_1 \sqrt{r_s} + \beta_2 r_s} with \gamma = -0.1423, \beta_1 = 1.0529, and \beta_2 = 0.3334 for unpolarized HEG (again, polarized variants differ slightly). These forms achieve fitting errors below 0.3% relative to QMC data across the sampled densities. No exact closed-form expression exists for \epsilon_c(\rho), necessitating empirical fits like PZ81, which, while benchmarked against 1980s QMC standards, exhibit limitations in capturing full HEG correlations, particularly near r_s \approx 1 where regimes overlap. In practice, PZ81 and similar LDA correlation functionals often overestimate molecular binding energies by 20-50% compared to experiment, as the local HEG assumption amplifies correlation in delocalized, low-density regions, leading to excessive attraction despite partial cancellation with exchange overestimation. This stems from the neglect of density gradients and self-interaction errors inherent in LDA.

Spin-Dependent Formulations

Spin Polarization

In the context of the local-density approximation (LDA) for , spin polarization accounts for the imbalance between spin-up and spin-down densities in magnetic systems, extending the uniform gas model to include magnetic effects. The spin polarization parameter is defined as \zeta = \frac{\rho_\uparrow - \rho_\downarrow}{\rho_\uparrow + \rho_\downarrow}, where \rho_\uparrow and \rho_\downarrow are the spin-up and spin-down densities, respectively, and the total is \rho = \rho_\uparrow + \rho_\downarrow. This parameter ranges from \zeta = 0 for paramagnetic (non-magnetic or equally populated spin channels) systems to \zeta = 1 for fully ferromagnetic states where all s have the same spin orientation, with no minority spin component. For the spin-polarized homogeneous gas (HEG), the exchange-correlation per particle is formulated as \varepsilon_{xc}(\rho, \zeta), which depends on both the total and the degree of . This increases (becomes less negative) with increasing |\zeta| due to reduced screening effects in the polarized case, where the imbalance limits the ability of opposite spins to screen interactions effectively. At \zeta = 0, the formulation reduces to the standard spin-unpolarized LDA, while at \zeta = 1, it describes a fully spin-polarized gas with no opposite-spin contributions. This \zeta-dependence introduces a physical analogy to the , where the stability of ferromagnetic order is assessed by the curvature of the with respect to spin at \zeta = 0. The incorporation of spin in LDA is motivated by its ability to describe in materials such as transition metals and magnetic insulators, where unpaired spins lead to net magnetic moments. For instance, it captures the itinerant in elements like iron, , and by allowing self-consistent determination of spin densities that stabilize magnetic ground states. However, LDA often overestimates the tendency toward , predicting metallic ferromagnetic phases in cases where more accurate methods indicate or weaker , such as in certain high-pressure phases of transition metals. This extension forms the basis for the local spin-density approximation (LSDA), which applies these concepts to inhomogeneous systems.

