Peierls transition

Last updated

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. [1]

Contents

Peierls’ theorem

Peierls' theorem [2] states that a one-dimensional equally spaced chain with one electron per ion is unstable.

The lowest Bloch band of an undistorted 1D lattice. Peierls instability before.jpg
The lowest Bloch band of an undistorted 1D lattice.
The lowest Bloch bands of a distorted 1D lattice. Energy gaps appear in
k
a
=
+-
p
/
2
{\displaystyle ka=\pm \pi /2}
as a result of the Peierls' instability. Peierls instability after.jpg
The lowest Bloch bands of a distorted 1D lattice. Energy gaps appear in as a result of the Peierls' instability.

This theorem was first espoused in the 1930s by Rudolf Peierls. It can be proven using a simple model of the potential for an electron in a 1-D crystal with lattice spacing . The periodicity of the crystal creates energy band gaps in the diagram at the edge of the Brillouin zone (similar to the result of the Kronig–Penney model, which helps to explain the origin of band gaps in semiconductors). If the ions each contribute one electron, then the band will be half-filled, up to values of in the ground state.

Peierls distortion of a 1-d periodic lattice. Peierl'sDistortion1-D.png
Peierls distortion of a 1-d periodic lattice.

Imagine a lattice distortion where every other ion moves closer to one neighbor and further away from the other, the unfavourable energy of the long bond between ions is outweighed by the energy gain of the short bond. The period has just doubled from to . In essence, the proof relies on the fact that doubling the period would introduce new band gaps located at multiples of ; see the figure in the right. This would cause small energy savings, based on the distortion of the bands in the vicinity of the new gaps. Approaching , the distortion due to the introduction of the new band gap will cause the electrons to be at a lower energy than they would be in the perfect crystal. Therefore, this lattice distortion becomes energetically favorable when the energy savings due to the new band gaps outweighs the elastic energy cost of rearranging the ions. Of course, this effect will be noticeable only when the electrons are arranged close to their ground state – in other words, thermal excitation should be minimized. Therefore, the Peierls transition should be seen at low temperature. This is the basic argument for the occurrence of the Peierls transition, sometimes called dimerization.

Historical background

The earliest written record of the Peierls transition was presented at the 1954 École de physique des Houches. These lecture notes (shown below) contain Rudolf Peierls' handwritten equations and figures, and can be viewed [3] in the library of the Institut Laue–Langevin, in Grenoble, France.

Notes from the 1954 Les Houches Conference presenting the Peierls transition. Rudolf Peierl's notes.jpg
Notes from the 1954 Les Houches Conference presenting the Peierls transition.

Peierls’ discovery gained experimental backing during the effort to find new superconducting materials. In 1964, Dr. William Little of the Stanford University Department of Physics theorized that a certain class of polymer chains may experience a high Tc superconducting transition. [4] The basis for his assertion was that the lattice distortions that lead to pairing of electrons in the BCS theory of superconductivity could be replaced instead by rearranging the electron density in a series of side chains. This means that now electrons would be responsible for creating the Cooper pairs instead of ions. Because the transition temperature is inversely proportional to the square root of the mass of the charged particle responsible for the distortions, the Tc should be improved by a corresponding factor:

The subscript i represents "ion", while e represents "electron". The predicted benefit in superconducting transition temperature was therefore a factor of about 300.

In the 1970s, various organic materials such as TTF-TCNQ were synthesized. [5] What was found is that these materials underwent an insulating transition rather than a superconducting one. Eventually it was realized that these were the first experimental observations of the Peierls transition. With the introduction of new band gaps after the lattice becomes distorted, electrons must overcome this new energy barrier in order to become free to conduct. The simple model of the Peierls distortion as a rearrangement of ions in a 1-D chain could describe why these materials became insulators rather than superconductors.

Peierls predicted that the rearrangement of the ion cores in a Peierls transition would produce periodic fluctuations in the electron density. These are commonly called charge density waves, and they are an example of collective charge transport. Several materials systems have verified the existence of these waves. Good candidates are weakly coupled molecular chains, where electrons can move freely along the direction of the chains, but motion is restricted perpendicular to the chains. NbSe3 and K0.3MoO3 are two examples in which charge density waves have been observed at relatively high temperatures of 145 K and 180 K respectively. [6]

Furthermore, the 1-D nature of the material causes a breakdown of the Fermi liquid theory for electron behavior. Therefore, a 1-D conductor should behave as a Luttinger liquid instead. A Luttinger liquid is a paramagnetic one-dimensional metal without Landau quasi-particle excitations.

