Fact-checked by Grok 2 weeks ago

GW approximation

The GW approximation is a cornerstone method in many-body for calculating the electronic in interacting many-electron systems, approximating the self-energy \Sigma as the product of the single-particle G and the dynamically screened interaction W. Introduced by Lars Hedin in through a set of self-consistent equations for the , it provides a systematic framework to go beyond mean-field approximations like Hartree-Fock by incorporating screening effects and electron correlations via perturbation expansion. This approach yields energies and lifetimes, essential for understanding electronic excitations in solids, molecules, and nanostructures. The method excels in predicting accurate structures and spectra, particularly addressing the underestimation of band gaps in (DFT) calculations using local or semilocal approximations. For instance, in semiconductors like and , GW computations yield fundamental gaps close to experimental values (e.g., 1.17 eV for versus 1.24–1.29 eV from GW), outperforming DFT's typical errors of 0.5–1 eV. Its applications span insulators, transition metals (e.g., and ), surfaces (e.g., Si(100)), clusters, fullerenes like C_{60}, and more recently, two-dimensional materials and molecular systems for ionization potentials and electron affinities. In , GW variants like G_0W_0@HF demonstrate high accuracy for weakly and strongly correlated molecules, with mean absolute errors for ionization potentials around 0.3 eV on the GW100 benchmark set. Despite its successes, the GW approximation has limitations, including computational cost scaling as O(N^4) (where N is the system size) in naive implementations, often mitigated by approximations like the plasmon-pole model or partial self-consistency. It neglects corrections in the (setting the vertex function \Gamma \approx 1), which can lead to overestimation of self-energy magnitudes by 10–20% in some cases, and struggles with strongly correlated systems exhibiting structures or Mott transitions. Extensions like self-consistent GW (SCGW), GW+DMFT, or inclusion of vertex effects aim to address these, but the perturbative G_0W_0 variant—using a starting DFT or Hartree-Fock —remains the most practical and widely used. Ongoing developments focus on efficient algorithms for large systems and integration with for screening.

Fundamentals

Definition

The GW approximation is a method within many-body used to compute energies and wavefunctions for the electronic structure of solids and molecules, going beyond mean-field approaches such as Hartree-Fock. It approximates the operator \Sigma, which accounts for electron-electron interactions, using the one-particle G and the screened interaction W, yielding \Sigma \approx iGW. This framework captures dynamic screening effects, where the Coulomb interaction is dressed by the response of the electron gas, providing a more accurate description of excitation energies compared to static mean-field treatments. The basic motivation for the GW approximation stems from the shortcomings of mean-field methods, which neglect many-body correlations and dynamic screening, often leading to underestimated band gaps in semiconductors and insulators. By incorporating these effects perturbatively, GW improves predictions for quasiparticle properties, such as ionization potentials and electron affinities, making it particularly valuable for materials where exchange-correlation interactions are significant. In the GW scheme, the self-energy is approximated as \Sigma \approx i G W, where the product denotes the convolution over space and time (or ) coordinates, and the vertex function \Gamma—which accounts for higher-order vertex corrections—is set to unity for simplicity. The method is named after its key ingredients, G (the ) and W (the screened interaction), and was originally introduced by Lars Hedin in 1965 for the electron gas, with seminal applications to semiconductors following in the .

Relation to Density Functional Theory

The GW approximation is frequently integrated with (DFT) as a post-processing step to enhance the accuracy of electronic structure predictions, particularly for excited-state properties where standard DFT falls short. In the commonly used G0W0 variant, the non-interacting Green's function G0 is constructed from the Kohn-Sham orbitals and eigenvalues obtained from a DFT ground-state calculation, typically employing local or semi-local functionals like PBE or LDA; this hybrid approach leverages DFT's computational efficiency while incorporating many-body to compute the . A primary motivation for adopting GW atop DFT is to rectify the notorious underestimation of band gaps in semiconductors and insulators by approximate DFT exchange-correlation functionals, which arise from their neglect of derivative discontinuities and inadequate treatment of long-range exchange. For instance, in bulk silicon, the PBE functional predicts an indirect band gap of approximately 0.6 eV, while G0W0@PBE yields about 1.2 eV, aligning closely with the experimental value of 1.17 eV. This correction stems from the GW self-energy's inclusion of screened exchange and correlation effects, which shift quasiparticle levels more accurately than DFT's static potential. The standard computational workflow begins with a DFT calculation to obtain the Kohn-Sham wave functions and densities, which serve as the basis for evaluating the and screened interaction W in the random-phase approximation; the is then computed perturbatively using these inputs, followed by of the to update the energies. Relative to pure DFT, this GW@DFT strategy provides superior spectra by accounting for non-local, dynamical exchange-correlation beyond semi-local s, though at a higher computational cost scaling as O(N4) in implementations, where N is the .

