Fact-checked by Grok 2 weeks ago

Hubbard model

The Hubbard model is a foundational theoretical framework in that describes the behavior of strongly interacting on a , capturing essential phenomena such as metal-insulator transitions driven by electron correlations. Introduced independently in 1963 by John Hubbard, Martin Gutzwiller, and Junjiro Kanamori, it extends the tight-binding model by incorporating an on-site repulsion term U that penalizes double occupancy of sites, alongside a hopping term -t allowing to move between nearest-neighbor sites. The model's takes the form
H = -t \sum_{\langle i,j \rangle, \sigma} (c_{i\sigma}^\dagger c_{j\sigma} + \text{h.c.}) + U \sum_i n_{i\uparrow} n_{i\downarrow},
where c_{i\sigma}^\dagger (c_{i\sigma}) creates (annihilates) an at site i with spin \sigma, and n_{i\sigma} = c_{i\sigma}^\dagger c_{i\sigma} is the number operator.
This minimal model has profound significance for understanding correlated electron materials, including Mott , high-temperature superconductors, and itinerant ferromagnets, as the ratio U/t determines whether the system behaves as a metal or an . Exactly solvable in one dimension via the method, as demonstrated by and F. Y. Wu in 1968, the model reveals a rich with charge and gaps but no Mott transition in 1D. In higher dimensions, however, it remains analytically intractable, necessitating numerical approaches like and simulations to explore its ground states and excitations. Beyond theory, the Hubbard model has been experimentally realized using ultracold atoms in optical lattices, enabling direct observation of its phases such as the superfluid-to-Mott insulator transition in bosonic variants and fermionic analogs of . Its extensions, including the t-J model (derived in the strong-coupling limit U >> t), underpin studies of and other doped Mott insulators. Over decades, the model has influenced diverse fields, from to , serving as a for testing many-body algorithms and approximation schemes.

Introduction and Overview

Definition and Basic Principles

The Hubbard model serves as an archetype for strongly correlated fermionic systems, modeling electrons that hop between lattice sites while subject to on-site repulsion. It provides a minimal framework to investigate the competition between and electron-electron interactions in , particularly for systems with partially filled narrow energy bands. The standard single-band Hubbard Hamiltonian takes the form \hat{H} = -t \sum_{\langle i,j \rangle, \sigma} \left( c_{i\sigma}^\dagger c_{j\sigma} + \mathrm{h.c.} \right) + U \sum_i n_{i\uparrow} n_{i\downarrow}, where t > 0 denotes the hopping parameter for electrons between nearest-neighbor sites \langle i,j \rangle, U represents the strength of the on-site repulsion, c_{i\sigma}^\dagger (c_{i\sigma}) is the creation (annihilation) operator for an electron of \sigma at site i, and n_{i\sigma} = c_{i\sigma}^\dagger c_{i\sigma} is the corresponding number operator. The hopping term captures the delocalization of electrons across the , enabling formation in the non-interacting (U = 0), while the interaction term penalizes double occupancy of a site by electrons of opposite spins. Key regimes include half-filling, where the average is one per site (n = 1), balancing kinetic and potential energies to probe insulating states driven by correlations. Away from half-filling, doping introduces deviations such as holes (n < 1) or extra electrons (n > 1), altering the electronic structure and enabling studies of doped Mott insulators. This model deliberately omits longer-range interactions and couplings to isolate the role of local on-site correlations in generating emergent phenomena.

Physical Significance and Applications

The Hubbard model plays a central role in elucidating the physics of Mott insulators, where the on-site repulsion parameter U is sufficiently large to localize electrons and suppress conduction, even at half-filling where band theory would predict . This phenomenon arises from the competition between kinetic hopping t and the repulsive interaction U, leading to an insulating state characterized by antiferromagnetic order in the undoped case. The model's prediction of such charge localization has been instrumental in understanding real materials exhibiting metal-insulator transitions driven by electron correlations rather than structural changes. Beyond insulators, the Hubbard model provides insights into , particularly in materials, where doping away from half-filling introduces charge carriers that mediate d-wave pairing. In the strong-coupling limit (U \gg t), the effective low-energy theory reduces to the t-J model, capturing the spin-charge separation and pairing mechanism observed in underdoped cuprates like La_{2-x}Sr_xCuO_4. Numerical simulations of the doped Hubbard model on square lattices have demonstrated the emergence of superconducting states with pairing symmetries matching experimental observations in these layered perovskites. The model serves as a prototypical framework for strongly correlated electron systems across diverse material classes, including transition metal oxides such as and V_2O_3, where it describes charge ordering and magnetic properties. In organic conductors like the κ-(BEDT-TTF)_2X salts, the Hubbard model accounts for the interplay of bandwidth narrowing and interactions leading to insulating or metallic phases under pressure. For heavy fermion compounds, such as CeCu_6 or UPt_3, extensions of the model incorporate hybridization effects to explain the large effective masses and Kondo lattice behavior, highlighting its versatility in capturing correlation-driven enhancements of masses. Despite its successes, the Hubbard model remains an idealized single-orbital description that captures essential physics but necessitates extensions for realistic applications, such as multi-orbital terms for band structures or inclusion of disorder to model impurities in actual solids. These limitations underscore its role as a starting point for more comprehensive theories rather than a complete quantitative tool for all material properties.

Historical Development

Origins in the 1950s–1960s

In the late , conventional band theory faced significant limitations when applied to semiconductors and compounds featuring narrow d- or f-electron bands, where the of electrons is small compared to the repulsion between them, resulting in strong correlation effects that single-particle approximations could not adequately describe. This situation motivated the development of simplified models to capture the dominant role of electron-electron interactions in determining electronic properties such as and in these materials. Pioneering theoretical work laid the groundwork for addressing these correlations. In 1949, Nevill F. Mott argued that strong electron-electron repulsion in transition metals could drive a metal-insulator transition, with electrons localizing on atomic sites to minimize interaction energy, thereby challenging the delocalized picture of band theory. Complementing this, Philip W. Anderson's 1958 analysis showed that even weak disorder in a potential leads to exponentially localized wavefunctions, preventing and contributing to insulating behavior in disordered systems. These insights highlighted the need for models incorporating both interaction and disorder-like effects arising from repulsion. The Hubbard model was proposed independently in 1963 by John Hubbard, Martin Gutzwiller, and Junjiro Kanamori as a minimal single-band model to approximate the complex multi-orbital interactions in narrow d-bands of transition metals. By reducing the problem to hopping between sites with an on-site repulsion term, the model captured essential correlation physics while simplifying calculations for real materials like . In an initial approximation, Hubbard drew an analogy to a disordered , treating the repulsion parameter as inducing a random potential distribution similar to impurities, which allowed estimation of correlation-induced band splitting and localization tendencies.

