Fact-checked by Grok 2 weeks ago

Wannier function

Wannier functions are a complete, orthonormal set of localized wavefunctions used in to represent the electronic structure of periodic crystalline systems in real space, serving as the Fourier transforms of extended Bloch wavefunctions over the . Formally defined for a band index n and lattice vector \mathbf{R} as | \mathbf{R} n \rangle = \frac{V}{(2\pi)^3} \int_{\mathrm{BZ}} d\mathbf{k} \, e^{-i \mathbf{k} \cdot \mathbf{R}} | \psi_{n \mathbf{k}} \rangle, where V is the unit cell volume and the integral is over the first , they transform the delocalized momentum-space description of electrons into spatially confined orbitals centered at lattice sites. Introduced by Gregory H. Wannier in 1937 to analyze electronic excitation levels in insulating crystals, these functions bridge the gap between the periodic boundary conditions of and intuitive atomic-like orbitals, enabling a more chemically meaningful interpretation of solid-state phenomena. Although the original formulation provided a powerful real-space basis, the non-uniqueness of Wannier functions arises from the gauge freedom in the phases of Bloch states, which can lead to delocalized or non-intuitive forms. This issue was resolved in the modern era through the development of maximally localized Wannier functions (MLWFs), pioneered by Nicola Marzari and David Vanderbilt in 1997, who introduced a variational to minimize the quadratic spread of the functions in space while preserving . Subsequent extensions, such as disentanglement methods for entangled bands, have made MLWFs applicable to complex materials beyond isolated bands. Wannier functions have become indispensable in for a wide array of applications, including the analysis of chemical bonding and covalency in crystals like and perovskites, the computation of macroscopic electric via Wannier centers, and the evaluation of orbital in topological materials. They facilitate Wannier interpolation techniques for efficient band structure calculations on dense \mathbf{k}-point grids, reducing computational cost in first-principles simulations. Beyond electrons, the formalism extends to phonons, photonic crystals, and cold-atom lattices, while enabling model Hamiltonians for transport properties, electron-phonon couplings in superconductors, and studies of strongly correlated systems like cuprates. Integrated into software packages such as Wannier90 and , these functions continue to drive advancements in materials design and theoretical predictions.

Fundamentals

Definition

In , Wannier functions provide a localized representation of states in periodic lattices, serving as an alternative to the extended Bloch wave functions that describe delocalized electrons under the Bloch theorem. Specifically, a Wannier function for band index n and lattice site \mathbf{R} is defined as the of the corresponding Bloch functions \psi_{n\mathbf{k}}(\mathbf{r}) over the , applicable to an isolated band or an entangled group of bands. The mathematical expression for the Wannier function is given by w_{n\mathbf{R}}(\mathbf{r}) = \frac{1}{\sqrt{N}} \sum_{\mathbf{k}} e^{-i\mathbf{k}\cdot\mathbf{R}} \psi_{n\mathbf{k}}(\mathbf{r}), where N denotes the number of unit cells in the , the sum runs over discrete wave vectors \mathbf{k} in the , and the phase factor ensures localization. This construction yields functions centered at vectors \mathbf{R}, capturing the spatial extent of orbitals around sites while preserving the periodicity of the underlying potential. Wannier functions were introduced by Gregory Wannier in 1937 as a means to model localized excitations, such as electron-hole pairs, in insulating crystals by transforming the extended states into compact, site-specific orbitals.