Theoretical Framework

Green's Function

In many-body perturbation theory, the one-electron Green's function serves as the central quantity for describing the propagation of electrons in interacting systems. It is defined as the time-ordered propagator G(1,2;\omega), where $1 = (\mathbf{r}, \tau) denotes space-time coordinates with \mathbf{r} the position and \tau the imaginary time, and \omega the frequency. Physically, G represents the probability amplitude for adding an electron at position \mathbf{r}_2 and time \tau_2 and subsequently removing it at \mathbf{r}_1 and \tau_1, or vice versa, thereby capturing single-particle excitations that change the particle number by one. The spectral representation provides a Lehman expansion of the Green's function in terms of quasiparticle states, linking it directly to the excitation spectrum: G(\mathbf{r},\mathbf{r}';\omega) = \sum_n \frac{\psi_n(\mathbf{r}) \psi_n^*(\mathbf{r}')}{\omega - \varepsilon_n + i\eta \operatorname{sgn}(\varepsilon_n - \mu)}, where \{\psi_n, \varepsilon_n\} are the quasiparticle wavefunctions and energies, \mu is the , and \eta is a positive ensuring . The poles of G at \omega = \varepsilon_n correspond to the quasiparticle energies, while the residues give the quasiparticle strengths, quantifying the extent to which the excitations resemble non-interacting particles. This representation is particularly useful for interpreting the effects of interactions on the electronic structure. In practical calculations, the time-ordered Green's function is often evaluated using the Matsubara formalism for finite-temperature systems, where imaginary frequencies i\omega_n (with \omega_n = (2n+1)\pi/\beta for fermions, \beta = 1/k_B T) replace real frequencies to avoid singularities and enable to the retarded form via techniques like the . The Matsubara approach facilitates the summation over thermal Matsubara frequencies in perturbation expansions, making it suitable for equilibrium properties at nonzero temperatures. Within the GW approximation, computations typically begin with the non-interacting Green's function G_0, constructed from Kohn-Sham orbitals and eigenvalues obtained via (DFT), which provides an efficient starting point for the perturbative . Subsequent iterations may update G through Dyson's equation to incorporate interactions, though the single-shot G_0 W_0 variant often suffices for energies in solids. This strategy leverages DFT's accuracy for ground-state properties while correcting band gaps and dispersions via many-body effects.

Screened Coulomb Interaction W

The screened Coulomb interaction W, a central quantity in the GW approximation, describes the effective electron-electron in a many-body , incorporating dielectric screening effects from the surrounding electrons. It is defined in space, time, and frequency as W(\mathbf{r},\mathbf{r}';\omega) = v(\mathbf{r},\mathbf{r}') + \int d\mathbf{r}_1 d\mathbf{r}_2 \, v(\mathbf{r},\mathbf{r}_1) P(\mathbf{r}_1,\mathbf{r}_2;\omega) v(\mathbf{r}_2,\mathbf{r}'), where v(\mathbf{r},\mathbf{r}') = 1/|\mathbf{r} - \mathbf{r}'| is the bare potential and P is the irreducible function representing the density response of the . This form arises from the Dyson equation for the screened , truncating higher-order terms in the polarization . In practice, the polarizability P is often evaluated within the random phase approximation (RPA), where it is approximated as the independent-particle response P(\mathbf{r},\mathbf{r}';\omega) \approx -i \int \frac{d\omega'}{2\pi} G(\mathbf{r},\mathbf{r}';\omega + \omega') G(\mathbf{r}',\mathbf{r};\omega'), using the non-interacting Green's function G_0 and neglecting vertex corrections. This RPA level for P simplifies computations while capturing the leading screening effects through ring diagrams in the perturbation expansion. The W is dynamically screened and frequency-dependent, reflecting the response of the electronic medium to perturbations at different scales. At high frequencies (\omega \to \infty), screening diminishes, and W(\omega) approaches the bare potential v, as rapid oscillations prevent charge rearrangement. At low frequencies, W(\omega) incorporates collective excitations such as plasmons in metals or excitonic effects tied to the bandgap in insulators, leading to enhanced screening and a softened . In metallic systems, such as , the imaginary part of W exhibits a characteristic edge, manifesting as peaks around 20–30 corresponding to satellites that broaden and renormalize bands. In insulators like , screening is weaker and bandgap-dependent, with W contributing to gap corrections through effects near atomic bonds, increasing the fundamental gap by approximately 0.7 relative to local-density approximations.