Evolution and Key Milestones

In the 1960s and 1970s, significant progress was made in developing variational methods to handle electron correlations in the Hubbard model. Martin Gutzwiller introduced a variational wavefunction in 1963 that incorporates correlation effects by projecting out double occupancies on lattice sites, providing a framework to study the impact of strong interactions on metallic properties. This approach was later refined in the Gutzwiller approximation, notably by Brinkman and Rice in 1970, who applied it to the Hubbard model to predict a metal-insulator transition at half-filling for finite on-site repulsion U, where the weight vanishes and double occupancy is suppressed. These developments highlighted the model's challenges in capturing Mott localization without exact solvability, establishing variational projections as a for approximate treatments. The 1980s saw the Hubbard model gain prominence in the context of following the discovery of cuprates. In , proposed that doping a Mott insulating state described by the half-filled Hubbard model could lead to pairing mechanisms in copper oxides, linking the undoped antiferromagnetic phase to resonating valence bond states and suggesting the doped t-J model as an effective low-energy description. This perspective elevated the model from a theoretical toy to a paradigm for understanding unconventional , spurring extensive studies on doped variants despite ongoing debates about its quantitative accuracy for real materials. During the 1990s and 2000s, (DMFT) emerged as a breakthrough for tackling strong correlations non-perturbatively. Developed by Antoine Georges and Gabriel Kotliar in 1992, DMFT maps the lattice problem onto an exactly solvable quantum impurity model in the limit of infinite spatial dimensions, capturing local dynamical correlations while treating non-local effects at a mean-field level. This method resolved key features like the Mott transition and spectral properties in the Hubbard model, enabling realistic simulations of correlated materials and bridging microscopic models to finite-dimensional systems through extensions like cluster DMFT. In recent years up to 2025, the Hubbard model has been integrated with techniques to derive material-specific parameters, enhancing its predictive power for real compounds. Methods such as linear response theory within have allowed self-consistent determination of on-site U and inter-site V parameters for oxides, improving descriptions of electronic structure in materials like nickelates and manganites. Concurrently, quantum simulations using ultracold fermionic atoms in optical lattices have validated theoretical predictions, achieving cryogenic temperatures to probe Mott insulators and doping effects with unprecedented control over parameters. These advancements underscore the model's enduring role in , facilitating connections between theory, computation, and experiment.

Model Formulation

Hamiltonian and Parameters

The Hubbard model is defined by a single-band that captures the essential competition between and on-site repulsion in strongly correlated systems. The standard form for a with N sites is given by \hat{H} = -t \sum_{\langle i,j \rangle, \sigma} \left( c_{i\sigma}^\dagger c_{j\sigma} + \mathrm{h.c.} \right) + U \sum_i n_{i\uparrow} n_{i\downarrow} - \mu \sum_{i,\sigma} n_{i\sigma}, where c_{i\sigma}^\dagger (c_{i\sigma}) creates (annihilates) an electron with spin \sigma = \uparrow, \downarrow at site i, n_{i\sigma} = c_{i\sigma}^\dagger c_{i\sigma} is the number operator, \langle i,j \rangle denotes nearest-neighbor pairs, and h.c. indicates the Hermitian conjugate. This formulation arises from a tight-binding approximation for the non-interacting band structure combined with a local interaction term, motivated by the need to model narrow-band electron correlations without long-range potentials. The first term describes the of electrons hopping between nearest-neighbor sites with t > 0, which sets the scale for electron delocalization and determines the non-interacting ; for a two-dimensional , this is W = 8t. The second term introduces the on-site Hubbard repulsion U > 0, representing the energy cost for double occupancy of a site by two s of opposite spins, which promotes Mott insulating behavior at half-filling when U is large. In high-temperature superconductors like cuprates, typical values are U \approx 2--$8 eV and t \approx 0.3--$0.4 eV, reflecting screened interactions in oxides. The third term, involving the \mu, controls the average electron density n = \sum_{i,\sigma} \langle n_{i\sigma} \rangle / N, enabling studies away from half-filling (n=1) to model doping effects. The model is typically defined on bipartite lattices such as the one-dimensional chain, two-dimensional square lattice, or the infinite-coordination-number Bethe lattice, which facilitates analytical and numerical treatments in different dimensions. Simulations often employ periodic boundary conditions to mimic infinite systems and respect translational invariance, though open boundaries are used in finite-cluster methods like exact diagonalization to avoid edge effects. Parameters are conventionally expressed in units where the lattice spacing a = 1 and \hbar = 1, with energies scaled by t (often setting t=1); the dimensionless ratio U/t quantifies the relative strength of interactions, with weak coupling at U/t \ll 1 yielding perturbative Fermi liquid behavior and strong coupling at U/t \gg 1 leading to localized states.

Single-Site and Lattice Variants

The single-site limit of the Hubbard model, also known as the atomic limit, occurs when the hopping parameter t = 0, isolating each lattice site and eliminating contributions. In this case, the for a single site simplifies to H = U n_{\uparrow} n_{\downarrow} - \mu (n_{\uparrow} + n_{\downarrow}), where U is the on-site repulsion, \mu is the , and n_{\sigma} (\sigma = \uparrow, \downarrow) are the number operators for spin-up and spin-down s. At half-filling, corresponding to one per site (\mu = U/2), the consists of singly occupied sites with degenerate configurations, while excitations to doubly occupied (or empty) sites require energy U, opening a Mott insulating gap of size U between the lower and upper Hubbard bands. This limit captures the essence of strong correlations leading to Mott localization without itinerancy, serving as a foundational for understanding insulating behavior in correlated systems. Lattice variants of the Hubbard model extend the standard form by incorporating additional terms to better approximate real materials, while retaining the core on-site U and nearest-neighbor hopping t. One common modification is the inclusion of next-nearest-neighbor hopping t', which introduces frustration in antiferromagnetic ordering and modifies the non-interacting , potentially suppressing nesting instabilities at half-filling. For instance, in two-dimensional square modeling hole-doped high-temperature superconductors like cuprates, typical t'/t \approx -0.15 can stabilize incommensurate magnetic phases or enhance doping effects. Another variant adds a Zeeman term -h \sum_i (n_{i\uparrow} - n_{i\downarrow}), which couples to the total magnetization and breaks time-reversal , enabling studies of polarization, , or field-induced transitions in disordered or clean . These adaptations maintain the model's simplicity while addressing limitations in describing band structure asymmetries or external perturbations observed in oxides. Cluster extensions, such as the two-site Hubbard dimer, provide tractable finite-size models for benchmarking approximations and understanding short-range correlations beyond the single-site case. The dimer , H = -t (c_{1\sigma}^\dagger c_{2\sigma} + \text{h.c.}) + U \sum_{i=1,2} n_{i\uparrow} n_{i\downarrow} - \mu \sum_{i=1,2,\sigma} n_{i\sigma}, admits exact diagonalization in a two-electron (at half-filling), yielding eigenvalues that reveal a crossover from metallic to insulating behavior as U/t increases, with the ground-state energy E_0 = -2\mu + \frac{1}{2} \left( U - \sqrt{U^2 + 16 t^2} \right) for the symmetric case. These exact solutions highlight double occupancy suppression and spin correlations, serving as testbeds for methods like or variational approaches in larger systems. In contrast to these variants, the standard Hubbard model neglects inter-site Coulomb repulsion terms of the form V \sum_{\langle i,j \rangle} n_i n_j, which are central to the extended Hubbard model and account for longer-range charge interactions leading to phenomena like charge-density waves. The omission of V in the standard form simplifies analysis of purely local correlations but limits applicability to systems where nearest-neighbor repulsions (V \sim 0.1 U) play a role in phase competition.

