Dynamical mean-field theory

Last updated

Dynamical mean-field theory (DMFT) is a method to determine the electronic structure of strongly correlated materials. In such materials, the approximation of independent electrons, which is used in density functional theory and usual band structure calculations, breaks down. Dynamical mean-field theory, a non-perturbative treatment of local interactions between electrons, bridges the gap between the nearly free electron gas limit and the atomic limit of condensed-matter physics. [1]

Contents

DMFT consists in mapping a many-body lattice problem to a many-body local problem, called an impurity model. [2] While the lattice problem is in general intractable, the impurity model is usually solvable through various schemes. The mapping in itself does not constitute an approximation. The only approximation made in ordinary DMFT schemes is to assume the lattice self-energy to be a momentum-independent (local) quantity. This approximation becomes exact in the limit of lattices with an infinite coordination. [3]

One of DMFT's main successes is to describe the phase transition between a metal and a Mott insulator when the strength of electronic correlations is increased. It has been successfully applied to real materials, in combination with the local density approximation of density functional theory. [4] [5]

Relation to mean-field theory

The DMFT treatment of lattice quantum models is similar to the mean-field theory (MFT) treatment of classical models such as the Ising model. [6] In the Ising model, the lattice problem is mapped onto an effective single site problem, whose magnetization is to reproduce the lattice magnetization through an effective "mean-field". This condition is called the self-consistency condition. It stipulates that the single-site observables should reproduce the lattice "local" observables by means of an effective field. While the N-site Ising Hamiltonian is hard to solve analytically (to date, analytical solutions exist only for the 1D and 2D case), the single-site problem is easily solved.

Likewise, DMFT maps a lattice problem (e.g. the Hubbard model) onto a single-site problem. In DMFT, the local observable is the local Green's function. Thus, the self-consistency condition for DMFT is for the impurity Green's function to reproduce the lattice local Green's function through an effective mean-field which, in DMFT, is the hybridization function of the impurity model. DMFT owes its name to the fact that the mean-field is time-dependent, or dynamical. This also points to the major difference between the Ising MFT and DMFT: Ising MFT maps the N-spin problem into a single-site, single-spin problem. DMFT maps the lattice problem onto a single-site problem, but the latter fundamentally remains a N-body problem which captures the temporal fluctuations due to electron-electron correlations.

Description of DMFT for the Hubbard model

The DMFT mapping

Single-orbital Hubbard model

The Hubbard model [7] describes the onsite interaction between electrons of opposite spin by a single parameter, . The Hubbard Hamiltonian may take the following form:

where, on suppressing the spin 1/2 indices , denote the creation and annihilation operators of an electron on a localized orbital on site , and .

The following assumptions have been made:

  • only one orbital contributes to the electronic properties (as might be the case of copper atoms in superconducting cuprates, whose -bands are non-degenerate),
  • the orbitals are so localized that only nearest-neighbor hopping is taken into account

The auxiliary problem: the Anderson impurity model

The Hubbard model is in general intractable under usual perturbation expansion techniques. DMFT maps this lattice model onto the so-called Anderson impurity model (AIM). This model describes the interaction of one site (the impurity) with a "bath" of electronic levels (described by the annihilation and creation operators and ) through a hybridization function. The Anderson model corresponding to our single-site model is a single-orbital Anderson impurity model, whose hamiltonian formulation, on suppressing some spin 1/2 indices , is:

where

  • describes the non-correlated electronic levels of the bath
  • describes the impurity, where two electrons interact with the energetical cost
  • describes the hybridization (or coupling) between the impurity and the bath through hybridization terms

The Matsubara Green's function of this model, defined by , is entirely determined by the parameters and the so-called hybridization function , which is the imaginary-time Fourier-transform of .

This hybridization function describes the dynamics of electrons hopping in and out of the bath. It should reproduce the lattice dynamics such that the impurity Green's function is the same as the local lattice Green's function. It is related to the non-interacting Green's function by the relation:

(1)

Solving the Anderson impurity model consists in computing observables such as the interacting Green's function for a given hybridization function and . It is a difficult but not intractable problem. There exists a number of ways to solve the AIM, such as

Self-consistency equations

The self-consistency condition requires the impurity Green's function to coincide with the local lattice Green's function :

where denotes the lattice self-energy.