Self-Energy Operator

In the GW approximation, the self-energy operator \Sigma encapsulates the effects of many-body electron-electron interactions and serves as an effective, non-local, and frequency-dependent potential in the description of quasiparticles. It is defined in the space-time representation as \Sigma(1,2) \approx i \, [G](/page/G)(1,2) [W](/page/W)(1^+,2), where $1 = (\mathbf{r}_1, t_1) and $1^+ = (\mathbf{r}_1, t_1 + 0^+) denotes an infinitesimal positive time shift to ensure ordering, [G](/page/G) is the one-particle , and [W](/page/W) is the screened . This form arises from a perturbative expansion where the vertex function is approximated by the identity, reducing higher-order diagrams. The self-energy \Sigma can be decomposed into an exchange part \Sigma_x and a correlation part \Sigma_c, such that \Sigma = \Sigma_x + \Sigma_c. The exchange component \Sigma_x is static and resembles the Fock operator but uses the bare Coulomb interaction v, obtained by setting W \to v in the GW expression, capturing the non-local exchange effects akin to Hartree-Fock theory. In contrast, the correlation part \Sigma_c is dynamic and stems from the screening effects in W - v, accounting for the frequency-dependent response of the electron gas to perturbations and including phenomena like plasmons. This enters the equation, which linearizes the equation for practical computation: [h_0 + V_{xc} + \operatorname{Re} \Sigma(\varepsilon)] \psi = \varepsilon \psi, where h_0 is the non-interacting (kinetic plus external potential), V_{xc} is the exchange-correlation potential from , and the real part of \Sigma at the \varepsilon corrects the single-particle eigenvalues. Due to the dependence of \Sigma, the wave functions and are renormalized by the factor Z = \left[1 - \left. \frac{\partial \operatorname{Re} \Sigma}{\partial \omega} \right|_{\omega = \varepsilon} \right]^{-1}, which represents the residue or spectral weight and typically lies between 0 and 1, reflecting the finite lifetime from imaginary parts of \Sigma. The renormalization Z < 1 indicates that carry only a of the total spectral weight, with the remainder distributed to satellites due to many-body excitations. For computational efficiency in large systems or when full GW calculations are prohibitive, the scissor operator approximation simplifies the self-energy effects by applying a uniform energy shift to the conduction bands relative to the valence bands, typically by the GW-corrected bandgap difference, without resolving band- or momentum-dependent details. This approach mimics the average quasiparticle correction from GW while avoiding the need to compute \Sigma for each state, though it neglects variations in \Sigma_x and \Sigma_c across the spectrum.

Formalism and Derivation

Hedin's Equations

In 1965, Lars Hedin derived a set of five coupled equations that provide a systematic framework for computing the one-particle Green's function in interacting many-electron systems using many-body perturbation theory. These equations, known as Hedin's equations, relate the Green's function G, the self-energy \Sigma, the irreducible polarizability P, the screened Coulomb interaction W, and the vertex function \Gamma in a self-consistent manner, enabling an exact description of electronic correlations starting from the bare Coulomb interaction. The formulation is foundational for ab initio many-body theory, as it transforms the intractable many-body problem into a hierarchy of response functions that can, in principle, be solved iteratively. The first equation is Dyson's equation, which expresses the interacting in terms of the non-interacting Green's function G_0 and the : G(1,2) = G_0(1,2) + \int d3\, d4\, G_0(1,3) \Sigma(3,4) G(4,2) Here, the arguments denote space-time coordinates, $1 \equiv (\mathbf{r}_1, z_1) where z is a frequency, and integrals are over all such coordinates. This equation captures propagation under interactions. The second equation defines the vertex function \Gamma, which accounts for screening beyond the random-phase approximation, via a of the with respect to the Green's function: \Gamma(1,2;3) = \delta(1,3)\delta(2,3) + \int d4\, d5\, d6\, d7\, \frac{\delta \Sigma(1,2)}{\delta G(4,5)} G(4,6) G(7,5) \Gamma(6,7;3) This irreducible vertex function introduces higher-order correlations in the response of the system. The third equation gives the irreducible polarizability P, representing the density-density response: P(1,2) = -i \int d3\, d4\, G(1,3) G(4,1) \Gamma(3,4;2) It describes how the electron density fluctuates in response to perturbations, incorporating vertex corrections. The fourth equation relates the screened Coulomb interaction W to the bare Coulomb potential v through the dielectric function \varepsilon = 1 - v P, yielding W = \varepsilon^{-1} v, or equivalently in integral form: W(1,2) = v(1,2) + \int d3\, d4\, v(1,3) P(3,4) W(4,2) This screening accounts for the reduction of the long-range repulsion due to electronic . Finally, the fifth equation expresses the self-energy \Sigma, which encodes all many-body effects beyond the mean field: \Sigma(1,2) = i \int d3\, d4\, G(1,4) W(1^+,3) \Gamma(4,2;3) The $1^+ indicates a time ordering where the time of the first argument is infinitesimally later, ensuring proper contour integration. Solving Hedin's equations exactly is computationally intractable due to the nonlinear coupling and the need for functional derivatives, which generate an infinite hierarchy of response functions. This complexity motivates approximations, such as setting \Gamma = 1 to truncate the hierarchy, which leads to practical schemes like the GW approximation.