Relation to Bloch Functions

Bloch functions serve as the fundamental solutions to the time-independent for electrons moving in a periodic crystal potential, characterized by their delocalized nature across the crystal lattice. These wave functions, labeled by band index n and crystal momentum \mathbf{k} in the first , take the form \psi_{n\mathbf{k}}(\mathbf{r}) = e^{i\mathbf{k}\cdot\mathbf{r}} u_{n\mathbf{k}}(\mathbf{r}), where u_{n\mathbf{k}}(\mathbf{r}) is a periodic function with the lattice periodicity, ensuring \psi_{n\mathbf{k}}(\mathbf{r} + \mathbf{R}) = e^{i\mathbf{k}\cdot\mathbf{R}} \psi_{n\mathbf{k}}(\mathbf{r}) for any lattice vector \mathbf{R}. This Bloch theorem underpins the band structure of solids, describing extended states that reflect the of the infinite crystal. Wannier functions provide a complementary localized , obtained through a unitary of the Bloch functions over the . Specifically, the Wannier function w_{n\mathbf{R}}(\mathbf{r}) centered at lattice site \mathbf{R} for n is given by w_{n\mathbf{R}}(\mathbf{r}) = \frac{1}{\sqrt{N}} \sum_{\mathbf{k}} e^{-i\mathbf{k}\cdot\mathbf{R}} \psi_{n\mathbf{k}}(\mathbf{r}), where N is the number of unit cells and the sum runs over the discrete \mathbf{k}-points in the . The inverse relation expresses each Bloch function as a superposition of Wannier functions: \psi_{n\mathbf{k}}(\mathbf{r}) = \frac{1}{\sqrt{N}} \sum_{\mathbf{R}} e^{i\mathbf{k}\cdot\mathbf{R}} w_{n\mathbf{R}}(\mathbf{r}). This duality preserves the and of the basis, allowing seamless transitions between momentum-space (Bloch) and real-space (Wannier) descriptions while maintaining the periodic structure of the . The applicability of this depends on the structure. For an isolated energy , where states at different \mathbf{k} do not mix with other bands, the Wannier functions are directly constructed from the Bloch states of that single , yielding well-defined localized orbitals. However, in cases of entangled or composite bands—where bands overlap or hybridize across the —direct leads to delocalized Wannier functions; disentanglement procedures are required to project onto an optimal of Bloch states, enabling the formation of localized functions for effective multi-band descriptions. A key aspect of this relation is the gauge freedom inherent in the Bloch basis. Bloch functions are defined up to a \mathbf{k}-dependent , \psi_{n\mathbf{k}}(\mathbf{r}) \to e^{i\phi_n(\mathbf{k})} \psi_{n\mathbf{k}}(\mathbf{r}), which does not alter their physical eigenvalues but influences the resulting Wannier functions' spatial localization and . This unitary ambiguity allows for optimization of Wannier localization by selecting appropriate phases, bridging the delocalized Bloch picture to tailored real-space representations suited for tight-binding models or local property calculations.

Properties and Formulation

Key Properties

Wannier functions possess several fundamental properties that render them invaluable for describing electronic states in periodic solids, bridging the gap between delocalized Bloch functions and localized representations. These attributes include orthogonality, completeness, and translation invariance, which ensure they form a robust basis for expansions. Additionally, their real-space localization facilitates the study of local physical phenomena, while their matrix elements with the enable connections to tight-binding models for energy calculations. A defining feature of Wannier functions is their , which allows them to serve as an without overlap between functions centered at different sites or s. Mathematically, this is expressed as \int w_{n\mathbf{R}}^*(\mathbf{r}) w_{m\mathbf{R'}}(\mathbf{r}) \, d\mathbf{r} = \delta_{nm} \delta_{\mathbf{R}\mathbf{R'}}, where w_{n\mathbf{R}}(\mathbf{r}) denotes the Wannier function for band index n and lattice vector \mathbf{R}, ensuring non-overlapping contributions in expansions of wavefunctions or operators. This property arises from their construction as unitary transformations of Bloch functions and holds for Wannier functions, providing a complete and non-redundant set for representing electronic states. Wannier functions also form a complete basis for the of periodic systems, spanning the same as the underlying Bloch functions through the projection P = \sum_{\mathbf{R}} |w_{n\mathbf{R}}\rangle \langle w_{n\mathbf{R}} |. This completeness guarantees that any state within the relevant energy bands can be exactly reconstructed from the Wannier basis, offering a localized alternative to momentum-space descriptions without loss of information. It underpins their utility in rigorous expansions, such as those for operators or response functions in . Translation invariance is another core property, reflecting the periodicity of the crystal lattice: translating a Wannier function by a lattice vector \mathbf{R} yields another Wannier function in the set, such that w_{n, \mathbf{0}}(\mathbf{r} - \mathbf{R}) = w_{n\mathbf{R}}(\mathbf{r}). This symmetry ensures that the functions are equivalent up to lattice shifts, facilitating Bloch-like superpositions via Fourier transforms, \psi_{n\mathbf{k}}(\mathbf{r}) = \sum_{\mathbf{R}} e^{i \mathbf{k} \cdot \mathbf{R}} w_{n\mathbf{R}}(\mathbf{r}), and preserves the translational symmetry of the underlying Hamiltonian. In contrast to the extended nature of Bloch functions, Wannier functions provide a real-space representation that is particularly advantageous for analyzing local phenomena, such as chemical bonding, defect interactions, or localized excitations in solids. Their spatial concentration around atomic sites allows for intuitive interpretations of and overlaps in real space, making them suitable for modeling short-range interactions that are challenging in momentum space. For instance, this localization aids in visualizing orbital hybridization or local response to perturbations like impurities. Finally, Wannier functions connect directly to energy properties through expectation values of the Hamiltonian, forming the basis for tight-binding models. The on-site and hopping matrix elements, \langle w_{n\mathbf{0}} | \hat{H} | w_{m\mathbf{R}} \rangle, quantify local energies and inter-site couplings, respectively, enabling the Hamiltonian in the Wannier basis to be expressed as H^W_{\mathbf{k}, nm} = \sum_{\mathbf{R}} e^{i \mathbf{k} \cdot \mathbf{R}} \langle w_{n\mathbf{0}} | \hat{H} | w_{m\mathbf{R}} \rangle. These elements approximate band structures via nearest-neighbor interactions, providing a simplified yet accurate framework for computing energy eigenvalues and electronic properties in periodic systems.