DMFT approximation: locality of the lattice self-energy

The only DMFT approximations (apart from the approximation that can be made in order to solve the Anderson model) consists in neglecting the spatial fluctuations of the lattice self-energy, by equating it to the impurity self-energy:

This approximation becomes exact in the limit of lattices with infinite coordination, that is when the number of neighbors of each site is infinite. Indeed, one can show that in the diagrammatic expansion of the lattice self-energy, only local diagrams survive when one goes into the infinite coordination limit.

Thus, as in classical mean-field theories, DMFT is supposed to get more accurate as the dimensionality (and thus the number of neighbors) increases. Put differently, for low dimensions, spatial fluctuations will render the DMFT approximation less reliable.

Spatial fluctuations also become relevant in the vicinity of phase transitions. Here, DMFT and classical mean-field theories result in mean-field critical exponents, the pronounced changes before the phase transition are not reflected in the DMFT self-energy.

The DMFT loop

In order to find the local lattice Green's function, one has to determine the hybridization function such that the corresponding impurity Green's function will coincide with the sought-after local lattice Green's function. The most widespread way of solving this problem is by using a forward recursion method, namely, for a given , and temperature :

  1. Start with a guess for (typically, )
  2. Make the DMFT approximation:
  3. Compute the local Green's function
  4. Compute the dynamical mean field
  5. Solve the AIM for a new impurity Green's function , extract its self-energy:
  6. Go back to step 2 until convergence, namely when .

Applications

The local lattice Green's function and other impurity observables can be used to calculate a number of physical quantities as a function of correlations , bandwidth, filling (chemical potential ), and temperature :

In particular, the drop of the double-occupancy as increases is a signature of the Mott transition.

Extensions of DMFT

DMFT has several extensions, extending the above formalism to multi-orbital, multi-site problems, long-range correlations and non-equilibrium.

Multi-orbital extension

DMFT can be extended to Hubbard models with multiple orbitals, namely with electron-electron interactions of the form where and denote different orbitals. The combination with density functional theory (DFT+DMFT) [4] [8] then allows for a realistic calculation of correlated materials. [9]

Extended DMFT

Extended DMFT yields a local impurity self energy for non-local interactions and hence allows us to apply DMFT for more general models such as the t-J model.

Cluster DMFT

In order to improve on the DMFT approximation, the Hubbard model can be mapped on a multi-site impurity (cluster) problem, which allows one to add some spatial dependence to the impurity self-energy. Clusters contain 4 to 8 sites at low temperature and up to 100 sites at high temperature.

Diagrammatic extensions

Spatial dependencies of the self energy beyond DMFT, including long-range correlations in the vicinity of a phase transition, can be obtained also through diagrammatic extensions of DMFT [10] using a combination of analytical and numerical techniques. The starting point of the dynamical vertex approximation [11] and of the dual fermion approach is the local two-particle vertex.

Non-equilibrium

DMFT has been employed to study non-equilibrium transport and optical excitations. [12] Here, the reliable calculation of the AIM's Green function out of equilibrium remains a big challenge. DMFT has also been applied to ecological models in order to describe the mean-field dynamics of a community with a thermodynamic number of species. [13]