Local Spin-Density Approximation

The local spin-density approximation (LSDA) generalizes the local-density approximation to account for polarization in systems, treating spin-up (\rho_\uparrow) and spin-down (\rho_\downarrow) densities as variables. This extension enables the of magnetic properties by incorporating the effects of spin imbalance into the exchange-correlation functional. Introduced for spin-polarized cases, LSDA relies on parametrizations derived from the uniform gas with varying spin polarization. The LSDA exchange-correlation energy functional takes the form E_{xc}[\rho_\uparrow, \rho_\downarrow] = \int \epsilon_{xc}(\rho(\mathbf{r}), \zeta(\mathbf{r})) \rho(\mathbf{r}) \, d\mathbf{r}, where \rho(\mathbf{r}) = \rho_\uparrow(\mathbf{r}) + \rho_\downarrow(\mathbf{r}) is the total and \zeta(\mathbf{r}) = (\rho_\uparrow(\mathbf{r}) - \rho_\downarrow(\mathbf{r}))/\rho(\mathbf{r}) is the local parameter, ranging from -1 (fully down-polarized) to +1 (fully up-polarized). This formulation differs from the spin-unpolarized LDA by explicitly depending on \zeta, which modulates the exchange-correlation per particle \epsilon_{xc} to reflect spin-dependent interactions in the slowly varying . For the exchange contribution, LSDA employs spin-scaling relations exact for the homogeneous electron gas, expressing the exchange energy per particle as \epsilon_x(\rho, \zeta) = \epsilon_x^P(\rho) \frac{(1 + \zeta)^{4/3} + (1 - \zeta)^{4/3}}{2}, where \epsilon_x^P(\rho) = -\frac{3}{4} \left( \frac{3}{\pi} \right)^{1/3} \rho^{1/3} is the unpolarized exchange per particle from Dirac's . This scaling arises from treating the spin-up and spin-down components as separate unpolarized gases with densities \rho_\uparrow = \frac{\rho}{2}(1 + \zeta) and \rho_\downarrow = \frac{\rho}{2}(1 - \zeta), ensuring the correct high-polarization limit where \epsilon_x \propto \rho^{1/3} independently of . The correlation part in LSDA is approximated through parametrized interpolations between unpolarized (\zeta = 0) and fully polarized (|\zeta| = 1) limits of the uniform gas, often showing increased correlation energy at higher |\zeta| due to reduced screening in polarized systems. Von Barth and Hedin proposed an early form using the same scaling as exchange for correlation, \epsilon_c(\rho, \zeta) = \epsilon_c^P(\rho) \frac{(1 + \zeta)^{4/3} + (1 - \zeta)^{4/3}}{2}, derived from perturbation theory and dielectric response. A more accurate and widely adopted parametrization is the spin extension by Perdew and Zunger (PZ81), which fits quantum Monte Carlo simulations of Ceperley and Alder to provide \epsilon_c(r_s, \zeta) as a function of the density parameter r_s = (3/(4\pi \rho))^{1/3} and \zeta, with explicit polynomial forms ensuring smooth interpolation and enhanced correlation for |\zeta| > 0. LSDA's primary advantages lie in its ability to capture spin-split bands arising from interactions in magnetic materials, enabling reliable predictions of magnetic moments and band structures in collinear spin configurations. It is routinely applied in Kohn-Sham calculations for collinear magnetism, where the spin polarization \zeta enters the self-consistent equations to yield separate effective potentials for up and down .

Exchange-Correlation Potential

Derivation

In the Kohn-Sham framework of , the exchange-correlation potential is defined as the functional derivative of the exchange-correlation energy functional with respect to the :
v_{\mathrm{xc}}(\mathbf{r}) = \frac{\delta E_{\mathrm{xc}}[\rho]}{\delta \rho(\mathbf{r})}.
This potential enters the Kohn-Sham equations as an effective one-electron term, acting as a local, multiplicative operator that shifts the eigenvalues of the non-interacting reference system to approximate the true many-electron spectrum.
Within the local-density approximation (LDA), the exchange-correlation is expressed by integrating the contribution from a homogeneous gas at the local :
E_{\mathrm{xc}}^{\mathrm{LDA}}[\rho] = \int d\mathbf{r} \, \rho(\mathbf{r}) \, \varepsilon_{\mathrm{xc}}\bigl(\rho(\mathbf{r})\bigr),
where \varepsilon_{\mathrm{xc}}(\rho) denotes the exchange-correlation per in a uniform gas of \rho. The functional derivative then yields a local potential:
v_{\mathrm{xc}}^{\mathrm{LDA}}(\mathbf{r}) = \varepsilon_{\mathrm{xc}}\bigl(\rho(\mathbf{r})\bigr) + \rho(\mathbf{r}) \frac{d \varepsilon_{\mathrm{xc}}}{d \rho} \bigg|_{\rho=\rho(\mathbf{r})}.
For the exchange-only contribution in the unpolarized case, \varepsilon_x(\rho) = -\frac{3}{4} \left( \frac{3}{\pi} \right)^{1/3} \rho^{1/3}, which leads to v_x(\rho) = \frac{4}{3} \varepsilon_x(\rho).
For spin-polarized systems, the local spin-density approximation (LSDA) extends this form using spin densities \rho_\uparrow(\mathbf{r}) and \rho_\downarrow(\mathbf{r}):
E_{\mathrm{xc}}^{\mathrm{LSDA}}[\rho_\uparrow, \rho_\downarrow] = \int d\mathbf{r} \, e_{\mathrm{xc}}\bigl(\rho_\uparrow(\mathbf{r}), \rho_\downarrow(\mathbf{r})\bigr),
where e_{\mathrm{xc}}(\rho_\uparrow, \rho_\downarrow) is the exchange-correlation energy per volume for a uniform spin-polarized gas. The spin-resolved potentials are obtained via partial derivatives:
v_{\mathrm{xc},\sigma}(\mathbf{r}) = \frac{\partial e_{\mathrm{xc}}}{\partial \rho_\sigma} \bigg|_{\rho_\uparrow=\rho_\uparrow(\mathbf{r}), \, \rho_\downarrow=\rho_\downarrow(\mathbf{r})},
with \sigma = \uparrow, \downarrow. These derivatives incorporate dependence on the local spin \zeta(\mathbf{r}) = \frac{\rho_\uparrow(\mathbf{r}) - \rho_\downarrow(\mathbf{r})}{\rho(\mathbf{r})}, reflecting the varying exchange-correlation behavior with spin imbalance.