Mathematical Derivation

The time-independent Schrödinger equation for an electron in a periodic crystal potential is given by H \psi(\mathbf{r}) = E \psi(\mathbf{r}), where the Hamiltonian H = -\frac{\hbar^2}{2m} \nabla^2 + V(\mathbf{r}) and the potential satisfies V(\mathbf{r} + \mathbf{R}) = V(\mathbf{r}) for any lattice vector \mathbf{R}. This periodicity implies translational symmetry, leading to solutions labeled by a band index n and a wavevector \mathbf{k} in the first Brillouin zone (BZ). According to , the eigenfunctions can be expressed as Bloch states \psi_{n\mathbf{k}}(\mathbf{r}) = e^{i \mathbf{k} \cdot \mathbf{r}} u_{n\mathbf{k}}(\mathbf{r}), where u_{n\mathbf{k}}(\mathbf{r}) is a with the periodicity, i.e., u_{n\mathbf{k}}(\mathbf{r} + \mathbf{R}) = u_{n\mathbf{k}}(\mathbf{r}). Substituting into the yields an effective equation for u_{n\mathbf{k}}: \left[ \frac{ (\mathbf{p} + \hbar \mathbf{k})^2 }{2m} + V(\mathbf{r}) \right] u_{n\mathbf{k}}(\mathbf{r}) = E_{n\mathbf{k}} u_{n\mathbf{k}}(\mathbf{r}), where \mathbf{p} = -i \hbar \nabla. These Bloch states form a complete, for the of the crystal. Wannier functions are obtained by performing a of the Bloch states from the reciprocal-space to real space, centered at sites. For a finite crystal with N unit cells of \Omega, the Wannier function for n centered at vector \mathbf{R}_m is w_{n\mathbf{R}_m}(\mathbf{r}) = \frac{1}{N} \sum_{\mathbf{k}} e^{-i \mathbf{k} \cdot \mathbf{R}_m} \psi_{n\mathbf{k}}(\mathbf{r}), where the sum is over N discrete \mathbf{k}-points in the , and the factor $1/N ensures , incorporating the unit cell implicitly through the BZ discretization (in the , this becomes \frac{\Omega}{(2\pi)^3} \int_{\mathrm{BZ}} d\mathbf{k}). The inverse transformation reconstructs the Bloch states: \psi_{n\mathbf{k}}(\mathbf{r}) = \sum_m e^{i \mathbf{k} \cdot \mathbf{R}_m} w_{n\mathbf{R}_m}(\mathbf{r}). This unitary preserves : \langle w_{n\mathbf{R}_m} | w_{n'\mathbf{R}_{m'}} \rangle = \delta_{nn'} \delta_{mm'}, as can be verified by substituting the expressions and using the completeness of the Bloch basis along with the of plane waves \frac{1}{N} \sum_{\mathbf{k}} e^{i \mathbf{k} \cdot (\mathbf{R}_m - \mathbf{R}_{m'})} = \delta_{mm'}. For isolated bands, the transformation is fully unitary. However, in cases of composite bands (e.g., a set of J bands forming an analytically connected ), the Bloch states within the subspace are related by a \mathbf{k}-dependent unitary matrix U_{mn}(\mathbf{k}), allowing the Wannier functions to be projected onto this subspace: w_{n\mathbf{R}_m}(\mathbf{r}) = \frac{1}{N} \sum_{\mathbf{k}} e^{-i \mathbf{k} \cdot \mathbf{R}_m} \sum_{m'=1}^J U_{m'n}(\mathbf{k}) \psi_{m'\mathbf{k}}(\mathbf{r}). This ensures the Wannier functions span the desired while maintaining unitarity within it, though the full projection may require disentanglement procedures to isolate smooth, isolated manifolds from entangled bands. The resulting basis remains orthogonal and complete within the subspace, facilitating real-space representations of operators like the .

