Coulomb gap

Last updated

First introduced by M. Pollak, [1] the Coulomb gap is a soft gap in the single-particle density of states (DOS) of a system of interacting localized electrons. Due to the long-range Coulomb interactions, the single-particle DOS vanishes at the chemical potential, at low enough temperatures, such that thermal excitations do not wash out the gap.

Contents

Theory

At zero temperature, a classical treatment of a system gives an upper bound for the DOS near the Fermi energy, first suggested by Efros and Shklovskii. [2] The argument is as follows: Let us look at the ground state configuration of the system. Defining as the energy of an electron at site , due to the disorder and the Coulomb interaction with all other electrons (we define this both for occupied and unoccupied sites), it is easy to see that the energy needed to move an electron from an occupied site to an unoccupied site is given by the expression:

.

The subtraction of the last term accounts for the fact that contains a term due to the interaction with the electron present at site , but after moving the electron this term should not be considered. It is easy to see from this that there exists an energy such that all sites with energies above it are empty, and below it are full (this is the Fermi energy, but since we are dealing with a system with interactions it is not obvious a-priori that it is still well-defined). Assume we have a finite single-particle DOS at the Fermi energy, . For every possible transfer of an electron from an occupied site to an unoccupied site , the energy invested should be positive, since we are assuming we are in the ground state of the system, i.e., . Assuming we have a large system, consider all the sites with energies in the interval The number of these, by assumption, is As explained, of these would be occupied, and the others unoccupied. Of all pairs of occupied and unoccupied sites, let us choose the one where the two are closest to each other. If we assume the sites are randomly distributed in space, we find that the distance between these two sites is of order: , where is the dimension of space. Plugging the expression for into the previous equation, we obtain the inequality: where is a coefficient of order unity. Since , this inequality will necessarily be violated for small enough . Hence, assuming a finite DOS at led to a contradiction. Repeating the above calculation under the assumption that the DOS near is proportional to shows that . This is an upper bound for the Coulomb gap. Efros [3] considered single electron excitations, and obtained an integro-differential equation for the DOS, showing the Coulomb gap in fact follows the above equation (i.e., the upper bound is a tight bound).

Other treatments of the problem include a mean-field numerical approach, [4] as well as more recent treatments such as, [5] also verifying the upper bound suggested above is a tight bound. Many Monte Carlo simulations were also performed, [6] [7] some of them in disagreement with the result quoted above. Few works deal with the quantum aspect of the problem. [8] Classical Coulomb gap in clean system without disorder is well captured within Extended Dynamical Mean Field Theory (EDMFT) supported by Metropolis Monte Carlo simulations. [9]

Experimental observations

Direct experimental confirmation of the gap has been done via tunneling experiments, which probed the single-particle DOS in two and three dimensions. [10] [11] The experiments clearly showed a linear gap in two dimensions, and a parabolic gap in three dimensions. Another experimental consequence of the Coulomb gap is found in the conductivity of samples in the localized regime. The existence of a gap in the spectrum of excitations would result in a lowered conductivity than that predicted by Mott variable-range hopping. If one uses the analytical expression of the single-particle DOS in the Mott derivation, a universal dependence is obtained, for any dimension. [12] The observation of this is expected to occur below a certain temperature, such that the optimal energy of hopping would be smaller than the width of the Coulomb gap. The transition from Mott to so-called Efros–Shklovskii variable-range hopping has been observed experimentally for various systems. [13] Nevertheless, no rigorous derivation of the Efros–Shklovskii conductivity formula has been put forth, and in some experiments behavior is observed, with a value of that fits neither the Mott nor the Efros–Shklovskii theories.

See also

Related Research Articles

The quantum Hall effect is a quantized version of the Hall effect which is observed in two-dimensional electron systems subjected to low temperatures and strong magnetic fields, in which the Hall resistance Rxy exhibits steps that take on the quantized values

Fermi liquid theory Theoretical model of interacting fermions

Fermi liquid theory is a theoretical model of interacting fermions that describes the normal state of most metals at sufficiently low temperatures. The interactions among the particles of the many-body system do not need to be small. The phenomenological theory of Fermi liquids was introduced by the Soviet physicist Lev Davidovich Landau in 1956, and later developed by Alexei Abrikosov and Isaak Khalatnikov using diagrammatic perturbation theory. The theory explains why some of the properties of an interacting fermion system are very similar to those of the ideal Fermi gas, and why other properties differ.

Polaron 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. For comparison of the models proposed in these papers see M. I. Dykman and E. I. Rashba, The roots of polaron theory, Physics Today 68, 10 (2015). This lowers the electron mobility and increases the electron's effective mass.

Scanning tunneling spectroscopy (STS), an extension of scanning tunneling microscopy (STM), is used to provide information about the density of electrons in a sample as a function of their energy.

Koopmans' theorem states that in closed-shell Hartree–Fock theory (HF), the first ionization energy of a molecular system is equal to the negative of the orbital energy of the highest occupied molecular orbital (HOMO). This theorem is named after Tjalling Koopmans, who published this result in 1934.