Analytical Approaches in Simple Systems

One-Dimensional Exactly Solvable Cases

The one-dimensional Hubbard model, defined on a chain lattice with nearest-neighbor hopping amplitude t and on-site repulsion U, admits an exact solution via the for arbitrary U/t > 0 and electron filling fraction n. This seminal result, obtained by and F. Y. Wu in 1968, maps the onto a set of nonlinear integral equations for pseudomomenta, enabling the determination of all eigenstates, including the , without approximations. The wavefunction is constructed as a of plane waves modulated by spin-dependent phase factors, accounting for the interaction through auxiliary parameters known as charge and spin rapidities. In the (N \to \infty, where N is the number of sites), these equations simplify to coupled Fredholm integral equations for the densities of the pseudomomenta distributions. The exact solution reveals a rich characterized by quantum correlations without long-range . At half-filling (n = 1), the system exhibits a charge gap \Delta_c \propto \exp(-2\pi t / U) for any U > 0, rendering it a , while the spin sector remains gapless with antiferromagnetic correlations decaying as power laws. Away from half-filling ($0 < n < 1), a charge gap is absent, and the ground state is a gapless Luttinger liquid with separate charge and spin velocities v_c and v_s, leading to power-law decay in density-density and spin-spin correlation functions: \langle n(x) n(0) \rangle - \langle n \rangle^2 \sim \cos(2k_F x)/x^{K_c + 1/K_c} for charge and \sim (-1)^x / x for staggered spin, where K_c < 1 is the Luttinger parameter controlled by U/t. Crucially, no Mott metal-insulator transition occurs as a function of U, as the insulating phase at half-filling emerges continuously from the non-interacting metallic state without a critical point. A concrete physical realization of the one-dimensional arises in a linear chain of hydrogen atoms, where the tight-binding approximation captures the low-energy electron dynamics. In the non-interacting limit (U = 0), the single-particle wavefunctions are Bloch states \psi_k(\mathbf{r}) = \frac{1}{\sqrt{N}} \sum_j e^{i k j a} \phi_{1s}(\mathbf{r} - \mathbf{R}_j), with \phi_{1s} the hydrogen 1s orbital centered at site \mathbf{R}_j and lattice constant a \approx 1.4 Å for equilibrium spacing. The corresponding energy band is E(k) = \epsilon_0 - 2t \cos(ka), where \epsilon_0 is the on-site energy and t \approx 1-2 eV is the hopping derived from orbital overlaps, yielding a half-filled band that is metallic under band theory. Including the on-site repulsion U \approx 20-30 eV from hydrogen atom Coulomb integrals maps the system directly to the , allowing experimental probes of correlation effects like the charge gap via spectroscopy. The ground-state energy from the Bethe ansatz is computed by summing contributions from the charge pseudomomenta k_l, yielding E = \sum_{l=1}^{M} (-2t \cos k_l) in the finite system, where M = nN is the total number of electrons (assuming even N). In the thermodynamic limit at arbitrary filling, this becomes the integral e = E/N = -\frac{2t}{\pi} \int_{-\pi/2}^{\pi/2} \cos k \, \rho(k) \, dk, where the density \rho(k) satisfies the integral equation \rho(k) + \int_{-\pi/2}^{\pi/2} dk' A(k - k') \rho(k') = \frac{1}{2\pi} + \frac{1}{2\pi} \int_{-\Lambda}^{\Lambda} dl \, B(k, l) \sigma(l). Here, A and B are Bethe ansatz kernels involving \tan^{-1} functions with argument scaled by U/(4t), \sigma(l) is the spin rapidity density solving a subordinate equation up to cutoff \Lambda, and parameters are fixed by filling n = \int \rho(k) dk. At half-filling, the expressions simplify due to particle-hole symmetry, with numerical solution showing e(0) = -4t/\pi \approx -1.273 t and e(\infty) = 0.

Infinite-Dimensional Limit

In the limit of infinite spatial dimensions d \to \infty, the nearest-neighbor hopping amplitude t in the is rescaled as t^* = t / \sqrt{2d} to maintain a finite kinetic energy scale and variance in the non-interacting density of states. This scaling suppresses long-range spatial correlations, causing the self-energy to become purely local, with non-local components vanishing as $1/d. As a result, the physics is dominated by local interactions at each site, simplifying the treatment of strong correlations. The Bethe lattice with infinite coordination number serves as the canonical geometry for this limit, yielding a semicircular non-interacting density of states: \rho(\epsilon) = \frac{1}{2\pi (t^*)^2} \sqrt{4(t^*)^2 - \epsilon^2}, where the bandwidth is W = 4t^* (or half-bandwidth D = 2t^*). This density of states facilitates analytical tractability and captures the essential features of the infinite-dimensional dynamics. The infinite-dimensional maps exactly onto an effective single-site action describing a quantum impurity embedded in a self-consistent bath. This mapping arises from diagrammatic perturbation theory, where vertex corrections and non-local diagrams become negligible, reducing the lattice problem to a local self-energy approximation. Alternatively, the cavity method derives the effective action by excluding one site from the lattice and treating the influence of the remaining "cavity" as a dynamic mean field acting on the impurity site. At half filling, the model undergoes a Mott metal-insulator transition at a finite critical on-site repulsion U_c \approx 3W, where W denotes the half-. The transition features a region of coexistence between metastable metallic and insulating phases, spanning approximately U_{c1} \approx 2.4W (closure of the metallic solution) and U_{c2} \approx 3W (onset of the insulating solution).