Localization

Localization Criteria

Wannier functions, derived from Bloch states through transformation, are not uniquely defined due to the freedom inherent in the choice of phases for the Bloch functions at each wavevector \mathbf{k}. This arbitrariness allows for various representations, but localization refers to selecting a that minimizes the spatial spread of the Wannier functions, concentrating them around sites to resemble or molecular orbitals. Early efforts to achieve such locality date back to Gregory Wannier's original formulation, where he proposed phase choices to construct localized functions for insulating crystals, though explicit criteria were limited. Subsequent work, such as Kohn's analysis in one dimension, emphasized selecting phases to ensure exponential localization for isolated bands. A quantitative measure of localization is provided by the spread functional \Omega = \sum_n \left[ \langle r^2 \rangle_n - \langle \mathbf{r} \rangle_n^2 \right], which sums the variances of the position operator over the Wannier functions labeled by band index n. This functional decomposes into an invariant part \Omega_I, which is independent of the gauge and reflects the underlying band dispersion in \mathbf{k}-space, and a localizable part \tilde{\Omega} = \Omega_D + \Omega_{OD}, where \Omega_D captures diagonal contributions from individual Wannier centers and \Omega_{OD} accounts for off-diagonal overlaps between different sites or bands; minimizing \tilde{\Omega} achieves optimal localization. Criteria for "good" localization typically require the Wannier functions to decay rapidly away from their centers, with characterizing well-localized functions in insulating materials, modulated by a power-law prefactor. In metallic systems, however, the decay is algebraic (power-law), reflecting the delocalized nature of states and limiting the degree of achievable localization. The extent of localization is further constrained by symmetries, which impose relations on the Bloch phases, and by topological invariants, such as Chern numbers, that can obstruct exponential localization in non-trivial structures.

Maximally Localized Wannier Functions

The Marzari-Vanderbilt algorithm, introduced in , provides an to construct maximally localized Wannier functions (MLWFs) by minimizing a spread functional through unitary rotations of Bloch states within a composite energy band. This approach addresses the non-uniqueness of Wannier functions by optimizing their localization in real space, making them suitable for applications requiring spatially compact orbitals. For isolated bands, the algorithm directly applies unitary transformations to minimize the total spread \Omega = \Omega_I + \tilde{\Omega}, where \Omega_I is the invariant part depending only on the Bloch states, and \tilde{\Omega} is the non-invariant part adjustable via gauge choices. The minimization proceeds using a method in the space of unitary matrices, with updates derived from the functional's gradients to iteratively reduce \tilde{\Omega} while preserving \Omega_I. When dealing with entangled bands—where states mix across the desired energy window—a preliminary disentanglement procedure projects the Bloch states onto a of interest, selecting an optimal set of N states from a larger manifold of M > N bands to enable subsequent localization. This step ensures the resulting MLWFs correspond to physically meaningful orbitals without artificial delocalization from band entanglement. In semiconductors like , the Marzari-Vanderbilt method yields MLWFs resembling sp^3-hybridized orbitals centered on atomic sites, with spreads on the order of a few squared, capturing the tetrahedral bonding geometry effectively. Recent refinements, particularly post-2020, have extended the framework to systems with non-collinear magnetism and spin-orbit coupling by incorporating spinor-valued Wannier functions and projectability-disentangled protocols that enhance robustness in automated workflows. These advancements achieve high-fidelity interpolations, with average errors below 15 meV up to 2 eV above the across diverse magnetic materials.

Rigorous Results

A foundational result in the theory of Wannier functions is Nenciu's theorem, which establishes the existence of exponentially localized Wannier functions for analytic Bloch bands in insulators. Specifically, for a nondegenerate, isolated energy band in a periodic crystal potential that is analytic in the position variable, there exist Wannier functions that decay exponentially at infinity, ensuring spatial localization. This theorem applies to arbitrary-dimensional crystals and resolves long-standing questions about the realizability of such functions beyond one dimension. Topological properties of the band structure can impose obstructions to the existence of exponentially localized Wannier functions. In Chern insulators, where the Chern number of an isolated band is nonzero, the non-trivial phase around the prevents the construction of exponentially decaying Wannier functions, as the integrated Berry curvature leads to delocalized charge centers. This obstruction arises fundamentally from the of the Bloch bundle, rendering Wannier functions ill-defined in the exponential sense for such systems. For multi-band cases, the existence of localized Wannier functions is analyzed using algebraic and the concept of stable equivalence. Under stable equivalence, a set of bands admits compactly supported Wannier functions it is equivalent to a trivial bundle in the K-group of the , allowing the addition of auxiliary trivial bands to achieve localization without altering the essential spectrum. This framework classifies multi-band insulators where direct localization may fail but can be restored through stable trivialization. Recent developments since 2010 have extended these results to , particularly addressing fragile topology. In fragile topological phases, an isolated set of bands exhibits an obstruction to exponentially localized Wannier functions due to nontrivial in the Wilson loop spectrum, yet this obstruction is destroyed upon adding sufficient trivial bands, distinguishing fragile phases from stable topological ones like Chern insulators. These conditions highlight how subtle topological features in materials such as twisted can permit or forbid localization depending on the band subspace considered. Proofs of these localization results often rely on Kato's analytic perturbation theory to handle the analyticity of Bloch eigenfunctions. By treating the crystal potential as a perturbation around the free-electron Hamiltonian, Kato's theory guarantees the analytic continuation of projectors onto isolated bands across the , enabling the construction of a smooth gauge that yields exponentially decaying Wannier functions via . This approach underpins Nenciu's theorem and extensions to perturbed systems, ensuring rigorous control over decay rates in analytic insulators.

