Dynamical mean-field theory

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]

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 denote the creation and annihilation operators of an electron on a localized orbital on site , and .

The following assumptions have been made:

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 is the following:

where

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.

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. An intuitive method would be the following: for a given , and temperature

  1. First, compute the non-interacting lattice Green's function , and extract its local part
  2. The self-consistency condition requires it to be equal to the impurity Green's function,
  3. Compute the corresponding hybridization function through (1)
  4. Solve the AIM for a new impurity Green's function , extract its self-energy:
  5. Make the DMFT approximation:
  6. Compute the new lattice Green's function, extract its local part and go back to step 2 with a new

Self-consistency is reached 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.

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)[13] then allows for a realistic calculation of correlated materials.

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 a combination of analytical and numerical techniques. The starting point of the dynamical vertex approximation[14] 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. Here, the reliable calculation of the AIM Green function out of equilibrium remains a big challenge.

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. Bibcode:1992PhRvB..45.6479G. doi:10.1103/PhysRevB.45.6479.
  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.
  4. 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/0511085Freely accessible. Bibcode:2006RvMP...78..865K. doi:10.1103/RevModPhys.78.865.
  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. American Institute of Physics Conference. Lectures on the Physics of Highly Correlated Electron Systems VIII. 715 (1). pp. 3–74. arXiv:cond-mat/0403123Freely accessible. doi:10.1063/1.1800733.
  7. John Hubbard (1963). "Electron Correlations in Narrow Energy Bands". Proceedings of the Royal Society A. 276 (1365): 238. Bibcode:1963RSPSA.276..238H. doi:10.1098/rspa.1963.0204.
  8. A. N. Rubtsov; V. V. Savkin; A. I. Lichtenstein (2005). "Continuous-time quantum Monte Carlo method for fermions". Physical Review B. 72 (3): 035122. arXiv:cond-mat/0411344Freely accessible. Bibcode:2005PhRvB..72c5122R. doi:10.1103/PhysRevB.72.035122.
  9. Philipp Werner; Armin Comanac; Luca de’ Medici; Matthias Troyer; Andrew J. Millis (2006). "Continuous-Time Solver for Quantum Impurity Models". Physical Review Letters. 97 (7): 076405. arXiv:cond-mat/0512727Freely accessible. Bibcode:2006PhRvL..97g6405W. doi:10.1103/PhysRevLett.97.076405.
  10. Werner, Philipp; Millis, Andrew J. (2006). "Hybridization expansion impurity solver: General formulation and application to Kondo lattice and two-orbital models". Physical Review B. 74 (15): 155107. arXiv:cond-mat/0607136Freely accessible. Bibcode:2006PhRvB..74o5107W. doi:10.1103/PhysRevB.74.155107.
  11. K. Haule (2007). "Quantum Monte Carlo Impurity Solver for Cluster DMFT and Electronic Structure Calculations". Physical Review B. 75: 155113. arXiv:cond-mat/0612172Freely accessible. Bibcode:2007PhRvB..75o5113H. doi:10.1103/PhysRevB.75.155113.
  12. Gull, Emanuel and Millis, Andrew J. and Lichtenstein, Alexander I. and Rubtsov, Alexey N. and Troyer, Matthias and Werner, Philipp (2011). "Continuous-time Monte~Carlo methods for quantum impurity models". Reviews of Modern Physics. 83: 349. arXiv:1012.4474Freely accessible. Bibcode:2011RvMP...83..349G. doi:10.1103/RevModPhys.83.349.
  13. K. Held (2007). "Electronic Structure Calculations using Dynamical Mean Field Theory". Adv. Phys. 56: 829. arXiv:cond-mat/0511293Freely accessible. Bibcode:2007AdPhy..56..829H. doi:10.1080/00018730701619647.
  14. A. Toschi; A. Katanin; K. Held (2007). "Dynamical vertex approximation: A step beyond dynamical mean-field theory". Physical Review B. 75: 045118. arXiv:cond-mat/0603100Freely accessible. Bibcode:2007PhRvB..75d5118T. doi:10.1103/PhysRevB.75.045118.

See also

External links

This article is issued from Wikipedia - version of the 7/25/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.