Theoretical Approximations

Mean-Field and Hartree-Fock Treatments

The Hartree-Fock approximation provides a foundational mean-field treatment of the Hubbard model by decoupling the on-site interaction term n_{i\uparrow} n_{i\downarrow} into a form that linearizes the Hamiltonian, specifically n_{i\uparrow} n_{i\downarrow} \approx \langle n_{i\uparrow} \rangle n_{i\downarrow} + n_{i\uparrow} \langle n_{i\downarrow} \rangle - \langle n_{i\uparrow} \rangle \langle n_{i\downarrow} \rangle, where the expectation values are self-consistently determined. This decoupling generates an effective single-particle Hamiltonian with renormalized energy bands shifted by the mean-field potentials, leading to a quasiparticle description that captures weak-coupling instabilities. In particular, for ferromagnetism, the approximation yields the Stoner criterion U \rho(\epsilon_F) > 1, where \rho(\epsilon_F) is the noninteracting at the , signaling the onset of when the interaction strength U exceeds a set by the bandwidth. A variant known as restricted Hartree-Fock focuses on antiferromagnetic order, particularly relevant at half-filling on bipartite lattices, where the approximation assumes a staggered pattern that opens a gap in the for any U > 0. This restricts the wave function to maintain total spin and charge uniformity while allowing spatial modulation of spin densities, resulting in an insulating antiferromagnetic with a nesting-driven at the . The gap size scales with U in the weak-coupling limit, providing a simple picture of magnetic ordering that aligns with nesting properties of the noninteracting band structure. To address stronger correlations, slave-boson mean-field theories introduce auxiliary bosonic fields to represent empty and doubly occupied sites, effectively projecting out double occupancy in the infinite-U limit while allowing finite U treatments through saddle-point approximations. Pioneered by Barnes, this approach decomposes the operators into fermionic and bosonic auxiliaries, yielding a renormalized Fermi with suppression as U increases. In the half-filled case, the theory predicts a Brinkman-Rice metal-insulator transition at a critical U_c = 8t (for a bandwidth W = 8t), where the quasiparticle weight vanishes and the system becomes incompressible, marking the collapse of the metallic state into a Mott-like with finite double occupancy approaching zero. Despite these insights, mean-field and Hartree-Fock treatments exhibit significant limitations in capturing strong-correlation physics. They tend to overestimate the propensity for magnetic ordering, predicting instabilities like at arbitrarily small U in the half-filled case, which contradicts the absence of long-range order at weak coupling in low dimensions. Moreover, without explicit projections to enforce no double occupancy, these approximations fail to produce a proper in the paramagnetic sector, retaining metallic behavior even at large U and neglecting short-range correlations that suppress charge fluctuations.

Strong-Coupling and t-J Model Derivations

In the strong-coupling limit of the Hubbard model, where the on-site repulsion U greatly exceeds the hopping amplitude t (i.e., U \gg t), second-order provides an effective description by treating the hopping term as a perturbation on the dominant U term. At half-filling, the ground state is a with one electron per site, and virtual hopping processes between neighboring sites generate an antiferromagnetic interaction. This leads to an effective Heisenberg Hamiltonian H = J \sum_{\langle i,j \rangle} \mathbf{S}_i \cdot \mathbf{S}_j, where the exchange constant is J = 4t^2 / U, derived from processes in which an electron hops to a neighboring site, creating a doubly occupied site (costing energy U), and then hops back, effectively exchanging spins on the sites. For doped systems away from half-filling, the strong-coupling expansion must account for mobile holes while prohibiting double occupancy to avoid high-energy states. The resulting effective model is the t-J model, obtained by projecting the onto states without double occupancies and deriving the low-energy dynamics via second-order in t/U. The hopping term allows holes to move without creating double occupancies, while the spin-exchange term arises similarly to the Heisenberg case. The t-J takes the form H_{t-J} = -t \sum_{\langle i,j \rangle, \sigma} (1 - n_{i,-\sigma}) c^\dagger_{i\sigma} c_{j\sigma} + \text{H.c.} + J \sum_{\langle i,j \rangle} \left( \mathbf{S}_i \cdot \mathbf{S}_j - \frac{1}{4} n_i n_j \right), where the factor (1 - n_{i,-\sigma}) enforces the no-double-occupancy constraint, \mathbf{S}_i are operators, n_i = n_{i\uparrow} + n_{i\downarrow}, and J = 4t^2 / U. This model captures the physics of lightly doped Mott insulators, such as the . At half-filling (\delta = 0), the t-J model reduces to the Heisenberg antiferromagnet, exhibiting long-range Néel antiferromagnetic order in two dimensions due to the J > 0. Upon doping with holes ($0 < \delta \ll 1), the undoped antiferromagnetic background is disrupted, but resonating valence bond (RVB) mean-field theories applied to the t-J model predict a tendency toward d-wave superconductivity, where holes pair in a spin-singlet state mediated by spin fluctuations. This RVB picture posits that the ground state involves a superposition of singlet bonds, with doping introducing mobile holons that condense to form superconductivity. The t-J approximation is valid for U/t \gtrsim 8--$10, as estimated for cuprate materials where t \approx 0.4 eV and U \approx 4--$8 eV from cluster calculations and comparison to experimental exchange energies J \approx 0.1--$0.15 eV. Below this ratio, higher-order terms in the perturbation expansion become significant, and the full must be used to avoid inaccuracies in the charge and spin sectors.

Numerical and Computational Methods

Exact Diagonalization and Small-System Studies

Exact diagonalization (ED) provides an exact numerical solution to the Hubbard model by constructing and fully diagonalizing the Hamiltonian matrix in the complete Fock basis for finite clusters of lattice sites. This basis spans all possible electron occupancy configurations per site—empty, singly occupied with spin up or down, or doubly occupied—yielding a Hilbert space dimension of $4^N for N sites. Due to the exponential scaling of this dimension with system size, ED is practically limited to small clusters, typically up to N \approx 16–20 sites, depending on lattice geometry and computational resources; for example, 4×4 square clusters in two dimensions are commonly studied after exploiting symmetries like particle number and total spin conservation to reduce the effective matrix size. The method delivers all eigenvalues and eigenstates, enabling precise calculations of static properties such as ground-state energies, charge gaps, and correlation functions directly from the exact spectrum. In small-system studies, ED reveals cluster-size dependencies in key features like the Mott gap at half-filling; for instance, on 2×2 plaquette clusters, a charge gap that becomes prominent for interaction strengths U \gtrsim 4t marks insulating behavior, while larger clusters like 4×4 show a more gradual transition influenced by finite-size effects and antiferromagnetic correlations. These insights highlight how short-range correlations dominate in tiny systems, providing benchmarks for approximate theories. For ground-state properties and low-energy excitations, the Lanczos algorithm is frequently integrated with ED, iteratively building a tridiagonal representation of the starting from a trial vector to converge on extremal eigenvalues and eigenvectors without requiring full matrix diagonalization. This approach efficiently computes expectation values like magnetization and kinetic energy. To access dynamical responses, such as spectral functions and dynamical structure factors, Lanczos-generated continued fraction expansions of retarded are used; representative applications on 8–10 site clusters in the two-dimensional uncover hole motion as coherent quasiparticles at low doping, evolving into string-like confined states for stronger interactions, with characteristic dispersion relations tied to the underlying antiferromagnetic background. Despite its exactness, ED's primary limitation remains the exponential growth of computational cost, restricting analyses to systems far from the thermodynamic limit and necessitating careful extrapolation of finite-size results to infer bulk behavior. Early seminal works, such as those on pairing susceptibilities and single-hole dynamics in two-dimensional clusters, established ED as a foundational tool for probing strong-correlation physics in the .