Applications

Theory of Polarization

The modern theory of in crystalline solids, introduced by King-Smith and Vanderbilt in 1993, resolves the historical ambiguity in defining electric by formulating it as a topological quantity derived from the Berry phase of Bloch wavefunctions. This approach treats not as a simple of , which suffers from branch-choice dependence, but as a geometric phase accumulated over the (BZ), defined modulo the quantum e \mathbf{R}/\Omega, where e is the electron charge, \mathbf{R} is a lattice vector, and \Omega is the unit cell volume. Central to this theory is the connection to Wannier functions, which provide a localized of the states, allowing to be interpreted as the displacement of Wannier charge centers relative to ionic positions. The electronic contribution to the vector is expressed as \mathbf{P}_{el} = \frac{e}{(2\pi)^3} \sum_n \int_{\mathrm{BZ}} d^3\mathbf{k} \, \mathbf{A}_n(\mathbf{k}), where the sum runs over occupied bands n, and \mathbf{A}_n(\mathbf{k}) = i \langle u_{n\mathbf{k}} | \nabla_{\mathbf{k}} u_{n\mathbf{k}} \rangle is the Berry connection defined with respect to the cell-periodic Bloch functions u_{n\mathbf{k}}. This integral yields the Berry phase, and equivalently, \mathbf{P}_{el} = \frac{e}{\Omega} \sum_n \bar{\mathbf{r}}_n, where \bar{\mathbf{r}}_n = \langle w_n | \mathbf{r} | w_n \rangle is the center of the Wannier function w_n for band n. The total includes the ic part \mathbf{P}_{ion} = \frac{e}{\Omega} \sum_s Z_s \mathbf{R}_s, with Z_s the charge and \mathbf{R}_s the position of s. In insulating , where bands are separated from conduction bands, this formulation ensures a unique and gauge-invariant ; in metals, however, crossings render it ill-defined due to incomplete band filling. A striking feature emerges in one-dimensional insulating chains, where exhibits quantization. For example, in the Su-Schrieffer-Heeger model—a paradigmatic 1D tight-binding chain with alternating hoppings—the spontaneous takes discrete values of 0 (trivial ) or e/(2a) (topological ), corresponding to the position of Wannier centers at bond centers or atomic sites, respectively; here, a is the . This quantization reflects the topological invariant of the band structure and has been pivotal in understanding symmetry-protected topological phases. Wannier-based polarization calculations have enabled accurate predictions of ferroelectric and piezoelectric properties in materials. In their seminal work, King-Smith and Vanderbilt computed the spontaneous of ferroelectric KNbO₃ as 0.35 C/m², closely matching the experimental value of 0.37 C/m², demonstrating the theory's predictive power for macroscopic . For , the response tensor e_{ijk} = \frac{\partial P_i}{\partial \epsilon_{jk}} (with \epsilon ) is obtained by tracking shifts in Wannier centers under applied strain, allowing efficient first-principles evaluation in complex perovskites like BaTiO₃. Compared to direct Berry phase computations on the full BZ, the Wannier approach offers advantages in interpretability—visualizing charge displacements—and computational efficiency, particularly when using maximally localized Wannier functions to handle entangled bands.