Local-density approximations (LDA) are a class of approximations to the exchange–correlation (XC) energy functional in density functional theory (DFT) that depend solely upon the value of the electronic density at each point in space. Many approaches can yield local approximations to the XC energy. However, overwhelmingly successful local approximations are those that have been derived from the homogeneous electron gas (HEG) model. In this regard, LDA is generally synonymous with functionals based on the HEG approximation, which are then applied to realistic systems.

Wigner crystal

A Wigner crystal is the solid (crystalline) phase of electrons first predicted by Eugene Wigner in 1934. A gas of electrons moving in a uniform, inert, neutralizing background will crystallize and form a lattice if the electron density is less than a critical value. This is because the potential energy dominates the kinetic energy at low densities, so the detailed spatial arrangement of the electrons becomes important. To minimize the potential energy, the electrons form a bcc lattice in 3D, a triangular lattice in 2D and an evenly spaced lattice in 1D. Most experimentally observed Wigner clusters exist due to the presence of the external confinement, i.e. external potential trap. As a consequence, deviations from the b.c.c or triangular lattice are observed. A crystalline state of the 2D electron gas can also be realized by applying a sufficiently strong magnetic field. However, it is still not clear whether it is the Wigner crystallization that has led to observation of insulating behaviour in magnetotransport measurements on 2D electron systems, since other candidates are present, such as Anderson localization.

In solid-state physics, an energy gap is an energy range in a solid where no electron states exist, i.e. an energy range where the density of states vanishes.

A charge density wave (CDW) is an ordered quantum fluid of electrons in a linear chain compound or layered crystal. The electrons within a CDW form a standing wave pattern and sometimes collectively carry an electric current. The electrons in such a CDW, like those in a superconductor, can flow through a linear chain compound en masse, in a highly correlated fashion. Unlike a superconductor, however, the electric CDW current often flows in a jerky fashion, much like water dripping from a faucet due to its electrostatic properties. In a CDW, the combined effects of pinning and electrostatic interactions likely play critical roles in the CDW current's jerky behavior, as discussed in sections 4 & 5 below.

In materials science, effective medium approximations (EMA) or effective medium theory (EMT) pertain to analytical or theoretical modeling that describes the macroscopic properties of composite materials. EMAs or EMTs are developed from averaging the multiple values of the constituents that directly make up the composite material. At the constituent level, the values of the materials vary and are inhomogeneous. Precise calculation of the many constituent values is nearly impossible. However, theories have been developed that can produce acceptable approximations which in turn describe useful parameters including the effective permittivity and permeability of the materials as a whole. In this sense, effective medium approximations are descriptions of a medium based on the properties and the relative fractions of its components and are derived from calculations, and effective medium theory. There are two widely used formulae.

In solid-state physics, heavy fermion materials are a specific type of intermetallic compound, containing elements with 4f or 5f electrons in unfilled electron bands. Electrons are one type of fermion, and when they are found in such materials, they are sometimes referred to as heavy electrons. Heavy fermion materials have a low-temperature specific heat whose linear term is up to 1000 times larger than the value expected from the free electron model. The properties of the heavy fermion compounds often derive from the partly filled f-orbitals of rare-earth or actinide ions, which behave like localized magnetic moments. The name "heavy fermion" comes from the fact that the fermion behaves as if it has an effective mass greater than its rest mass. In the case of electrons, below a characteristic temperature (typically 10 K), the conduction electrons in these metallic compounds behave as if they had an effective mass up to 1000 times the free particle mass. This large effective mass is also reflected in a large contribution to the resistivity from electron-electron scattering via the Kadowaki–Woods ratio. Heavy fermion behavior has been found in a broad variety of states including metallic, superconducting, insulating and magnetic states. Characteristic examples are CeCu6, CeAl3, CeCu2Si2, YbAl3, UBe13 and UPt3.

The Anderson impurity model, named after Philip Warren Anderson, is a Hamiltonian that is used to describe magnetic impurities embedded in metals. It is often applied to the description of Kondo effect-type problems, such as heavy fermion systems and Kondo insulators. In its simplest form, the model contains a term describing the kinetic energy of the conduction electrons, a two-level term with an on-site Coulomb repulsion that models the impurity energy levels, and a hybridization term that couples conduction and impurity orbitals. For a single impurity, the Hamiltonian takes the form

In density functional theory (DFT), the Harris energy functional is a non-self-consistent approximation to the Kohn–Sham density functional theory. It gives the energy of a combined system as a function of the electronic densities of the isolated parts. The energy of the Harris functional varies much less than the energy of the Kohn–Sham functional as the density moves away from the converged density.

A Peierls transition or Peierls distortion is a distortion of the periodic lattice of a one-dimensional crystal. Atomic positions oscillate, so that the perfect order of the 1-D crystal is broken.

In condensed matter physics, biexcitons are created from two free excitons.