Quantum Monte Carlo and Dynamical Mean-Field Theory

Quantum Monte Carlo (QMC) methods provide a powerful stochastic approach for studying the finite-temperature properties of the , particularly through determinant QMC (DQMC), which was introduced for simulating interacting fermion systems including the . In this technique, the interaction term U \sum_i n_{i\uparrow} n_{i\downarrow} is decoupled using a into fluctuating auxiliary fields, allowing the problem to be mapped onto non-interacting fermions whose partition function is evaluated as a determinant of the single-particle Green's function matrix. The world-line algorithm then samples these configurations in imaginary time, enabling computations of thermodynamic quantities like equal-time correlations and susceptibilities for lattice sizes up to $16 \times 16 in two dimensions at moderate temperatures. Recent improvements, such as submatrix update algorithms, have extended capabilities to systems with up to 8,000 sites in three dimensions as of 2024. A key challenge in DQMC for the Hubbard model is the fermion sign problem, which arises from negative weights in the Monte Carlo sampling, leading to exponential computational cost with decreasing temperature or increasing system size. At half-filling on bipartite lattices, the sign problem is absent due to particle-hole symmetry, which ensures all determinants are positive, allowing reliable simulations down to low temperatures. However, away from half-filling, such as in doped systems, the sign problem becomes severe, severely limiting studies of phenomena like superconductivity in the underdoped regime. Dynamical mean-field theory (DMFT) addresses the limitations of finite-dimensional approximations by leveraging the exact solvability of the in the infinite-dimensional limit, where the self-energy becomes purely local, to map the lattice problem onto a self-consistent single-site quantum impurity model embedded in a self-consistent bath. The local Green's function is obtained via G(\omega) = \int d\epsilon \, \rho(\epsilon) \frac{1}{\omega + \mu - \epsilon - \Sigma(\omega)}, where \rho(\epsilon) is the non-interacting density of states, \Sigma(\omega) is the self-energy from solving the impurity model, and the bath hybridization function is iteratively adjusted for self-consistency. Impurity solvers such as (for finite temperatures) or (for zero temperature and dynamics) are employed to compute \Sigma(\omega), enabling access to spectral functions and transport properties in dimensions d \geq 2. DMFT reveals rich physics in the , including a first-order at low temperatures with hysteresis between coexisting metallic and insulating phases, as the interaction strength U is varied across a critical value. In the metallic phase, DMFT captures a crossover from a coherent at low U or high temperatures to an incoherent bad metal regime near the transition, characterized by a pseudogap in the spectral function and enhanced scattering rates that violate standard quasiparticle descriptions. These features highlight DMFT's success in describing strong-correlation effects beyond simple mean-field theories.

Extensions to Complex Systems

Multi-Orbital and Doped Hubbard Models

The multi-orbital extends the single-band version to systems with multiple degenerate or partially degenerate orbitals per site, capturing the essential physics of transition metal compounds where d-electrons occupy several orbitals. The Hamiltonian is given by H = -\sum_{\langle i,j \rangle, m,m',\sigma} t_{mm'}(c_{i m \sigma}^\dagger c_{j m' \sigma} + \text{h.c.}) + \sum_{i,m,m',\sigma} \epsilon_{mm'} c_{i m \sigma}^\dagger c_{i m' \sigma} + U \sum_{i,m} n_{i m \uparrow} n_{i m \downarrow} + \sum_{i,m < m',\sigma \neq \sigma'} U' n_{i m \sigma} n_{i m' \sigma'} + \sum_{i,m < m',\sigma} (U'-J) n_{i m \sigma} n_{i m' \sigma} + J \sum_{i,m \neq m'} (c_{i m \uparrow}^\dagger c_{i m \downarrow}^\dagger c_{i m' \downarrow} c_{i m' \uparrow} + \text{h.c.}), where c_{i m \sigma}^\dagger (c_{i m \sigma}) creates (annihilates) an electron with spin \sigma in orbital m at site i, n_{i m \sigma} = c_{i m \sigma}^\dagger c_{i m \sigma}, t_{mm'} is the inter-site orbital-dependent hopping amplitude between nearest neighbors \langle i,j \rangle, \epsilon_{mm'} is the intra-site orbital energy (crystal field), U is the intra-orbital Coulomb repulsion, U' is the inter-orbital repulsion for opposite spins, and J is Hund's exchange coupling. This form ensures rotational invariance through the Kanamori parametrization, with relations like U' = U - 2J, including pair-hopping and spin-flip terms. To ensure rotational invariance in the interaction sector, the Kanamori parametrization is commonly employed, expressing the Coulomb interactions in terms of Slater integrals F^0, F^2, and F^4 reduced to parameters U, U', and J. This approach, originally developed for transition metal ions, parameterizes the multi-orbital interactions such that U' = U - 2J for opposite spins and U' - J for parallel spins, while including explicit pair-hopping (J (c_{i m \uparrow}^\dagger c_{i m \downarrow}^\dagger c_{i m' \downarrow} c_{i m' \uparrow} + \text{h.c.})) and spin-flip terms. The parametrization simplifies computations in and cluster extensions, enabling studies of orbital-dependent correlations in realistic materials. Recent developments include nonequilibrium multi-orbital extensions using two-particle self-consistent theory and numerically exact methods like for simulating steady-state properties, as of 2024. Doping away from half-filling introduces charge carriers into the multi-orbital Hubbard model, defined by the doping level \delta = 1 - n where n is the average electron density per site. In cuprate systems, doping (\delta > 0) leads to the formation of Zhang-Rice singlets, where a doped on oxygen orbitals hybridizes with the d-orbital to create an effective spin-singlet state that behaves as a single mobile carrier. This mechanism reduces the effective model to a single-band t-J-like description at low doping, enhancing antiferromagnetic correlations and , though multi-orbital effects persist in modulating the singlet . Phase diagrams of multi-orbital Hubbard models reveal rich structures, including orbital-selective Mott transitions (OSMT), where individual orbitals undergo metal-insulator transitions at different critical interactions due to bandwidth differences or crystal-field splittings. In two-orbital models with unequal bandwidths, shows that the narrower band localizes first, forming an orbital-selective while the wider band remains metallic, stabilized by Hund's J which suppresses double occupancy across orbitals. These transitions occur at finite temperatures and can be tuned by doping, leading to bad-metal phases with coexisting itinerant and localized electrons, crucial for understanding iron-based superconductors.