Research topics

1-D metals have been the subject of much research. Here are a few examples of both theoretical and experimental research efforts to illustrate the broad range of topics:

See also

Related Research Articles

<span class="mw-page-title-main">BCS theory</span> Microscopic theory of superconductivity

In physics, theBardeen–Cooper–Schrieffer (BCS) theory is the first microscopic theory of superconductivity since Heike Kamerlingh Onnes's 1911 discovery. The theory describes superconductivity as a microscopic effect caused by a condensation of Cooper pairs. The theory is also used in nuclear physics to describe the pairing interaction between nucleons in an atomic nucleus.

<span class="mw-page-title-main">Fermi liquid theory</span> Theoretical model in physics

Fermi liquid theory is a theoretical model of interacting fermions that describes the normal state of the conduction electrons in most metals at sufficiently low temperatures. The theory describes the behavior of many-body systems of particles in which the interactions between particles may be strong. 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.

<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.

Jellium, also known as the uniform electron gas (UEG) or homogeneous electron gas (HEG), is a quantum mechanical model of interacting electrons in a solid where the positive charges are assumed to be uniformly distributed in space; the electron density is a uniform quantity as well in space. This model allows one to focus on the effects in solids that occur due to the quantum nature of electrons and their mutual repulsive interactions without explicit introduction of the atomic lattice and structure making up a real material. Jellium is often used in solid-state physics as a simple model of delocalized electrons in a metal, where it can qualitatively reproduce features of real metals such as screening, plasmons, Wigner crystallization and Friedel oscillations.

<span class="mw-page-title-main">Mott insulator</span> Materials classically predicted to be conductors, that are actually insulators

Mott insulators are a class of materials that are expected to conduct electricity according to conventional band theories, but turn out to be insulators. These insulators fail to be correctly described by band theories of solids due to their strong electron–electron interactions, which are not considered in conventional band theory. A Mott transition is a transition from a metal to an insulator, driven by the strong interactions between electrons. One of the simplest models that can capture Mott transition is the Hubbard model.

<span class="mw-page-title-main">Spin density wave</span>

Spin-density wave (SDW) and charge-density wave (CDW) are names for two similar low-energy ordered states of solids. Both these states occur at low temperature in anisotropic, low-dimensional materials or in metals that have high densities of states at the Fermi level . Other low-temperature ground states that occur in such materials are superconductivity, ferromagnetism and antiferromagnetism. The transition to the ordered states is driven by the condensation energy which is approximately where is the magnitude of the energy gap opened by the transition.

Surface states are electronic states found at the surface of materials. They are formed due to the sharp transition from solid material that ends with a surface and are found only at the atom layers closest to the surface. The termination of a material with a surface leads to a change of the electronic band structure from the bulk material to the vacuum. In the weakened potential at the surface, new electronic states can be formed, so called surface states.

<span class="mw-page-title-main">Wigner crystal</span>

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.

<span class="mw-page-title-main">Pseudogap</span> State at which a Fermi surface has a partial energy gap in condensed matter physics

In condensed matter physics, a pseudogap describes a state where the Fermi surface of a material possesses a partial energy gap, for example, a band structure state where the Fermi surface is gapped only at certain points.

Metal–insulator transitions are transitions of a material from a metal to an insulator. These transitions can be achieved by tuning various ambient parameters such as temperature, pressure or, in case of a semiconductor, doping.

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, 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.

Charge ordering (CO) is a phase transition occurring mostly in strongly correlated materials such as transition metal oxides or organic conductors. Due to the strong interaction between electrons, charges are localized on different sites leading to a disproportionation and an ordered superlattice. It appears in different patterns ranging from vertical to horizontal stripes to a checkerboard–like pattern , and it is not limited to the two-dimensional case. The charge order transition is accompanied by symmetry breaking and may lead to ferroelectricity. It is often found in close proximity to superconductivity and colossal magnetoresistance.