Numerical Implementation

In numerical implementations of the local-density approximation (LDA) within (DFT), the \rho(\mathbf{r}) is discretized either on a real-space grid for direct evaluation or expanded in a basis set such as plane waves for periodic systems or Gaussian functions for molecular calculations. In plane-wave approaches, prevalent for solids, the density and potentials are represented in reciprocal space and efficiently transformed to real space via fast transforms (FFTs), a key computational advance from the that enabled the practical use of plane-wave methods. The LDA exchange-correlation potential v_{xc}(\mathbf{r}) is then computed on the grid and interpolated back to the basis set for solving the Kohn-Sham equations. The core of the implementation is the self-consistent field (SCF) cycle, which iteratively solves the Kohn-Sham equations: an initial density guess is used to construct the (including , external, and v_{xc}), the resulting orbitals are obtained via , and a new is formed from these orbitals; this process repeats until the or total converges to a specified , often $10^{-5} to $10^{-7} /bohr^3. To promote and , particularly in oscillatory cases like metals, density mixing schemes blend the input and output densities—such as simple linear mixing with a , Kerker preconditioning for long-wavelength components, or more advanced Broyden- or Anderson-mixing methods that extrapolate based on prior iterations. Widely used codes for LDA calculations include and , both employing approximations to treat core electrons explicitly while focusing on valence electrons, thereby reducing basis set size and computational demands; however, pseudopotentials introduce approximations that can affect transferability and accuracy for core-related properties like energies. These codes support LDA functionals parameterized for the homogeneous electron gas, with using projector-augmented wave () methods and offering norm-conserving or ultrasoft variants. Regarding efficiency, LDA evaluation is particularly straightforward, scaling linearly O(M) with the number of grid points M since v_{xc} depends only on local \rho(\mathbf{r}), contrasting with the cubic O(N^3) scaling of Hamiltonian diagonalization where N is the basis size or number of electrons, which dominates costs for large systems; optimizations like FFTs further mitigate overhead in plane-wave codes.