Wannier Interpolation

Wannier interpolation is a computational technique that exploits the localized nature of Wannier functions to efficiently evaluate electronic properties throughout the by constructing a tight-binding-like from first-principles data obtained on a coarse k-point grid. This approach begins with calculations, such as (DFT), performed on a sparse set of k-points to determine the Bloch states and overlaps within an energy window of interest. These are then projected onto a set of maximally localized Wannier functions (MLWFs), yielding matrix elements of the in the Wannier basis that capture the real-space hopping between localized orbitals. The resulting tight-binding model allows interpolation to arbitrarily dense k-grids via Fourier transformation, drastically reducing the computational cost compared to direct on fine meshes, which would otherwise require prohibitive resources for properties demanding high k-point resolution. The core of the method lies in the representation of the in the Wannier basis. The matrix elements are defined as H_{mn}(\mathbf{R}) = \langle w_{m\mathbf{0}} | H | w_{n\mathbf{R}} \rangle, where w_{m\mathbf{0}} and w_{n\mathbf{R}} are the MLWFs centered at the origin and lattice vector \mathbf{R}, respectively, and H is the full . The k-dependent Hamiltonian is then obtained by Bloch summation: H_{mn}(\mathbf{k}) = \sum_{\mathbf{R}} e^{i\mathbf{k}\cdot\mathbf{R}} H_{mn}(\mathbf{0},\mathbf{R}), which yields the interpolated band energies and wavefunctions upon at any \mathbf{k}. This formulation ensures gauge invariance for smooth quantities and leverages the of Wannier overlaps for . The accuracy of the interpolation hinges on the quality of the MLWFs, typically generated via the Marzari-Vanderbilt scheme. This technique finds broad applications in , particularly for plotting detailed band structures, computing densities of states (DOS), and evaluating such as spectra, all without the need for exhaustive DFT calculations on dense grids. For instance, it enables the efficient visualization of band dispersions in complex materials like topological insulators or oxides, where fine k-sampling is essential for capturing subtle features near band edges. In optical contexts, interpolated velocity matrix elements facilitate the computation of interband transitions, providing insights into excitonic effects or photovoltaic responses with high resolution. The precision of Wannier interpolation depends strongly on the localization of the Wannier functions; well-localized bases yield errors that decay exponentially with increasing coarse-grid density, achieving sub-meV accuracy for energies in insulators. However, in metals, where details introduce stronger k-dependence, errors can be larger—typically on the order of 10-50 meV—necessitating finer initial grids or approaches to mitigate artifacts from avoided crossings. Despite these challenges, the method outperforms traditional schemes, especially for entangled s. Recent extensions have broadened Wannier interpolation to linear and nonlinear response functions, enabling the computation of constants and advanced with enhanced efficiency. For example, formulations incorporating long-range electron-phonon interactions have been applied to dielectric screening in two-dimensional materials, capturing polar effects beyond standard DFT. Similarly, adaptive high-order schemes now support precise evaluation of tensors, integrating seamlessly with many-body for corrections. These developments, emerging post-2020, underscore the method's evolving role in simulating realistic device responses under finite temperatures or external fields.

Tight-Binding Models

Wannier functions provide a localized basis for constructing effective tight-binding models of the electronic structure in crystalline solids, where the is expressed in terms of matrix elements between these functions. The core of such models are the hopping integrals, defined as t_{mn}(\mathbf{R}) = \langle w_{m\mathbf{0}} | H | w_{n\mathbf{R}} \rangle, which quantify the between Wannier orbitals w_{m\mathbf{0}} at the origin and w_{n\mathbf{R}} at lattice vector \mathbf{R}. These integrals primarily capture nearest-neighbor interactions, enabling a simplified representation of band structures while preserving key physical features from first-principles calculations. In semiconductors like , maximally localized Wannier functions resemble sp³ hybrid orbitals, facilitating tight-binding models that accurately describe valence band bonding and conduction properties. For transition metals and their compounds, such as oxides, Wannier functions centered on d-orbitals allow modeling of partially filled bands, incorporating correlation effects through extensions like the . These examples highlight how Wannier-based tight-binding captures atomic-like localization while accounting for crystalline periodicity. The primary advantages of Wannier tight-binding models lie in their interpretable parameters, which directly relate to physical processes: hopping terms govern like , on-site energies influence magnetic ordering in correlated systems, and defect perturbations can be modeled by modifying local integrals. This interpretability aids in understanding complex behaviors without full overhead. However, the models assume short-range hopping, limiting validity to systems where long-range interactions are negligible, such as insulators or narrow-gap materials; delocalized states in metals may require larger basis sets. Recent advancements integrate Wannier tight-binding into frameworks for materials discovery, where learned potentials parameterize hoppings to predict properties across composition spaces, as seen in post-2022 developments for defect landscapes and quantum simulations. These approaches leverage the sparsity of Wannier representations to train efficient models on high-throughput .