GW Approximation Scheme

The GW approximation scheme applies specific truncations to Hedin's equations to make the computation of the tractable. In this approach, the function is set to unity (\Gamma = 1), neglecting higher-order corrections that account for local field effects and exchange-correlation beyond the (RPA). The irreducible is approximated using the RPA, given by P = -i G G, where the product of two Green's functions captures the leading for screening. These simplifications yield the in the form \Sigma = i G W, where W is the screened interaction derived from the inverse function \varepsilon^{-1} = 1 - v P, with v the bare potential. This truncation balances computational feasibility with accuracy for properties in many materials, though it omits dynamic screening effects in the vertex that can be important for strongly correlated systems. Within the GW framework, calculations are performed at different levels of self-consistency to trade off accuracy and cost. The non-self-consistent G₀W₀ approximation uses a starting G₀ typically obtained from (DFT) calculations, with the \Sigma computed perturbatively as a one-shot correction to the DFT eigenvalues: \varepsilon^{QP} = \varepsilon^{DFT} + \langle\psi^{DFT}| \Sigma(\varepsilon^{QP}) - V^{xc} |\psi^{DFT}\rangle, where V^{xc} is the exchange-correlation potential. This level often provides band gaps and ionization potentials within 0.1–0.3 eV of experiment for semiconductors like and . More advanced is the GW₀ scheme, which iterates the G self-consistently while keeping the screened interaction W fixed at the initial RPA level from G₀; this eigenvalue self-consistency improves convergence for occupied states but may overestimate band gaps in some cases. The fully self-consistent GW approach iterates both G and W until convergence of the , enforcing laws and yielding total energies closer to coupled-cluster benchmarks, though it is computationally intensive and sometimes leads to unphysical metallic states in insulators due to over-screening.

Computational Aspects

Numerical Methods

The choice of basis set is crucial for representing the and screened Coulomb interaction in GW calculations, influencing both accuracy and computational efficiency. For periodic systems, such as solids and surfaces, plane-wave basis sets are widely employed due to their ability to efficiently describe delocalized electronic states via Bloch expansions. These basis sets naturally incorporate and are particularly suited for uniform electron gases or crystalline materials, though they require pseudopotentials to handle and large cutoffs to achieve convergence. In contrast, for molecular systems or isolated defects, Gaussian-type orbitals centered on atoms provide a compact representation, leveraging their analytical integrals for overlap and elements. These localized functions facilitate all-electron treatments but demand careful optimization of exponents to minimize basis set superposition errors. Numerical atomic orbitals, which use discretized radial functions on a grid, offer flexibility for including core states explicitly and are effective in mixed-basis approaches for large-scale simulations. Evaluating the frequency-dependent quantities in the GW self-energy requires careful numerical integration over the imaginary or real frequency axis to avoid singularities and ensure convergence. The contour deformation technique deforms the integration path into the , bypassing poles on the real axis associated with excitations and plasmons, which allows for a sparser sampling while maintaining accuracy for both and core levels. This method is particularly advantageous for systems with strong correlations or deep core states, as it reduces the number of frequency points needed compared to direct real-axis . Alternatively, analytic continuation approximates the real-frequency self-energy by fitting data computed on the imaginary axis using techniques like Padé , providing an efficient route for band calculations but requiring caution to control fitting errors. Both approaches mitigate the need for dense frequency grids, which would otherwise scale poorly with system size. In periodic systems, the integration for GW quantities demands accurate k-point sampling to capture momentum-dependent effects. The Monkhorst-Pack scheme generates uniform grids of k-points offset from high-symmetry points, offering a straightforward and convergent method for averaging over the reciprocal space in insulators and semiconductors. For improved precision, especially in metals or when evaluating sums involving , the tetrahedron method linearly interpolates integrands within tetrahedral volumes of the , enhancing convergence for smooth functions and reducing sampling errors without increasing grid . These techniques ensure reliable corrections by balancing computational cost with the need for dense sampling in regions of high electronic . The naive implementation of GW scales as O(N^4) with system size N, arising from the four-index summation over occupied and unoccupied states in the self-energy matrix elements, which limits applications to small systems. Advanced techniques reduce this to O(N^3) by exploiting locality and sparsity. The Sternheimer approach, based on density response functions, eliminates explicit unoccupied state summations through perturbative equations solved iteratively, drastically lowering the prefactor for plane-wave calculations. Low-rank approximations further optimize by projecting the polarizability or dielectric matrix onto a reduced basis of singular vectors, enabling efficient handling of the screened interaction for large basis sets. These methods have enabled GW studies on systems exceeding hundreds of atoms while preserving benchmark accuracy.