Real-Material Applications

The Hubbard model has been extensively applied to high-temperature superconducting cuprates, such as La_{2-x}Sr_xCuO_4 (LSCO), where typical parameters include an on-site repulsion U \approx 4-6 and nearest-neighbor hopping t \approx 0.4 . These values place the system in the intermediate to strong coupling regime, capturing the Mott insulating behavior of the undoped parent compound and the evolution to a doped metal upon introducing holes. (DMFT) treatments of the doped Hubbard model predict a pseudogap phase emerging from strong antiferromagnetic correlations, consistent with the suppression of low-energy weight observed in underdoped cuprates. In the same doping range, the t-J model—derived from the strong-coupling limit of the Hubbard —exhibits stripe order, characterized by alternating charge and spin density waves, which aligns with the nanoscale inferred from scanning tunneling in LSCO. In iron-based superconductors, or pnictides, multi-orbital extensions of the Hubbard model incorporate five d-orbitals per atom to describe the complex band structure and orbital . These models reveal that orbital fluctuations, enhanced by moderate electron-phonon coupling, mediate s-wave , but strong antiferromagnetic fluctuations favor an s^{\pm} with sign reversal between and pockets, matching the observed in compounds like LaFeAsO_{1-x}F_x. The multi-orbital Hubbard framework thus provides a unified description of the spin-density-wave instability in undoped pnictides and the emergent upon doping. Experimental validations of Hubbard-model predictions in these materials include optical conductivity measurements \sigma(\omega), which reveal prominent Hubbard bands as split lower and upper manifolds separated by \sim U, with the lower Hubbard band dominating the Drude-like response in metallic phases. In cuprates, mid-infrared features in \sigma(\omega) for underdoped LSCO confirm the transfer of spectral weight to high energies due to correlations, as anticipated by the model. Neutron scattering experiments further support the framework by detecting dispersive magnon excitations in the antiferromagnetic state of undoped La_2CuO_4, with a spin-wave velocity matching Hubbard-model estimates from linear spin-wave theory. In pnictides, inelastic neutron scattering reveals high-energy spin excitations persisting into the superconducting state, consistent with itinerant antiferromagnetic correlations in the multi-orbital Hubbard picture. Applying the Hubbard model to real materials faces challenges in parameter extraction, particularly via augmented with Hubbard correction (DFT+U), where determining U requires careful linear-response calculations to achieve convergence and avoid artifacts from approximations. Additionally, the model inherently neglects vibrations, omitting electron-phonon coupling that influences charge ordering in cuprates; for instance, the pure Hubbard model predicts stripes preceding charge stripes upon doping, contrary to observations where phonons enhance charge modulations first.

Quantum Simulation and Experimental Realizations

Analog Quantum Simulators

Analog quantum simulators for the Hubbard model primarily utilize ultracold fermionic atoms confined in optical lattices, which directly emulate the through the tight-binding . These platforms leverage the ability to map atomic motion and interactions onto the lattice sites, with ^6Li atoms serving as a prototypical choice due to their fermionic nature and broad Feshbach resonance for interaction tuning. The hopping amplitude t is adjusted by varying the optical lattice depth V_0, where deeper lattices suppress tunneling and reduce t exponentially as t \propto e^{-\sqrt{V_0}}; meanwhile, the on-site repulsion U is controlled via magnetic-field-tuned Feshbach resonances that modify the s-wave scattering length a_s, enabling U/t ratios up to about 10 in typical setups. A seminal experimental milestone was achieved by the Esslinger group in the late 2000s, culminating in the 2008 observation of the phase at unit filling (n = 1), where increasing U/t transitioned the system from a compressible metallic state to an incompressible , as evidenced by vanishing . Subsequent measurements of in the Mott regime, reported in 2015, quantified the low to density changes (\kappa \approx 0) across the , providing direct thermodynamic confirmation of the gapped . Key probing techniques include time-of-flight (TOF) imaging, where the is suddenly turned off to allow ballistic , yielding the momentum distribution n(\mathbf{k}) that exhibits a sharp drop in coherence peaks for the . Additionally, noise correlations—measured as density fluctuations in the expanded cloud—reveal short-range antiferromagnetic order and allow extraction of the double occupancy \langle n_{\uparrow} n_{\downarrow} \rangle, which suppresses to near zero in the strong-coupling limit as doubly occupied sites become energetically unfavorable. These analog simulators excel at capturing real-time , such as the light-cone spreading of correlations following quenches, which remain inaccessible to classical numerical methods due to exponential scaling. However, challenges persist in achieving homogeneity across and , where harmonic trapping potentials introduce density gradients and filling inhomogeneities, limiting system sizes to a few tens of lattice sites per dimension.

Digital Quantum Computing Approaches

Digital quantum computing approaches to the Hubbard model leverage universal gate-based quantum processors to simulate the model's fermionic , offering programmable flexibility for exploring its dynamics and ground states beyond the limitations of analog simulators. These methods typically map the fermionic operators in the Hubbard , H = -t \sum_{\langle i,j \rangle, \sigma} (c_{i\sigma}^\dagger c_{j\sigma} + \text{h.c.}) + U \sum_i n_{i\uparrow} n_{i\downarrow}, to qubits using transformations such as the Jordan-Wigner mapping, which encodes fermions on a chain while preserving locality for one-dimensional systems. Time evolution is approximated via Trotter-Suzuki decomposition into sequences of single- rotations and two- entangling gates, enabling the study of strongly correlated phenomena like Mott insulation and . Variational quantum eigensolvers (VQE) are often employed to approximate ground states by optimizing parameterized circuits against the Hamiltonian's expectation value. Early demonstrations established the feasibility of digital fermionic simulations on superconducting quantum circuits, focusing on simplified models that foreshadow Hubbard applications. In 2015, researchers used superconducting processors to simulate transverse-field Ising models mappable to free fermions, validating digital techniques for interacting systems. This work laid the groundwork for Hubbard simulations by demonstrating efficient encoding of hopping and interaction terms via Gaussian fermion-to-qubit mappings. Subsequent efforts extended to direct Hubbard implementations on noisy intermediate-scale quantum (NISQ) devices, prioritizing low-depth circuits to combat decoherence. NISQ-era protocols have enabled real-time dynamics simulations of the one-dimensional Fermi-Hubbard model on superconducting hardware, showcasing quantum utility over classical methods for entangled regimes. A 2025 proposal outlined a Z₂ lattice-gauge theory mapping for the 1D model, implementable on NISQ processors with constant circuit depth per Trotter step, using nearest- and next-nearest-neighbor gates tailored to heavy-hexagonal topologies like IBM's. Experimental validation on IBM's processors (up to 156 s) simulated systems of 20 to 104 s (10 to 52 sites), capturing staggered dynamics for evolution times up to τ=4 with first- and second-order Trotterization (depths of 23 and 46 gates per step, respectively). Error mitigation techniques, including zero-noise extrapolation and probabilistic error cancellation, yielded results aligning with simulations, demonstrating advantage in high-entanglement scenarios where classical scale poorly. Recent advances have scaled simulations to two-dimensional lattices, accessing Hubbard physics at sizes intractable classically. In , a programmable on Google's 105-qubit superconducting processor realized 2D Fermi-Hubbard dynamics on up to 6×6 lattices, probing parameters like interaction strength U/t and to observe magnetic polarons, orders, bond solids, and thermalization. Circuits employed optimized Trotterization with local Jordan-Wigner encoding, achieving evolution fidelities validated against exact diagonalization for small systems and for larger ones, highlighting platforms' versatility for parameter sweeps. For fault-tolerant prospects, resource-optimized algorithms reduce qubit and gate overheads for the Fermi-Hubbard model, targeting high-temperature superconductor insights; one analysis compiled circuits with O(1) qubits per site using low-depth QEC codes, estimating simulations of 16-site systems within 10^6 physical gates on future logical processors. These developments underscore quantum computing's potential for precise, scalable Hubbard studies, though current limitations in and gate restrict system sizes to dozens of sites.