Applications

Atomic and Molecular Systems

The local-density approximation (LDA) has been applied to atomic systems since the early days of density functional theory, providing reasonably accurate total energies, particularly for core electrons in high-density regions. In atomic calculations, LDA performs well for inner-shell electrons, yielding total energies close to Hartree-Fock values. However, LDA tends to underestimate bond lengths in molecular contexts by 1-2%, as seen in small diatomic species where predicted equilibrium distances are 0.01-0.03 Å shorter than experimental values, arising from overestimation of binding strengths. Early DFT calculations using LDA-like approximations, such as the Xα method, were performed on small molecules in the , enabling self-consistent treatments of electronic structure for systems like complexes and diatomics. These pioneering efforts, led by groups including Baerends and collaborators, demonstrated LDA's feasibility for molecular geometries and vibrational frequencies, often matching experimental data within 5% for bond lengths in simple hydrides. For molecular examples, LDA accurately describes the H₂ dissociation curve near , reproducing the and energy minimum effectively, but fails at large internuclear separations due to self-interaction error, predicting spurious residual binding instead of the correct to two neutral atoms. In benchmarks against exact methods like Hartree-Fock or configuration interaction, LDA ionization potentials for atoms and small molecules exhibit errors of 0.3-1.0 , typically underestimating values due to incorrect asymptotic behavior of the potential, though core-level are more accurate within 0.1 . LDA also overbinds rare-gas dimers excessively, computing binding energies for He₂ and Ne₂ that are an larger than van der Waals realities (e.g., ~1-2 meV experimental vs. ~10-20 meV LDA), highlighting its limitations for weakly interacting systems. These applications underscore LDA's utility for qualitative insights into atomic and molecular electronic structure despite quantitative shortcomings.

Solid-State Physics

In , the local-density approximation (LDA) is widely applied to extended periodic systems such as crystals, metals, and surfaces, where it facilitates the computation of electronic band structures and related properties within (DFT). By assuming the exchange-correlation energy depends locally on the , LDA enables efficient self-consistent field calculations under , making it a cornerstone for predicting material behaviors in infinite lattices. This approach has been particularly successful in describing delocalized electrons in metals and semiconductors since the 1990s, as computational advancements allowed for routine applications to complex solids. As of 2025, LDA remains widely used in high-throughput DFT calculations for materials discovery due to its computational efficiency. A notable limitation of LDA in semiconductors is its systematic underestimation of band gaps, typically by about 50%, arising from the absence of the derivative discontinuity in the exchange-correlation potential. For instance, LDA predicts a band gap of 0.5 for , compared to the experimental value of 1.1 , which hampers accurate modeling of optical and properties. This discrepancy stems from LDA's piecewise linearity in the total energy with respect to electron number, missing the jump at integer occupancy that corrects the gap in exact DFT. Despite issues, LDA excels in predicting structural properties of solids, yielding constants with typical mean absolute relative errors of about 1-2% across diverse materials, alongside reliable dispersions and cohesive energies, especially in metals. In alkali metals like sodium and , LDA provides excellent agreement with experimental cohesive energies and parameters, capturing the nearly free-electron behavior effectively. For transition metal oxides, the spin-dependent local spin-density approximation (LSDA) variant successfully describes magnetic ordering and moments, as seen in compounds like and FeO, where it reproduces antiferromagnetic ground states without additional corrections in many cases. To handle the computational demands of solid-state calculations, LDA is often paired with pseudopotential methods to treat approximately, or all-electron approaches for higher accuracy. The projector augmented-wave () method stands out for its efficiency in solids, combining pseudopotential speed with all-electron precision by augmenting plane-wave basis sets with localized projectors around atomic cores, enabling scalable simulations of large periodic systems like and surfaces.

Limitations and Extensions

Shortcomings of LDA

The local-density approximation (LDA) suffers from a significant self-interaction error, where an experiences a spurious repulsion with itself due to the approximate treatment of and effects. This error is particularly evident in one-electron systems, such as the , where the exact total energy is -0.5 , but LDA yields approximately -0.768 using standard parametrizations like Ceperley-Alder. Similarly, for the H_2^+ molecular , LDA fails to cancel this unphysical self-repulsion, leading to incorrect energies and bond lengths. LDA assumes a locally uniform , neglecting variations in the density gradient, which causes failures in systems with rapidly varying densities. This limitation is pronounced in weakly bound systems like van der Waals interactions, where LDA underbinds rare gas dimers and layered materials by missing long-range dispersion forces, often predicting zero or near-zero binding energies. At surfaces, such as the model, LDA overestimates the surface energy by up to 100% or more for low-density regimes (r_s < 4), predicting unphysically negative values that imply instability. A key theoretical flaw in LDA is its inability to capture the derivative discontinuity in the exchange-correlation energy per particle, ε_xc(ρ), at numbers. This results in severe underestimation of gaps in insulators and semiconductors, typically by 40-50%, and can even predict metallic behavior for true insulators like (LDA gap ~0.5 eV vs. experimental 1.1 eV). In molecular and solid-state systems, LDA generally overbinds atoms and molecules, with cohesive energies overestimated by 20-50% compared to experiment, as seen in diatomics like N_2 and bulk metals. This overbinding arises from the inadequate description of the exchange-correlation hole in nonuniform densities, leading to excessively attractive interactions.