References and notes

  1. A. Georges; G. Kotliar; W. Krauth; M. Rozenberg (1996). "Dynamical mean-field theory of strongly correlated fermion systems and the limit of infinite dimensions". Reviews of Modern Physics . 68 (1): 13. Bibcode:1996RvMP...68...13G. doi:10.1103/RevModPhys.68.13.
  2. A. Georges and G.Kotliar (1992). "Hubbard model in infinite dimensions". Physical Review B . 45 (12): 6479–6483. Bibcode:1992PhRvB..45.6479G. doi:10.1103/PhysRevB.45.6479. PMID   10000408.
  3. W. Metzner; D. Vollhardt (1989). "Correlated Lattice Fermions in d = ∞ Dimensions". Physical Review Letters . 62 (3): 324–327. Bibcode:1989PhRvL..62..324M. doi:10.1103/PhysRevLett.62.324. PMID   10040203.
  4. 1 2 G. Kotliar; S. Y. Savrasov; K. Haule; V. S. Oudovenko; O. Parcollet; C. A. Marianetti (2006). "Electronic structure calculations with dynamical mean-field theory". Reviews of Modern Physics . 78 (3): 865. arXiv: cond-mat/0511085 . Bibcode:2006RvMP...78..865K. doi:10.1103/RevModPhys.78.865. S2CID   119099745.
  5. D. Vollhardt (2012). "Dynamical mean-field theory for correlated electrons". Annalen der Physik . 524 (1): 1–19. Bibcode:2012AnP...524....1V. doi: 10.1002/andp.201100250 .
  6. Antoine Georges (2004). "Strongly Correlated Electron Materials: Dynamical Mean-Field Theory and Electronic Structure". AIP Conference Proceedings. Lectures on the Physics of Highly Correlated Electron Systems VIII. Vol. 715. American Institute of Physics. pp. 3–74. arXiv: cond-mat/0403123 . doi:10.1063/1.1800733.
  7. John Hubbard (1963). "Electron Correlations in Narrow Energy Bands". Proceedings of the Royal Society A . 276 (1365): 238–257. Bibcode:1963RSPSA.276..238H. doi:10.1098/rspa.1963.0204. S2CID   35439962.
  8. K. Held (2007). "Electronic Structure Calculations using Dynamical Mean Field Theory". Adv. Phys. 56 (6): 829–926. arXiv: cond-mat/0511293 . Bibcode:2007AdPhy..56..829H. doi:10.1080/00018730701619647. S2CID   15466043.
  9. "Embedded Dynamical Mean Field Theory, an electronic structure package implementing DFT+DMFT".
  10. G. Rohringer; H. Hafermann; A. Toschi; A. Katanin; A. E. Antipov; M. I. Katsnelson; A. I. Lichtenstein; A. N. Rubtsov; K. Held (2018). "Diagrammatic routes to nonlocal correlations beyond dynamical mean field theory". Reviews of Modern Physics . 90 (4): 025003. arXiv: 1705.00024 . Bibcode:2018RvMP...90b5003R. doi:10.1103/RevModPhys.90.025003. S2CID   119186041.
  11. A. Toschi; A. Katanin; K. Held (2007). "Dynamical vertex approximation: A step beyond dynamical mean-field theory". Physical Review B . 75 (4): 045118. arXiv: cond-mat/0603100 . Bibcode:2007PhRvB..75d5118T. doi:10.1103/PhysRevB.75.045118. S2CID   119538856.
  12. Aoki, Hideo; Tsuji, Naoto; Eckstein, Martin; Kollar, Marcus; Oka, Takashi; Werner, Philipp (2014-06-24). "Nonequilibrium dynamical mean-field theory and its applications". Reviews of Modern Physics. 86 (2): 779–837. arXiv: 1310.5329 . Bibcode:2014RvMP...86..779A. doi:10.1103/RevModPhys.86.779. ISSN   0034-6861. S2CID   119213862.
  13. Roy, F; Biroli, G; Bunin, G; Cammarota, C (2019-11-29). "Numerical implementation of dynamical mean field theory for disordered systems: application to the Lotka–Volterra model of ecosystems". Journal of Physics A: Mathematical and Theoretical. 52 (48): 484001. arXiv: 1901.10036 . Bibcode:2019JPhA...52V4001R. doi:10.1088/1751-8121/ab1f32. ISSN   1751-8113.

See also

Related Research Articles

In particle physics, the Dirac equation is a relativistic wave equation derived by British physicist Paul Dirac in 1928. In its free form, or including electromagnetic interactions, it describes all spin-1/2 massive particles, called "Dirac particles", such as electrons and quarks for which parity is a symmetry. It is consistent with both the principles of quantum mechanics and the theory of special relativity, and was the first theory to account fully for special relativity in the context of quantum mechanics. It was validated by accounting for the fine structure of the hydrogen spectrum in a completely rigorous way.