Software Implementations

Several packages implement the approximation for electronic structure calculations, often building on (DFT) inputs and supporting various basis sets and numerical methods. ABINIT, a plane-wave-based code, provides full GW capabilities, including self-consistent iterations for energies, and is widely used for solids and nanostructures. Yambo, a post-DFT community code, specializes in GW for corrections and extends to the Bethe-Salpeter equation for optical spectra, interfacing seamlessly with plane-wave codes like . , through its PWscf module, serves as a DFT engine for generating inputs to GW codes such as Yambo or BerkeleyGW, enabling plane-wave GW calculations for periodic systems. BerkeleyGW is an independent, massively parallel open-source package designed specifically for and GW-Bethe-Salpeter calculations, taking mean-field wavefunctions from codes like ABINIT or as input, and excels in computing spectra and optical responses for materials. (an for "Without Empty States ") is another open-source tool focused on full-frequency computations, supporting large-scale simulations with k-point sampling and GPU acceleration for efficient handling of the screened interaction W. The GW100 set, comprising ionization potentials and electron affinities of 100 molecules computed at the G0W0@PBE level using multiple independent codes, serves as a standard for validating implementations across these packages. Commercial software also supports GW, particularly for specialized applications. VASP implements projector-augmented wave (PAW)-based for materials, offering partial self-consistency and low-scaling variants suitable for . For molecular systems, implementations using Gaussian basis sets, such as in the Gaussian and plane waves scheme, enable G0W0 and eigenvalue-self-consistent GW calculations, providing accurate energies for small to medium-sized molecules. As of 2025, recent developments include integrations to accelerate computations, such as predicting dielectric functions and density-density response for faster evaluation of the screened interaction W, with applications demonstrated in high-throughput workflows. These enhancements, often leveraging neural networks trained on first-principles data, reduce computational costs while maintaining accuracy for large datasets. Data-driven low-rank approximations for the electron-hole further enable efficient time-dependent calculations.

Applications

Electronic Structure Calculations

The GW approximation is widely applied to compute quasiparticle band structures, densities of states (), and ionization potentials in solids and molecules, providing corrections to (DFT) results that often suffer from underestimation. Starting from DFT wave functions and eigenvalues as input, GW calculations yield renormalized energies that better align with experimental photoemission and inverse photoemission spectra. These applications are particularly valuable for semiconductors, insulators, and molecular systems, where GW reveals the many-body effects influencing electronic excitations. In semiconductors and oxides, GW excels at predicting band gaps with high accuracy. For gallium arsenide (GaAs), a prototypical III-V semiconductor, GW calculations yield a direct band gap of approximately 1.5 eV at the Γ point, closely matching the experimental low-temperature value of 1.52 eV, whereas standard DFT underestimates it by about 0.5 eV. Similarly, for monoclinic hafnium dioxide (HfO₂), an important high-κ dielectric, GW predicts a band gap of approximately 5.7 eV, in excellent agreement with the experimental optical value of about 5.8 eV, highlighting GW's ability to capture the insulating nature beyond DFT's typical underestimation to around 3.5 eV. These improvements stem from the inclusion of screened exchange and correlation effects, which open the gap without altering the overall band dispersion significantly. GW-derived DOS further refines the valence and conduction band features, enabling precise identification of material properties like effective masses. For compounds, provides essential corrections to DFT quasiparticle spectra, particularly in strongly correlated systems like 3d oxides (e.g., , MnO). DFT often fails to reproduce the charge-transfer insulating behavior and underestimates band gaps due to self-interaction errors in d-orbitals, but renormalizes the spectra, shifting occupied d-states downward and revealing satellite peaks arising from multi-electron excitations and plasmons. These satellites, observed 5–10 below the main peaks in photoemission experiments, reflect the breakdown of the single-particle picture and are captured in spectral functions, offering insights into correlation strengths not accessible via DFT. potentials in these materials, computed as removal energies, align well with experiment, typically within 0.3 . In molecular systems, GW accurately determines HOMO-LUMO gaps, corresponding to the fundamental gaps (ionization potential minus ), surpassing DFT's delocalization errors. For (C₆H₆), a π-conjugated , G₀W₀ calculations yield a gap of 10.4 eV, in near-perfect agreement with the experimental value of 10.4 eV derived from gas-phase photoemission data. This precision extends to the GW100 dataset of 100 small , where G₀W₀@PBE achieves a of approximately 0.2 eV for ionization potentials and about 0.4 eV for , validating GW as a gold standard for molecular electronic structure with errors far below those of . The resulting molecular and orbital energies facilitate reliable predictions of reactivity and charging processes in .

Optical and Transport Properties

The GW approximation, through its computation of the screened interaction W and the non-interacting G, enables the evaluation of the irreducible P(\omega) within the (RPA). This directly enters the frequency-dependent dielectric function as \varepsilon(\omega) = 1 - v P(\omega), where v denotes the bare potential. The imaginary part of the inverse dielectric function, -\operatorname{Im} \varepsilon^{-1}(\omega), yields the loss spectrum, providing insights into plasmons and interband transitions in materials such as semiconductors and metals. For accurate absorption spectra, the GW-derived energies and wave functions serve as inputs to the Bethe-Salpeter (BSE), which accounts for the attractive electron-hole interaction mediated by [W](/page/W). The BSE describes excitonic bound states, often resulting in a red-shift of the onset compared to independent-particle approximations; in , this shift can exceed 1 eV due to strong localization and reduced screening. Such excitonic effects are crucial for interpreting and optical gaps in low-dimensional systems. In charge transport, GW calculations yield renormalized band dispersions, allowing extraction of effective masses via m^* = \hbar^2 \left( \frac{\partial^2 \varepsilon}{\partial k^2} \right)^{-1}, which inform carrier mobilities through Boltzmann transport theory. For instance, in _2, GW predicts electron effective masses around 0.45 m_e (where m_e is the free-electron mass), contributing to room-temperature mobilities up to 200 cm²/V·s, consistent with experimental values enhanced by reduced scattering in . Representative applications include GW-BSE studies of bulk CdTe, where the method captures the optical near 1.5 eV, incorporating excitonic binding energies of approximately 10 meV that align with experimental spectra by resolving discrete peaks below the quasiparticle gap of ~1.6 eV. This approach highlights 's versatility in bridging static electronic structure to dynamic response properties. Recent extensions as of 2025 include GW applications to halide perovskites for solar cells and topological insulators, improving predictions of defect states and surface properties.