References

  1. [1]
    The Hubbard model at half a century | Nature Physics
    their original papers all appearing in 1963. The main motivation was the need for a way to tackle the behaviour of ...
  2. [2]
    Electron correlations in narrow energy bands - Journals
    To investigate this situation a simple, approximate model for the interaction of electrons in narrow energy bands is introduced.
  3. [3]
    The Hubbard Model: Basic Ideas and Some Results - PMC - NIH
    Jul 18, 2023 · The aim of this paper is to present a mini-review of the HM for the nonspecialized readers. The HM has become a complex research subject.Missing: history | Show results with:history
  4. [4]
  5. [5]
    [2104.00064] The Hubbard model: A computational perspective - arXiv
    Mar 31, 2021 · The Hubbard model is the simplest model of interacting fermions on a lattice and is of similar importance to correlated electron physics as the Ising model.Missing: definition | Show results with:definition
  6. [6]
    [PDF] 4 An Introduction to the Hubbard Hamiltonian - Richard T. Scalettar ...
    The Hubbard Hamiltonian (HH) offers one of the most simple ways to get insight into how the interactions between electrons give rise to insulating, ...
  7. [7]
    [PDF] The Hubbard Model: A Computational Perspective - NSF PAR
    Aug 10, 2023 · Abstract. The Hubbard model is the simplest model of interacting fermions on a lattice and is of similar importance to correlated electron ...
  8. [8]
    Mechanism of superconductivity in the Hubbard model at ... - PNAS
    We investigate the pairing glue of superconductivity in the two-dimensional Hubbard model, for interaction values relevant to cuprate physics, ...
  9. [9]
    Absence of Diffusion in Certain Random Lattices | Phys. Rev.
    This paper presents a simple model for such processes as spin diffusion or conduction in the impurity band.
  10. [10]
    Effect of Correlation on the Ferromagnetism of Transition Metals
    Effect of correlation on the ferromagnetism of transition metals. Martin C. Gutzwiller Research Laboratory. Zurich, International Business Machines Corporation.
  11. [11]
    Application of Gutzwiller's Variational Method to the Metal-Insulator ...
    Nov 15, 1970 · It is shown that the approximate variational calculation of Gutzwiller predicts a metal-insulator transition as the intra-atomic Coulomb interaction is ...
  12. [12]
    The Resonating Valence Bond State in La2CuO4 and ... - Science
    The Resonating Valence Bond State in La2CuO4 and Superconductivity. P. W. AndersonAuthors Info & Affiliations. Science. 6 Mar 1987.
  13. [13]
    Ab initio calculation of the Hubbard and Hund exchange in local ...
    Nov 2, 2022 · We show that the Z element (or element) in Heusler compounds significantly reduces the strength of the Hubbard parameter for Mn electrons ...
  14. [14]
    A neutral-atom Hubbard quantum simulator in the cryogenic regime
    Jun 11, 2025 · Here we demonstrate a several-fold reduction in temperature, bringing large-scale quantum simulations of the Hubbard model into an entirely new ...
  15. [15]
    Physics of cuprates with the two-band Hubbard model - arXiv
    Nov 3, 2004 · We conclude that a single-band tt'-U Hubbard model captures the basic physics of the cuprates concerning superconductivity, antiferromagnetism, pseudogap and ...
  16. [16]
    [PDF] Elementary Introduction to the Hubbard Model - Scalettar
    The Hubbard model offers one of the most simple ways to get insight into how the interac- tions between electrons can give rise to insulating, magnetic, and ...<|control11|><|separator|>
  17. [17]
    [PDF] Phase Diagram of the Multi-Orbital Hubbard Model
    Aug 3, 2018 · Accordingly, it takes out a single atomic site ... the Fermi energy lies within the gap, in the atomic limit we find the Mott insulating phase.
  18. [18]
    Two-dimensional Hubbard model with nearest
    Mar 1, 1987 · We study the magnetic properties of the two-dimensional Hubbard model with nearest-neighbor (t) and next-nearest-neighbor ( t 2 ) hopping and on-site repulsion ...Missing: variants | Show results with:variants<|separator|>
  19. [19]
    Magnetic phase diagram of the Hubbard model with next-nearest ...
    Aug 19, 2009 · We use the amplitude t2/t1 of the NNN hopping to tune the density of states (DOS) of the non-interacting system from a situation with particle– ...
  20. [20]
    [PDF] Frustration of antiferromagnetism in the t-t'-Hubbard model at weak ...
    The perfect-nesting instability towards antiferromagnetism of the Hubbard model is sup- pressed by next-nearest neighbor hopping t'. The asymptotic behavior of ...
  21. [21]
    Interacting Electrons in a Two-Dimensional Disordered Environment
    Jun 17, 2003 · In summary, applying a Zeeman magnetic field in the 2D disordered Hubbard model reduces the effect of the Hubbard interaction and is able to ...
  22. [22]
    [PDF] Effect of interactions, disorder and magnetic field in the Hubbard ...
    Abstract. The effects of both interactions and Zeeman magnetic field in disordered elec- tronic systems are explored in the Hubbard model on a square lattice.
  23. [23]
    [PDF] Hubbard Hamiltonian in the dimer representation. Large U limit - arXiv
    5. 2 Exact solution of the Hubbard dimer. The eigenvalues and eigenvectors of the dimer Hamiltonian (5) can be found by using a standard procedure (cf Refs ...
  24. [24]
    [PDF] The Hubbard Dimer: A density functional case study of a many-body ...
    This review explains the relationship between density functional theory and strongly correlated models using the simplest possible example, the two-site Hubbard ...
  25. [25]
    [PDF] Non-standard Hubbard models in optical lattices: a review
    The Hubbard model, named after John Hubbard, is the simplest model of interacting particles in a lattice, with only two terms in the Hamiltonian (Hubbard 1963) ...
  26. [26]
    [PDF] The Hubbard Model and its Properties
    The Hubbard model describes interacting quantum particles in a solid, using a tight-binding description with on-site interaction, and was proposed by J.  ...Missing: history origins
  27. [27]
    Mean-field approximation of the Fermi–Hubbard model expressed in ...