<span class="mw-page-title-main">Lattice model (physics)</span>

In mathematical physics, a lattice model is a mathematical model of a physical system that is defined on a lattice, as opposed to a continuum, such as the continuum of space or spacetime. Lattice models originally occurred in the context of condensed matter physics, where the atoms of a crystal automatically form a lattice. Currently, lattice models are quite popular in theoretical physics, for many reasons. Some models are exactly solvable, and thus offer insight into physics beyond what can be learned from perturbation theory. Lattice models are also ideal for study by the methods of computational physics, as the discretization of any continuum model automatically turns it into a lattice model. The exact solution to many of these models includes the presence of solitons. Techniques for solving these include the inverse scattering transform and the method of Lax pairs, the Yang–Baxter equation and quantum groups. The solution of these models has given insights into the nature of phase transitions, magnetization and scaling behaviour, as well as insights into the nature of quantum field theory. Physical lattice models frequently occur as an approximation to a continuum theory, either to give an ultraviolet cutoff to the theory to prevent divergences or to perform numerical computations. An example of a continuum theory that is widely studied by lattice models is the QCD lattice model, a discretization of quantum chromodynamics. However, digital physics considers nature fundamentally discrete at the Planck scale, which imposes upper limit to the density of information, aka Holographic principle. More generally, lattice gauge theory and lattice field theory are areas of study. Lattice models are also used to simulate the structure and dynamics of polymers.

<span class="mw-page-title-main">Drude model</span> Model of electrical conduction

The Drude model of electrical conduction was proposed in 1900 by Paul Drude to explain the transport properties of electrons in materials. Basically, Ohm's law was well established and stated that the current J and voltage V driving the current are related to the resistance R of the material. The inverse of the resistance is known as the conductance. When we consider a metal of unit length and unit cross sectional area, the conductance is known as the conductivity, which is the inverse of resistivity. The Drude model attempts to explain the resistivity of a conductor in terms of the scattering of electrons by the relatively immobile ions in the metal that act like obstructions to the flow of electrons.

<span class="mw-page-title-main">Polaron</span> Quasiparticle in condensed matter physics

A polaron is a quasiparticle used in condensed matter physics to understand the interactions between electrons and atoms in a solid material. The polaron concept was proposed by Lev Landau in 1933 and Solomon Pekar in 1946 to describe an electron moving in a dielectric crystal where the atoms displace from their equilibrium positions to effectively screen the charge of an electron, known as a phonon cloud. This lowers the electron mobility and increases the electron's effective mass.

Plasma oscillations, also known as Langmuir waves, are rapid oscillations of the electron density in conducting media such as plasmas or metals in the ultraviolet region. The oscillations can be described as an instability in the dielectric function of a free electron gas. The frequency depends only weakly on the wavelength of the oscillation. The quasiparticle resulting from the quantization of these oscillations is the plasmon.

A stochastic differential equation (SDE) is a differential equation in which one or more of the terms is a stochastic process, resulting in a solution which is also a stochastic process. SDEs have many applications throughout pure mathematics and are used to model various behaviours of stochastic models such as stock prices, random growth models or physical systems that are subjected to thermal fluctuations.

<span class="mw-page-title-main">Hubbard model</span> Approximate model used to describe the transition between conducting and insulating systems

The Hubbard model is an approximate model used to describe the transition between conducting and insulating systems. It is particularly useful in solid-state physics. The model is named for John Hubbard.

In physics and mathematics, the Gibbs measure, named after Josiah Willard Gibbs, is a probability measure frequently seen in many problems of probability theory and statistical mechanics. It is a generalization of the canonical ensemble to infinite systems. The canonical ensemble gives the probability of the system X being in state x as

In theoretical physics, a scalar–tensor theory is a field theory that includes both a scalar field and a tensor field to represent a certain interaction. For example, the Brans–Dicke theory of gravitation uses both a scalar field and a tensor field to mediate the gravitational interaction.

<span class="mw-page-title-main">Lattice Boltzmann methods</span> Class of computational fluid dynamics methods

The lattice Boltzmann methods (LBM), originated from the lattice gas automata (LGA) method (Hardy-Pomeau-Pazzis and Frisch-Hasslacher-Pomeau models), is a class of computational fluid dynamics (CFD) methods for fluid simulation. Instead of solving the Navier–Stokes equations directly, a fluid density on a lattice is simulated with streaming and collision (relaxation) processes. The method is versatile as the model fluid can straightforwardly be made to mimic common fluid behaviour like vapour/liquid coexistence, and so fluid systems such as liquid droplets can be simulated. Also, fluids in complex environments such as porous media can be straightforwardly simulated, whereas with complex boundaries other CFD methods can be hard to work with.