Improvements and Alternatives

The generalized gradient approximation (GGA) extends the local-density approximation (LDA) by incorporating the gradient of the electron density, ∇ρ, into the exchange-correlation functional, allowing for better treatment of density inhomogeneities. This improvement addresses key LDA shortcomings, such as overbinding in molecules and underestimation of band gaps in solids. The Perdew-Burke-Ernzerhof (PBE) functional, introduced in 1996, exemplifies GGA by providing more accurate bond lengths and atomization energies compared to LDA, with reduced errors in band gaps for semiconductors (typically 20-50% improvement over LDA predictions). However, GGAs like PBE introduce parameters derived from fundamental constants, which can lead to slight overestimation of lattice constants in some metals. Hybrid functionals further enhance accuracy by mixing a portion of exact Hartree-Fock with DFT , mitigating self-interaction errors more effectively than pure GGAs. The B3LYP functional, proposed by Becke in 1993, combines 20% exact with Becke's GGA and Lee-Yang-Parr correlation, yielding superior performance for and energies in molecular systems. For solids, the Heyd-Scuseria-Ernzerhof (HSE) screened hybrid, developed in 2004, attenuates long-range to reduce computational cost while improving predictions, often achieving errors of about 0.3 eV for semiconductors—roughly half the typical GGA error. These functionals excel in describing charge transfer and but demand higher computational resources due to the non-local exact term. Meta-generalized gradient approximations (meta-GGAs) build on GGAs by including the density or Laplacian of the , enabling satisfaction of more exact constraints without empirical fitting. The SCAN functional, introduced by Perdew and colleagues in 2016, incorporates these terms to improve descriptions of both weak interactions and strong correlations, outperforming PBE in cohesive energies and surface tensions while maintaining similar computational scaling. Beyond meta-GGAs, the (RPA) addresses correlation effects nonlocally by integrating over the response function, providing accurate van der Waals interactions and total energies when used atop semilocal functionals; it is particularly effective for layered materials and dispersion-bound systems. Despite these advances, LDA remains sufficient and widely used as a in computational codes for high-density, nearly uniform systems like simple metals (e.g., alkali metals), where gradients are small and error cancellations yield reliable structural and energetic properties. In 2025, LDA continues to serve as the starting point for many large-scale simulations in due to its simplicity and low cost.
ApproximationAccuracy (e.g., band gaps, bonds)Computational Cost
LDAModerate; underestimates gaps by ~50%, overbinds moleculesLow (O(N^3))
GGA (e.g., PBE)Good; reduces gap errors to ~30%, better bondsLow (similar to LDA)
Hybrid (e.g., )High; gap errors ~10-20%, accurate charge transferHigh (O(N^4))
This table summarizes the hierarchy conceptually, with hybrids offering the best balance for precision-critical applications despite increased expense.