History and Developments

Origins and Early Work

The GW approximation emerged from foundational work in , which provides a framework for treating electron-electron interactions in quantum many-body systems. Preceding the formal development of GW, the Bohm-Pines theory in the 1950s introduced concepts of collective excitations, or plasmons, and their role in screening the Coulomb interaction in the homogeneous electron gas, laying groundwork for later treatments of dynamic screening effects. The GW approximation was first formulated by Lars Hedin in his 1965 work on the one-particle for the electron gas, where he derived a set of coupled equations relating the Green's function G, the screened interaction W, and the \Sigma = iGW, applied as a perturbative scheme to compute properties in interacting electron systems. This initial calculation demonstrated the method's potential for the homogeneous electron gas, though computational limitations restricted its immediate broader use. Early applications followed in 1969, when Hedin and Stig Lundqvist extended the GW framework to study the homogeneous gas, incorporating - and -phonon s to evaluate one- properties such as effective masses and . Their work included derivations of -pole models to approximate the frequency dependence of the dielectric function, simplifying calculations of the screened W while capturing key many-body effects like satellites in the spectral function. The method saw a revival in the for ab initio calculations in solids, driven by advances in computational power that enabled applications beyond simple models, including the first fully ab initio calculations for by Hybertsen and Louie, which demonstrated its ability to accurately predict energies and band gaps, correcting the underestimation typical of , as later reviewed by Fanny Aryasetiawan and Olle Gunnarsson, who highlighted its efficacy for band structures in real materials.

Key Advances and Extensions

In the 1980s and 1990s, significant progress was made in applying the approximation to realistic materials. Concurrently, efforts to incorporate vertex beyond the basic GW were explored, which analyzed the impact of these on the self-energy and polarizability in semiconductors, revealing their role in improving for excited-state properties. The saw the development of efficient computational codes and approximations that broadened the applicability of GW methods. A comprehensive by Onida et al. highlighted the implementation of GW in codes like GW1, enabling routine calculations for a variety of solids and molecules with improved numerical stability. Refinements to the plasmon-pole model, originally proposed by Hybertsen and Louie in , were further advanced during this period to better approximate the frequency dependence of the screened interaction, reducing computational cost while maintaining accuracy for predictions in insulators. From the 2010s onward, self-consistent extensions of gained prominence, with the quasi-self-consistent (QSGW) approach developed by Kotani and Schilfgaarde providing a framework to iteratively update the and screened interaction, yielding more reliable electronic structures for metals and correlated systems without full inclusion. Additionally, techniques emerged to accelerate computations, such as models trained to predict the matrix elements directly from Kohn-Sham orbitals, as demonstrated in 2023 studies that achieved near- accuracy for energies in diverse materials with reduced scaling. By 2025, has enabled GW calculations for unprecedented system sizes, including supercells exceeding 1000 atoms, as showcased in implementations on platforms like and Aurora that leverage massively parallel algorithms for defect and transport properties in complex materials.

Limitations and Future Directions

Computational Challenges

The GW approximation imposes significant computational demands due to its formal scaling properties, which limit its applicability to relatively small systems. In canonical implementations, the computational time and memory requirements both scale as O(N^4), where N is the number of basis functions or orbitals, arising primarily from the evaluation of the screened Coulomb interaction and the self-energy matrix elements. This quartic scaling restricts routine GW calculations to systems of up to a few hundred atoms, such as small molecules, clusters, or unit cells of solids, beyond which the resource intensity becomes prohibitive without specialized optimizations. Achieving convergence in GW calculations presents additional hurdles, particularly with respect to basis set completeness and sampling. For plane-wave basis sets, convergence requires energy cutoffs exceeding 100 Ry to minimize basis set incompleteness errors, which can otherwise lead to systematic inaccuracies in quasiparticle energies. In metallic systems, dense k-point grids are essential to capture the adequately, as sparse sampling results in poor convergence of band structures and transport properties, often necessitating grids with hundreds of points for reliable results. The standard G_0W_0 approximation, while computationally lighter than fully self-consistent variants, introduces errors stemming from its perturbative nature and neglect of higher-order effects. It often slightly overestimates band gaps in semiconductors and insulators, with typical discrepancies of ~0.1 attributed to the omission of vertex corrections in the . These errors arise because the non-interacting G_0 and the independently screened interaction W_0 fail to fully account for dynamical screening and exchange-correlation beyond the random-phase approximation. Furthermore, the GW approximation exhibits gaps in its predictive power for strongly correlated electron systems, such as Mott insulators, where standard formulations inadequately describe localization and strong electron-electron interactions. In these cases, the method often overestimates metallic behavior or fails to capture insulating gaps without additional extensions, highlighting inherent limitations in its treatment of correlation effects.