<i>t</i>-<i>J</i> model

In solid-state physics, the t-J model is a model first derived by Józef Spałek to explain antiferromagnetic properties of Mott insulators, taking into account experimental results about the strength of electron-electron repulsion in this materials.

In fluid dynamics, Airy wave theory gives a linearised description of the propagation of gravity waves on the surface of a homogeneous fluid layer. The theory assumes that the fluid layer has a uniform mean depth, and that the fluid flow is inviscid, incompressible and irrotational. This theory was first published, in correct form, by George Biddell Airy in the 19th century.

In mathematics, the Lindström–Gessel–Viennot lemma provides a way to count the number of tuples of non-intersecting lattice paths, or, more generally, paths on a directed graph. It was proved by Gessel–Viennot in 1985, based on previous work of Lindström published in 1973.

In physics, and especially scattering theory, the momentum-transfer cross section is an effective scattering cross section useful for describing the average momentum transferred from a particle when it collides with a target. Essentially, it contains all the information about a scattering process necessary for calculating average momentum transfers but ignores other details about the scattering angle.

Heat transfer physics describes the kinetics of energy storage, transport, and energy transformation by principal energy carriers: phonons, electrons, fluid particles, and photons. Heat is thermal energy stored in temperature-dependent motion of particles including electrons, atomic nuclei, individual atoms, and molecules. Heat is transferred to and from matter by the principal energy carriers. The state of energy stored within matter, or transported by the carriers, is described by a combination of classical and quantum statistical mechanics. The energy is different made (converted) among various carriers. The heat transfer processes are governed by the rates at which various related physical phenomena occur, such as the rate of particle collisions in classical mechanics. These various states and kinetics determine the heat transfer, i.e., the net rate of energy storage or transport. Governing these process from the atomic level to macroscale are the laws of thermodynamics, including conservation of energy.

<span class="mw-page-title-main">Superradiant phase transition</span> Process in quantum optics

In quantum optics, a superradiant phase transition is a phase transition that occurs in a collection of fluorescent emitters, between a state containing few electromagnetic excitations and a superradiant state with many electromagnetic excitations trapped inside the emitters. The superradiant state is made thermodynamically favorable by having strong, coherent interactions between the emitters.

Dieter Vollhardt is a German physicist and Professor of Theoretical Physics at the University of Augsburg.

In solid state physics, the Luttinger–Ward functional, proposed by Joaquin Mazdak Luttinger and John Clive Ward in 1960, is a scalar functional of the bare electron-electron interaction and the renormalized one-particle propagator. In terms of Feynman diagrams, the Luttinger–Ward functional is the sum of all closed, bold, two-particle irreducible diagrams, i.e., all diagrams without particles going in or out that do not fall apart if one removes two propagator lines. It is usually written as or , where is the one-particle Green's function and is the bare interaction.

<span class="mw-page-title-main">Lorentz oscillator model</span> Theoretical model describing the optical response of bound charges

The Lorentz oscillator model describes the optical response of bound charges. The model is named after the Dutch physicist Hendrik Antoon Lorentz. It is a classical, phenomenological model for materials with characteristic resonance frequencies for optical absorption, e.g. ionic and molecular vibrations, interband transitions (semiconductors), phonons, and collective excitations.

The Typical Medium Dynamical Cluster Approximation (TMDCA) is a non-perturbative approach designed to model and obtain the electronic ground state of strongly correlated many-body systems. It addresses critical aspects of mean-field treatments of strongly correlated systems, such as the lack of an intrinsic order parameter to characterize quantum phase transitions and the description of spatial dependent features. Additionally, the TMDCA tackles the challenge of accurately modeling strongly correlated systems when imperfections disrupt the fundamental assumptions of band theory, as seen in density functional theory, such as the independent particle approximation and material homogeneity.