References

  1. [1]
    Self-Consistent Equations Including Exchange and Correlation Effects
    KOHN AND L. J. SHAM. Unieersity of Ca/Bfornia, San Diego, la Jolta, California. (Received 21 June 1965l. From a theory of Hohenberg and Kohn, approximation.
  2. [2]
    [PDF] Density Functional Theory - Purdue Math
    A more accurate approximation is the Local Density Approximation (LDA), described below. This is much more accurate but overcompensates, leading to binding ...
  3. [3]
    [PDF] Density Functional Theory (DFT) - Rutgers Physics
    The Local Density Approximation is very successful for ”wide-band” systems with open s and p orbitals and sometimes also for d systems. It works also for band ...
  4. [4]
  5. [5]
  6. [6]
    Ground State of the Electron Gas by a Stochastic Method
    - **Study**: Ground State of the Electron Gas by a Stochastic Method
  7. [7]
    [PDF] II. My first density functional: Thomas-Fermi Theory
    Feb 2, 2011 · , the energy per particle is: (2.2.19). Plugging into Eq. (2.2.3) ... Let us study an unperturbed homogeneous electron gas using Thomas-Fermi.
  8. [8]
    Simple and Accurate Exchange Energy for Density Functional Theory
    E x [ ρ ] = ∫ ρ ε x F ( s ) d 3 r . Here, ε x = − 3 4 ( 3 π ρ ) 1 / 3 was the Dirac exchange energy for uniform electron density [18].Missing: particle | Show results with:particle
  9. [9]
    Note on Exchange Phenomena in the Thomas Atom
    Oct 24, 2008 · However, as you have access to this content, a full PDF is available via the 'Save PDF' action button. For dealing with atoms involving many ...
  10. [10]
    A Simplification of the Hartree-Fock Method | Phys. Rev.
    This simplified field is being applied to problems in atomic structure, with satisfactory results, and is adapted as well to problems of molecules and solids.Missing: Xα | Show results with:Xα
  11. [11]
    5.3.2 Exchange Functionals - Q-Chem Manual
    ∘. Local Spin-Density Approximation (LSDA). ∙. Slater: Slater-Dirac exchange functional (Xα method with α=2/3) · ∘. Generalized Gradient Approximation (GGA). ∙.<|control11|><|separator|>
  12. [12]
    Self-interaction correction to density-functional approximations for ...
    May 15, 1981 · Self-interaction correction to density-functional approximations for many-electron systems. J. P. Perdew · Alex Zunger.
  13. [13]
    Self-interaction corrections in density functional theory - AIP Publishing
    Mar 4, 2014 · The self-interaction corrections get rid of the self-interaction error, which is the sum of the Coulomb and exchange self-interactions that ...
  14. [14]
    How accurate are the parametrized correlation energies of the ...
    May 15, 2018 · Unlike the simpler PZ81, the other parametrizations and the DPI employ a single analytic form for the whole range 0 ≤ r s ≤ ∞ , although this ...
  15. [15]
    Large-Scale Benchmark of Exchange–Correlation Functionals for ...
    Jul 15, 2019 · In the current work we make use of the Perdew–Zunger (35) (PZ81) approximation, which captures the spin dependency by interpolating between the ...
  16. [16]
    A local exchange-correlation potential for the spin polarized case. i
    Citation U von Barth and L Hedin 1972 J. Phys. C: Solid State Phys. 5 1629DOI 10.1088/0022-3719/5/13/012. Download Article PDF. Article metrics. 9284 Total ...
  17. [17]
    [PDF] Density functional theory - University of Washington
    A Local density approximation . ... 4/3. + n. 4/3. ) (26). The correlation energy of a spin-polarized electron gas ...
  18. [18]
  19. [19]
    [PDF] Local-density functional calculations of the energy of atoms
    Jan 22, 1997 · The Kohn-Sham local-density approximation (LDA) and its variants are widely used in ab initio computations of materials properties [14].
  20. [20]
    Density functional theory: Its origins, rise to prominence, and future
    Aug 25, 2015 · This paper reviews the development of density-related methods back to the early years of quantum mechanics and follows the breakthrough in their application ...
  21. [21]
    [PDF] arXiv:1303.6185v2 [physics.chem-ph] 21 May 2013
    May 21, 2013 · For He2 and Ne2, LDA leads to binding en- ergies which are one order of magnitude too large and to distances which are about 0.5 Å too small.<|control11|><|separator|>
  22. [22]
    Quantifying the Accuracy of Density Functionals on Transition Metal ...
    Nov 9, 2023 · For instance, LDA functionals, which overestimate them, have stronger cohesive energies, which are directly connected to the surface energies ...