Beyond-GW Methods

To address the limitations of the GW approximation in capturing strong electron correlations and many-body effects in complex systems, such as oxides or molecules with d/f electrons, beyond-GW methods introduce higher-order corrections or alternative formulations for improved energies and spectral functions. These approaches extend Hedin's equations by incorporating vertex functions or treatments, enhancing accuracy for systems where standard GW underestimates band gaps or fails to describe satellite structures. Vertex corrections represent a key beyond-GW strategy by including the vertex function Γ in the Σ to account for additional screening and beyond the random-phase . In the GWΓ , often implemented via a test-electron approach, the irreducible is modified to incorporate effects, leading to better descriptions of exchange-correlation holes and improved ionization potentials in semiconductors like . This approach has been shown to converge self-consistently through iterative solutions of Hedin's equations. Self-consistent schemes offer another pathway beyond perturbative , with the COHSEX (Coulomb Hole plus Screened ) approximation serving as a static limit that captures short-range effects more effectively. COHSEX decomposes the self-energy into a screened and a hole contribution, which approximates the dynamical screening in while remaining computationally lighter; it has been extended in self-consistent frameworks to stabilize calculations for insulators, demonstrating superior performance in predicting cohesive energies over standard for covalent solids. The expansion provides a beyond- method particularly suited for describing satellite structures in photoemission spectra, which arise from multi-plasmon excitations and are poorly captured by perturbative . By expanding the in a form using the self-energy as input, this approach generates multiple shake-up satellites with intensities and positions matching experiments, as demonstrated for photoemission spectra in simple metals like sodium and aluminum, where corrections shift satellite peaks by 10-15 eV with relative intensities accurate to within 20%. Applied to valence bands, plus reduces errors to below 0.05 eV for sp-bonded materials, offering a diagrammatic resummation that avoids divergences in standard for deeply bound states. Emerging hybrid methods as of 2025 integrate with (DMFT) to tackle , such as f-electron systems, by combining GW's band structure with DMFT's local correlations. In GW+DMFT schemes, the is partitioned into a GW-derived non-local part and a local DMFT solver, achieving spectral functions for Mott insulators like with band gaps accurate to 0.3 eV and proper description of Hubbard bands. Additionally, surrogates are being developed to approximate full vertex corrections, reducing computational cost while maintaining accuracy for molecular quasiparticles. For systems involving relativistic effects, extensions like GW+SOX incorporate spin-orbit coupling (SOC) through screened optimized exchange, enhancing accuracy for heavy elements with d/f orbitals. This method applies SOC within the self-energy calculation, yielding band splittings in semiconductors like HgTe within 0.05 eV of experiment, and proves particularly effective for topological insulators where standard GW neglects spin-dependent correlations.