Computational Methods

Numerical Implementation

The numerical implementation of Wannier functions commences with a density functional theory (DFT) calculation of the Bloch states on a finite Monkhorst-Pack k-point grid, providing the overlap matrices and Hamiltonian projections needed for subsequent Wannierization. This workflow proceeds in two primary stages: first, disentanglement to isolate a smooth subspace of interest from potentially entangled bands, and second, unitary optimization to minimize the quadratic spread functional, thereby obtaining maximally localized Wannier functions (MLWFs). The disentanglement step employs energy windows—an outer window encompassing the full set of bands and an inner (frozen) window defining the target subspace—to project trial localized functions (e.g., atomic-like orbitals) onto the Bloch basis, iteratively refining the subspace via singular value decomposition or similar techniques until convergence. Following disentanglement, the unitary optimization applies a steepest-descent algorithm on the unitary rotations at each k-point, updating the gauge via exponential maps to reduce the spread, typically converging within 10–50 iterations for isolated bands. Finite k-point grids introduce discretization errors that affect the accuracy of the Wannier functions, particularly the computed spreads, necessitating with respect to grid density. For instance, in semiconductors like , the total spread Ω converges rapidly with grids denser than 8×8×8, achieving values below 1 Ų, but coarser grids (e.g., 4×4×4) can overestimate delocalization by up to 20%. In systems with complex band structures, such as oxides, denser grids (12×12×12 or higher) are required to resolve fine features in the , with adaptive sampling schemes improving efficiency for non-uniform . The choice of grid balances computational cost in the initial DFT step against the precision of interpolated properties, often verified by monitoring the invariance of MLWF centers and spreads. Significant challenges arise in systems with band entanglement, particularly those involving d or f electrons, where hybridization across energy levels precludes simple band isolation. In transition metals like or perovskites with d orbitals, bands from different angular momenta overlap extensively, requiring wide disentanglement windows (e.g., spanning 10–20 ) and careful projection choices to avoid spurious delocalization, with spreads increasing by factors of 2–5 if entanglement is mishandled. For metallic systems, partial band occupations demand smearing techniques (e.g., Fermi-Dirac or Gaussian with widths of 0.01–0.1 Ry) during the DFT stage to stabilize the electronic structure, followed by frozen inner windows up to or near the to exclude conduction states while preserving metallic character. These approaches mitigate discontinuities but can introduce sensitivity to smearing parameters, potentially altering Wannier spreads by 10–15% if not converged. Post-processing of the resulting MLWFs involves diagnostic quantities such as the total spread Ω (decomposed into gauge-invariant and off-diagonal components) to assess localization quality, typically targeting Ω < 5 Ų for well-localized functions. Projections onto atomic orbitals are evaluated via overlap integrals between the MLWFs and pseudo-atomic basis sets, quantifying chemical bonding (e.g., sp³ hybridization in yields projection centers aligned with bonds). These steps facilitate the construction of tight-binding Hamiltonians by interpolating elements on denser grids, with spreads serving as a criterion for the overall procedure. Recent advances have enabled efficient computation for large systems exceeding 1000 atoms, incorporating GPU acceleration through hybrid algorithms like selected columns of the (SCDM) combined with MLWF optimization, achieving speedups of 5–10× on multi-GPU clusters for supercells up to 2000 atoms. These methods reduce the iterative burden of traditional disentanglement, making Wannierization feasible for disordered alloys or defects where CPU-based approaches scale poorly (O(N³) for N bands).

Software Tools

Wannier90 serves as the foundational open-source library for computing maximally localized Wannier functions (MLWFs), enabling the generation of localized basis sets from delocalized Bloch states obtained from first-principles calculations. It supports key features such as the disentanglement procedure for handling entangled bands and extensive post-processing capabilities for deriving properties like curvatures and tight-binding Hamiltonians. The library interfaces seamlessly with major electronic structure codes, including for plane-wave pseudopotential methods and for all-electron and projector-augmented wave approaches, allowing users to compute MLWFs directly within standard workflows. Complementing Wannier90, specialized tools extend its functionality for specific applications. WanT provides an integrated suite for studying coherent electronic transport in nanodevices, leveraging Wannier functions to compute conductance and in one-dimensional systems. Postw90, included as a module within Wannier90, facilitates the calculation of advanced properties such as orbital magnetizations, quantum metric tensors, and spin Hall conductivities from the output MLWFs. For developers seeking to embed Wannier functionality into custom codes, the libwannier library offers a modular , enabling parallel (MPI) invocations and integration with external programs for tasks like wavefunction projections. The Wannier90 ecosystem integrates with broader frameworks to streamline workflows. Libraries such as the Atomic Simulation Environment (ASE) support reading and writing Wannier90 file formats, facilitating structure manipulation and automation in Python-based pipelines. Similarly, pymatgen provides handling for Wannier90 data, including Unk files for real-space wavefunctions, enabling seamless incorporation into and analysis tools. Recent versions, including v3.1.0 released in March 2020, introduce support for spinor wavefunctions via the selected columns of density matrix (SCDM) method and topological invariants like spin Hall , enhancing applicability to spin-orbit coupled and non-collinear systems. Ongoing developments toward version 4.0, discussed at the Wannier Developers Meeting in February 2024 at the , aim to expand library interfaces and automation features. Benchmarks demonstrate the high accuracy of Wannier90-generated MLWFs in representative materials. In , MLWFs accurately interpolate metallic bands and van Hove singularities across the , achieving near-perfect agreement with dense k-point sampling even from coarse grids, with spread values on the order of 10^{-2} Ų. For GaAs, a prototypical III-V , Wannier interpolation reproduces valence and conduction band structures with errors below 10 meV relative to direct DFT calculations, validating its use in tight-binding models for optoelectronic properties. Community-driven developments sustain the ecosystem's growth through open-source contributions on and collaborative events. The Wannier90 repository has seen numerous pull requests addressing optimizations and bug fixes since 2020, fostering robustness across diverse platforms. The Wannier 2022 Summer School and Developers Meeting at ICTP united over 100 researchers to advance integration with emerging codes and automated Wannierization protocols. A comprehensive review of the Wannier software ecosystem was published in 2024.