  23. [23]
    [PDF] Is the density functional theory band gap problem truly a ... - OSTI
    Si: expt: 1.2 eV, DFT/LDA: 0.5 eV. GaAs: expt. 1.5 eV, DFT/LDA: 0.5 eV. The band gap defines the energy scale for defect levels. Fundamental impediment to ...
  24. [24]
    Density functional theory and the band gap problem - Perdew - 1985
    Mar 23, 1985 · (2) The derivative discontinuity of the exchange-correlation energy, which is responsible for substantial underestimation of the fundamental gap ...
  25. [25]
    Accurate lattice geometrical parameters and bulk moduli from a ...
    Sep 11, 2018 · We find that the TM functional is able to produce very accurate lattice constants, with a mean absolute error of 0.038 Å ⁠, and bulk moduli ...
  26. [26]
    Gaussian-basis LDA and GGA calculations for alkali-metal ...
    May 15, 1998 · Our results indicate that the Gaussian-LCAO method should be able to give accurate results for nearly any crystalline solid, since it succeeds ...
  27. [27]
    Insulating gap in the transition-metal oxides: A calculation using the ...
    Apr 15, 1994 · In contrast to LSDA, the LSDA+U calculations reveal the experimentally observed antiferromagnetic and insulating ground state. The values of the ...<|control11|><|separator|>
  28. [28]
    Challenges for Density Functional Theory | Chemical Reviews
    Although LDA gives good geometries, it massively overbinds molecules. The seminal work of Becke, Perdew, Langreth, and Parr in the 1980s, which introduced the ...
  29. [29]
    Not Found
    **Summary:**
  30. [30]
    Toward a benchmark for the jellium surface energy | Phys. Rev. B
    Jan 15, 2000 · LDA makes compensating errors at intermediate and small wave vectors. Studies of small jellium clusters also support the density-functional ...
  31. [31]
    Generalized Gradient Approximation - ScienceDirect.com
    The generalized gradient approximation (GGA) is defined as a method that incorporates the gradient of the electron density to account for the inhomogeneity of ...
  32. [32]
    Generalized Gradient Approximation Made Simple | Phys. Rev. Lett.
    Oct 28, 1996 · We present a simple derivation of a simple GGA, in which all parameters (other than those in LSD) are fundamental constants.
  33. [33]
    Exchange-correlation functionals for band gaps of solids - Nature
    Jul 10, 2020 · This is due to the large overestimation of band gaps smaller than 1 eV, while beyond this range it underestimates band gaps, but performs ...<|control11|><|separator|>
  34. [34]
    Density‐functional thermochemistry. III. The role of exact exchange
    A semiempirical exchange‐correlation functional containing local‐spin‐density, gradient, and exact‐exchange terms is tested on 56 atomization energies.Missing: paper URL
  35. [35]
    Efficient hybrid density functional calculations in solids: Assessment ...
    Efficient hybrid density functional calculations in solids: Assessment of the Heyd–Scuseria–Ernzerhof screened Coulomb hybrid functional Available. Jochen Heyd;.Missing: paper | Show results with:paper
  36. [36]
    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 ...
  37. [37]
    Random-Phase Approximation Methods - Annual Reviews
    May 5, 2017 · Random-phase approximation (RPA) methods are rapidly emerging as cost-effective validation tools for semilocal density functional computations.Missing: paper | Show results with:paper
  38. [38]
    Local density approximation combined with Gutzwiller method for ...
    Feb 17, 2009 · This is the reason why LDA works well for simple metals, such as Na and K, where wide s band crosses the Fermi level, but it fails for ...
  39. [39]
    Effects of density functional semilocality and van der Waals nonlocality
    LDA works for metal surfaces through two known error cancellations. The Perdew–Burke–Ernzerhof generalized gradient approximation tends to underestimate both ...
  40. [40]
    Comparison of the Performance of Density Functional Methods for ...
    Apr 15, 2023 · This work analyzes the performance of 250 electronic structure theory methods (including 240 density functional approximations) for the description of spin ...
  41. [41]
    Thirty years of density functional theory in computational chemistry
    ABSTRACT. In the past 30 years, Kohn–Sham density functional theory has emerged as the most popular electronic structure method in computational chemistry.