References

  1. [1]
  2. [2]
  3. [3]
  4. [4]
    Part 1: An overview of available functionals - VASP
    The PBE0 band gap (1.84 eV) is larger than the PBE one (0.62 eV). As a general trend in many materials, hybrid functionals lead to larger band gap and partially ...
  5. [5]
  6. [6]
    New Method for Calculating the One-Particle Green's Function with ...
    A set of successively more accurate self-consistent equations for the one-electron Green's function have been derived.
  7. [7]
    Electronic excitations: density-functional versus many-body Green's ...
    Jun 7, 2002 · This review compares the theoretical and practical aspects of the two approaches and their specific numerical implementations.
  8. [8]
    Electron correlation in semiconductors and insulators: Band gaps ...
    Oct 15, 1986 · The full dielectric matrix is used to evaluate the self-energy operator in the GW approximation: the first term in an expansion of the self- ...
  9. [9]
  10. [10]
    The GW Compendium: A Practical Guide to Theoretical ... - Frontiers
    The GW approximation in electronic structure theory has become a widespread tool for predicting electronic excitations in chemical compounds and materials.
  11. [11]
  12. [12]
  13. [13]
  14. [14]
  15. [15]
  16. [16]
  17. [17]
  18. [18]
    GW - abinit
    GW. This page gives hints on how to perform a GW calculation, including self-consistency with the ABINIT package. Introduction¶.Missing: software open source Yambo VASP Quantum ESPRESSO BerkeleyGW WEST
  19. [19]
    BerkeleyGW – A Massively Parallel GW and GW-BSE Code
    The BerkeleyGW package takes mean-field solutions (for example, DFT wavefunctions and eigenvalues) as input for the many-body perturbation theory calculations.Documentation · Download · BerkeleyGW software package... · BerkeleyGW 2.1Missing: open source Yambo VASP WEST
  20. [20]
    GW100: Benchmarking G0W0 for Molecular Systems
    GW100 is a benchmark set of the ionization potentials and electron affinities of 100 molecules computed with the GW method using three independent GW codes and ...Introduction · The GW100 Set · Computational Methodology · Results
  21. [21]
    Practical guide to GW calculations - VASP Wiki
    May 2, 2025 · The GW approximation is an approximation to the self-energy. ... This method goes back to Hybertsen and Louie and can be achieved in two ways.
  22. [22]
    Accelerating GW calculations through machine-learned dielectric ...
    Oct 7, 2023 · The GW approach produces highly accurate quasiparticle energies, but its application to large systems is computationally challenging due to ...Missing: review | Show results with:review
  23. [23]
    Towards fully automated GW band structure calculations - Nature
    Jan 29, 2021 · We analyze a data set comprising 370 GW band structures of two-dimensional (2D) materials covering 14 different crystal structures and 52 chemical elements.Missing: review | Show results with:review
  24. [24]
    [PDF] Quasiparticle band structures and the GW approximation
    GaAs band structure. Direct band gap at Γ in LDA nonrelativistic. 1.1 eV ... experimental band gap. 0.0. 2.0. 4.0. 6.0. 8.0. 10.0 theoretical band gap. GW. LDA. C.
  25. [25]
    Practical GW scheme for electronic structure of 3d-transition-metal ...
    Here, we calculate the quasiparticle spectra of closed- and open-shell molecular anions with partially and completely filled 3d shells (shallow and deep 3d ...
  26. [26]
  27. [27]
    Excitonic Character in Optical Properties of Tetrahedral CdX (X = S ...
    Computed (with the G0W0–BSE method, represented by circles) and experimentally determined size-dependent optical absorption energy of the lowest exciton state ...<|control11|><|separator|>
  28. [28]
    Phys. Rev. 85, 338 (1952) - A Collective Description of Electron ...
    Collective vs Individual Particle Aspects of the Interactions. David Pines and David Bohm. Phys. Rev. 85, 338 – Published 15 January 1952.
  29. [29]
    [PDF] Systematic Vertex Corrections through Iterative Solution of Hedin's ...
    Feb 23, 1998 · The GW approximation neglects diagram- matic vertex corrections both in the polarization propagator and the self-energy. Its theoretical ...
  30. [30]
    Quasiparticle Self-Consistent Theory | Phys. Rev. Lett.
    Jun 6, 2006 · In the latter case the electron-hole mediated state making P = - i G × G is suppressed by the square of the renormalization factor Z × Z ...
  31. [31]
    [PDF] arXiv:2305.02990v3 [cond-mat.mtrl-sci] 24 Jul 2023
    Jul 24, 2023 · The GW approach produces highly accurate quasiparticle energies, but its application to large systems is computationally challenging, ...
  32. [32]
    Advancing Quantum Many-Body GW Calculations on Exascale ...
    Sep 26, 2025 · Here, we present innovative implementations of advanced GW methods within the BerkeleyGW package, enabling large-scale simulations on Frontier ...
  33. [33]
  34. [34]
    Low-Scaling GW with Benchmark Accuracy and Application to ...
    Feb 23, 2021 · In a conventional GW implementation, however, its computational cost is O(N4) in the system size N, which prohibits its application to many ...
  35. [35]
    Ab Initio Full Cell for Correlated Materials | Phys. Rev. X
    Apr 6, 2021 · We apply the full cell G W + DMFT method to compute the spectral properties of Si, the spectra, magnetic moments, and spin correlation functions ...
  36. [36]
    the Coulomb-hole plus screened exchange approach - arXiv
    Aug 27, 2020 · In this work we show the advantages of using the Coulomb-hole plus screened-exchange (COHSEX) approach in the calculation of potential energy surfaces.
  37. [37]
    Multiple Plasmon Satellites in Na and Al Spectral Functions from Ab ...
    Sep 9, 1996 · In this Letter we use the cumulant expansion approach to obtain the spectral functions of Na and Al from ab initio calculations including the effects of band ...<|separator|>
  38. [38]
    On the combined use of GW approximation and cumulant expansion ...
    Jul 5, 2016 · On the combined use of GW approximation and cumulant expansion in the calculations of quasiparticle spectra: The paradigm of Si valence bands.
  39. [39]
    ComDMFT v.2.0: Fully Self-Consistent ab initio GW+EDMFT ... - arXiv
    Oct 6, 2023 · ComDMFT v.2.0 is a computational package for studying correlated quantum materials using GW+EDMFT, treating correlated electrons within full GW ...Missing: hybrid 2020-2025
  40. [40]
    Implementation and Validation of Fully Relativistic GW Calculations
    Jun 22, 2016 · We present an implementation of G 0 W 0 calculations including spin–orbit coupling (SOC) enabling investigations of large systems, with thousands of electrons.Method · Bulk Semiconductor Calculations · Validity of Separate G0W0...Missing: SOX | Show results with:SOX