A composite fermion is the topological bound state of an electron and an even number of quantized vortices, sometimes visually pictured as the bound state of an electron and, attached, an even number of magnetic flux quanta. Composite fermions were originally envisioned in the context of the fractional quantum Hall effect, but subsequently took on a life of their own, exhibiting many other consequences and phenomena.

Superradiant phase transition

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.

The Spitzer resistivity is an expression describing the electrical resistance in a plasma, which was first formulated by Lyman Spitzer in 1950. The Spitzer resistivity of a plasma decreases in proportion to the electron temperature as .

Dark photon Hypothetical force carrier particle connected to dark matter

The dark photon is a hypothetical hidden sector particle, proposed as a force carrier similar to the photon of electromagnetism but potentially connected to dark matter. In a minimal scenario, this new force can be introduced by extending the gauge group of the Standard Model of Particle Physics with a new abelian U(1) gauge symmetry. The corresponding new spin-1 gauge boson can then couple very weakly to electrically charged particles through kinetic mixing with the ordinary photon and could thus be detected. The dark photon can also interact with the Standard Model if some of the fermions are charged under the new abelian group. The possible charging arrangements are restricted by a number of consistency requirements such as anomaly cancellation and constraints coming from Yukawa matrices.

Ionic Coulomb blockade (ICB) is an electrostatic phenomenon that appears in ionic transport through mesoscopic electro-diffusive systems and manifests itself as oscillatory dependences of the conductance on the fixed charge in the pore.

References

  1. M. Pollak (1970). "Effect of carrier-carrier interactions on some transport properties in disordered semiconductors". Discussions of the Faraday Society. 50: 13. doi:10.1039/DF9705000013.
  2. A L Efros and B I Shklovskii (1975). "Coulomb gap and low temperature conductivity of disordered systems". Journal of Physics C. 8 (4): L49. Bibcode:1975JPhC....8L..49E. doi:10.1088/0022-3719/8/4/003.
  3. A. L. Efros (1976). "Coulomb gap in disordered systems". Journal of Physics C: Solid State Physics. 9 (11): 2021. Bibcode:1976JPhC....9.2021E. doi:10.1088/0022-3719/9/11/012.
  4. M. Grunewald, B. Pohlmann, L. Schweitzer, and D.Wurtz (1982). "Mean field approach to the electron glass". Journal of Physics C: Solid State Physics. 15 (32): L1153. doi:10.1088/0022-3719/15/32/007.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  5. M. Muller and S. Pankov (2007). "Mean-field theory for the three-dimensional Coulomb glass". Physical Review B. 75 (14): 144201. arXiv: cond-mat/0611021 . Bibcode:2007PhRvB..75n4201M. doi:10.1103/PhysRevB.75.144201. S2CID   119419036.
  6. J. H. Davies, P. A. Lee, and T. M. Rice (1982). "Electron Glass". Physical Review Letters. 49 (10): 758-761. Bibcode:1982PhRvL..49..758D. doi:10.1103/PhysRevLett.49.758.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  7. A. Mobius, M. Richter, and B. Drittler (1992). "Coulomb gap in two- and three-dimensional systems: Simulation results for large samples". Physical Review B. 45 (20): 11568–11579. Bibcode:1992PhRvB..4511568M. doi:10.1103/PhysRevB.45.11568. PMID   10001170.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  8. G. Vignale (1987). "Quantum electron glass". Physical Review B. 36 (15): 8192–8195. Bibcode:1987PhRvB..36.8192V. doi:10.1103/PhysRevB.36.8192. PMID   9942629.
  9. Pramudya, Y.; Terletska, H.; Pankov, S.; Manousakis, E.; Dobrosavljević, V. (2011-09-12). "Nearly frozen Coulomb liquids". Physical Review B. 84 (12): 125120. arXiv: 1012.2396 . Bibcode:2011PhRvB..84l5120P. doi: 10.1103/PhysRevB.84.125120 .
  10. J. G. Massey and M. Lee (1995). "Direct Observation of the Coulomb Correlation Gap in a Nonmetallic Semiconductor, Si: B". Physical Review Letters. 75 (23): 4266–4269. Bibcode:1995PhRvL..75.4266M. doi:10.1103/PhysRevLett.75.4266. PMID   10059861.
  11. V. Y. Butko, J. F. Ditusa, and P. W. Adams (2000). "Coulomb Gap: How a Metal Film Becomes an Insulator". Physical Review Letters. 84 (7): 1543–6. arXiv: cond-mat/0006025 . Bibcode:2000PhRvL..84.1543B. doi:10.1103/PhysRevLett.84.1543. PMID   11017563. S2CID   40065110.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  12. B. Shklovskii and A. Efros, Electronic properties of doped semiconductors (Springer-Verlag, Berlin, 1984).
  13. Rogatchev, A.Yu.; Mizutani, U. (2000). "Hopping conductivity and specific heat in insulating amorphousTixSi100−x alloys". Physical Review B. 61 (23): 15550–15553. Bibcode:2000PhRvB..6115550R. doi:10.1103/PhysRevB.61.15550. ISSN   0163-1829.