References

  1. [1]
    Maximally localized Wannier functions: Theory and applications - arXiv
    Dec 22, 2011 · A unique set of maximally localized Wannier functions, accomplishing the solid-state equivalent of constructing localized molecular orbitals, or "Boys orbitals ...Missing: definition | Show results with:definition
  2. [2]
  3. [3]
    Maximally localized Wannier functions: Theory and applications
    Oct 10, 2012 · In the earlier solid-state theory literature, WFs were typically introduced in order to carry out some formal derivation, for example, of the ...
  4. [4]
    The Structure of Electronic Excitation Levels in Insulating Crystals
    In this article, a method is devised to study the energy spectrum for an excited electron configuration in an ideal crystal.
  5. [5]
  6. [6]
    Wannier90 as a community code: new features and applications
    Jan 23, 2020 · In this article we describe these new functionalities, that include the implementation of new features for wannierisation and disentanglement ( ...New features for... · Enhancements in functionality · New post-processing features
  7. [7]
    Robust Wannierization including magnetization and spin-orbit ...
    Jul 9, 2025 · Here, we extend the applicability of PDWFs to magnetic systems and/or those including spin-orbit coupling, and implement such extensions in ...
  8. [8]
    Existence of the exponentially localised Wannier functions
    As a consequence the existence of exponentially localised Wannier functions corresponding to nondegenerated bands of arbitrary three-dimensional crystals is ...Missing: Nenciu's | Show results with:Nenciu's
  9. [9]
    Fragile Topology and Wannier Obstructions | Phys. Rev. Lett.
    Sep 18, 2018 · Marzari, A. A. Mostofi, J. R. Yates, I. Souza, and D. Vanderbilt, Maximally localized wannier functions: Theory and applications, Rev. Mod ...
  10. [10]
    Compactly supported Wannier functions and algebraic $K$-theory
    Mar 20, 2017 · A Wannier function is a wave packet in position space, taking values in the space spanned by the orbitals for each site; thus it is a vector ...
  11. [11]
    Theory of polarization of crystalline solids | Phys. Rev. B
    Jan 15, 1993 · We show that physically ΔP can be interpreted as a displacement of the center of charge of the Wannier functions.
  12. [12]
    [PDF] Theory of Polarization: A Modern Approach - Rutgers Physics
    May 2, 2025 · The modern theory of polarization provides the correct definition of Ps, as well as the theoretical framework allowing one to compute it from.
  13. [13]
    [PDF] Wannier function and polarization 1 Free fermions - Kohei Kawabata
    Nov 9, 2024 · In one dimension, this localization exhibits exponential decay in band insulators and algebraic decay in band metals [3]. Example.—As the ...
  14. [14]
    [PDF] Dynamical mean-field theory using Wannier functions
    Sep 29, 2006 · Hopping integrals between the t2g Wannier orbitals of BaVS3 in the maximally localized basis. The term 001. 2 shall denote the hopping to the ...
  15. [15]
    Database of Wannier tight-binding Hamiltonians using high ... - Nature
    Apr 13, 2021 · Wannier tight-binding Hamiltonians (WTBH) provide a computationally efficient way to predict electronic properties of materials.
  16. [16]
    A machine learning approach to predict tight-binding parameters for ...
    Jun 11, 2025 · The method opens a path towards understanding complicated defect landscapes using a computationally affordable semi-empirical approach without sacrificing ...
  17. [17]
  18. [18]
    None
    Below is a merged summary of the recent extensions of Wannier interpolation for response functions post-2020, consolidating all information from the provided segments into a single, detailed response. To maximize clarity and density, I will use a table format where appropriate to organize key details, followed by a narrative summary that integrates the remaining information. The response retains all references, URLs, and specific advancements mentioned across the segments.
  19. [19]
  20. [20]
  21. [21]
    QEF/want - GitHub
    WanT is an open-source, GNU General Public License suite of codes that provides an integrated approach for the study of coherent electronic transport in ...
  22. [22]
    wannier-developers/wannier90: Official repository of the ... - GitHub
    The Maximally-Localised Generalised Wannier Functions Code. The homepage of the Wannier90 code is https://www.wannier.org. The code is hosted on GitHub.
  23. [23]