    Jul 12, 2023 · The mean-field approximation (MF) is one of the simplest approximations for the FH model. It consists in replacing the two-body interaction term ...INTRODUCTION · Fermi–Hubbard model · Mean-field approximation and...
  28. [28]
    [PDF] arXiv:1901.02617v2 [cond-mat.str-el] 10 Jan 2019
    Jan 10, 2019 · The Hartree-Fock approximation discussed above leads one to the so called Stoner criterion. It says that the Hubbard model exhibits ...
  29. [29]
    [PDF] Update of Hartree--Fock theory for Hubbard-like models - arXiv
    Jun 19, 2025 · We note in passing that, for the Hubbard model at half-filling, the restriction of Hartree–Fock theory to AF states can be done without loss ...
  30. [30]
    [PDF] Mean-field solution of the Hubbard model: the magnetic phase ...
    Feb 26, 2024 · Bandwidth is W = 8t. ... In the previous section we have analyzed formalism and phase diagram of the one-band Hubbard model on a square lattice in ...
  31. [31]
    Exact and many-body perturbation solutions of the Hubbard model ...
    Mar 17, 2022 · However, the MF approach misses all the correlation effects introduced by the full Hubbard model. ... Hartree–Fock (“usHF”) method employed in Ref ...
  32. [32]
  33. [33]
  34. [34]
    Exact diagonalization study of optical conductivity in the two ...
    Jul 11, 2005 · In this paper, we perform the exact diagonalization study of σ ( ω ) in the 2D Hubbard model up to a 20-site cluster. The diagonalization of the ...
  35. [35]
    [PDF] 7. Exact diagonalization
    Similarly we find for the t-J model 3N states and for the Hubbard model 4N states. This exponential growth of the matrix with lattice size makes even small.
  36. [36]
    [PDF] 7 Exact Diagonalization and Lanczos Method - Erik Koch J¨ulich ...
    Exact diagonalization and Lanczos method are discussed, with exact diagonalization including variational principles and matrix eigenvalue problem. Lanczos ...
  37. [37]
    Spectral function in the two-dimensional Hubbard model | Phys. Rev. B
    Nov 1, 1991 · The spectral function of a single hole in the two-dimensional Hubbard model is studied numerically through exact diagonalization for a nine-site ...
  38. [38]
    Pairing in the two-dimensional Hubbard model: An exact ...
    May 1, 1988 · We have studied the pair susceptibilities for all possible pair wave functions that fit on a two-dimensional (2D) eight-site Hubbard cluster ...Missing: papers | Show results with:papers
  39. [39]
    Quantum critical points and the sign problem - Science
    Jan 27, 2022 · We propose a reinterpretation of the low average sign for the Hubbard model on the square lattice away from half filling in terms of the onset ...
  40. [40]
    [PDF] Multiorbital Hubbard model in infinite dimensions: Quantum Monte ...
    Using the quantum Monte Carlo technique we compute thermodynamics and spectra for the orbitally de- generate Hubbard model in infinite spatial dimensions.Missing: seminal | Show results with:seminal
  41. [41]
    [PDF] Spectral and transport properties of doped Mott-Hubbard systems ...
    in Table I: the Hubbard model with U = 8t (full line) and for the model parameters of La2−xSrxCuO4 (dotted line), and. YBa2Cu3O6+x (long-dashed line). The ...
  42. [42]
    Intrinsic cluster-shaped density waves in cellular dynamical mean ...
    Dec 27, 2019 · In particular, cellular dynamical mean-field theory (CDMFT) [18, 19] predicts a pseudogap in the Hubbard model of cuprates without any need for ...
  43. [43]
    Stripes and spin-density waves in the doped two-dimensional ...
    Mar 28, 2022 · Stripe order is commonly observed in cuprates [26] . There is evidence suggesting that the stripe order might be the origin of pseudogap phase ...
  44. [44]
    Orbital-Fluctuation-Mediated Superconductivity in Iron Pnictides
    Apr 15, 2010 · In this Letter, we introduce the five-orbital Hubbard-Holstein (HH) model for iron pnictides, considering the e -ph interaction by Fe-ion ...
  45. [45]
    Electrodynamics of correlated electron materials | Rev. Mod. Phys.
    Jun 2, 2011 · The complex optical conductivity σ ˜ ( ω ) = σ 1 ( ω ) + i σ 2 ( ω ) of FL quasiparticles residing in a partially filled parabolic band is ...
  46. [46]
    Persistent high-energy spin excitations in iron-pnictide ... - Nature
    Feb 12, 2013 · Orbital-fluctuation-mediated superconductivity in iron pnictides: analysis of the filve-orbital Hubbard-Holstein model. ... Neutron scattering ...Results · Fe 3d Fluorescence... · Fitting Of The Spin...
  47. [47]
    Hubbard parameters from density-functional perturbation theory
    Aug 16, 2018 · We present a transparent and computationally efficient approach for the first-principles calculation of Hubbard parameters from linear-response theory.Missing: challenges | Show results with:challenges
  48. [48]
    Stripe correlations in the two-dimensional Hubbard-Holstein model
    Dec 2, 2022 · Many experiments, however, also indicate that the cuprates have strong electron-phonon (e-ph) coupling, and it is unclear how this interaction ...
  49. [49]
    Quantum simulation of the Hubbard model with ultracold fermions in ...
    Sep 3, 2018 · Here, we review recent progress of the quantum simulation of the emblematic Fermi-Hubbard model with ultracold atoms.Missing: Doppler | Show results with:Doppler
  50. [50]
    Compressibility of a Fermionic Mott Insulator of Ultracold Atoms
    Feb 20, 2015 · We characterize the Mott insulating regime of a repulsively interacting Fermi gas of ultracold atoms in a three-dimensional optical lattice.
  51. [51]
    Digital quantum simulation of fermionic models with a ... - Nature
    Jul 8, 2015 · How to cite this article: Barends, R. et al. Digital quantum simulation of fermionic models with a superconducting circuit. Nat. Commun. 6 ...
  52. [52]
    [2509.14196] Quantum Utility in Simulating the Real-time Dynamics ...
    Sep 17, 2025 · Quantum Utility in Simulating the Real-time Dynamics of the Fermi-Hubbard Model using Superconducting Quantum Computers. Authors:Talal Ahmed ...
  53. [53]
  54. [54]
    Resource-optimized fault-tolerant simulation of the Fermi-Hubbard ...
    Aug 11, 2025 · The Fermi-Hubbard (FH) model is a longstanding model of interacting electronic systems, which despite its ostensibly simple definition, contains ...Results · Discussion · MethodsMissing: digital | Show results with:digital