Type-1.5 superconductors are multicomponent superconductors characterized by two or more coherence lengths, at least one of which is shorter than the magnetic field penetration length , and at least one of which is longer. This is in contrast to single-component superconductors, where there is only one coherence length and the superconductor is necessarily either type 1 or type 2. When placed in magnetic field, type-1.5 superconductors should form quantum vortices: magnetic-flux-carrying excitations. They allow magnetic field to pass through superconductors due to a vortex-like circulation of superconducting particles. In type-1.5 superconductors these vortices have long-range attractive, short-range repulsive interaction. As a consequence a type-1.5 superconductor in a magnetic field can form a phase separation into domains with expelled magnetic field and clusters of quantum vortices which are bound together by attractive intervortex forces. The domains of the Meissner state retain the two-component superconductivity, while in the vortex clusters one of the superconducting components is suppressed. Thus such materials should allow coexistence of various properties of type-I and type-II superconductors.

<span class="mw-page-title-main">Luttinger's theorem</span>

In condensed matter physics, Luttinger's theorem is a result derived by J. M. Luttinger and J. C. Ward in 1960 that has broad implications in the field of electron transport. It arises frequently in theoretical models of correlated electrons, such as the high-temperature superconductors, and in photoemission, where a metal's Fermi surface can be directly observed.

Lithium molybdenum purple bronze is a chemical compound with formula Li
0.9
Mo
6
O
17
, that is, a mixed oxide of molybdenum and lithium. It can be obtained as flat crystals with a purple-red color and metallic sheen.

Electronic entropy is the entropy of a system attributable to electrons' probabilistic occupation of states. This entropy can take a number of forms. The first form can be termed a density of states based entropy. The Fermi–Dirac distribution implies that each eigenstate of a system, i, is occupied with a certain probability, pi. As the entropy is given by a sum over the probabilities of occupation of those states, there is an entropy associated with the occupation of the various electronic states. In most molecular systems, the energy spacing between the highest occupied molecular orbital and the lowest unoccupied molecular orbital is usually large, and thus the probabilities associated with the occupation of the excited states are small. Therefore, the electronic entropy in molecular systems can safely be neglected. Electronic entropy is thus most relevant for the thermodynamics of condensed phases, where the density of states at the Fermi level can be quite large, and the electronic entropy can thus contribute substantially to thermodynamic behavior. A second form of electronic entropy can be attributed to the configurational entropy associated with localized electrons and holes. This entropy is similar in form to the configurational entropy associated with the mixing of atoms on a lattice.

<span class="mw-page-title-main">Fulleride</span> Chemical compound

Fullerides are chemical compounds containing fullerene anions. Common fullerides are derivatives of the most common fullerenes, i.e. C60 and C70. The scope of the area is large because multiple charges are possible, i.e., [C60]n (n = 1, 2...6), and all fullerenes can be converted to fullerides. The suffix "-ide" implies their negatively charged nature.

The term Dirac matter refers to a class of condensed matter systems which can be effectively described by the Dirac equation. Even though the Dirac equation itself was formulated for fermions, the quasi-particles present within Dirac matter can be of any statistics. As a consequence, Dirac matter can be distinguished in fermionic, bosonic or anyonic Dirac matter. Prominent examples of Dirac matter are graphene and other Dirac semimetals, topological insulators, Weyl semimetals, various high-temperature superconductors with -wave pairing and liquid helium-3. The effective theory of such systems is classified by a specific choice of the Dirac mass, the Dirac velocity, the gamma matrices and the space-time curvature. The universal treatment of the class of Dirac matter in terms of an effective theory leads to a common features with respect to the density of states, the heat capacity and impurity scattering.

<span class="mw-page-title-main">Electron-on-helium qubit</span> Quantum bit

An electron-on-helium qubit is a quantum bit for which the orthonormal basis states |0⟩ and |1⟩ are defined by quantized motional states or alternatively the spin states of an electron trapped above the surface of liquid helium. The electron-on-helium qubit was proposed as the basic element for building quantum computers with electrons on helium by Platzman and Dykman in 1999. 

References

  1. "Peierls transition". IUPAC Goldbook. 2014. doi: 10.1351/goldbook.P04468 .
  2. Fowler, Michael (28 Feb 2007). "Electrons in One Dimension: the Peierls Transition".
  3. {http://www.epn-campus.eu/library/joint-ill-esrf-library/}
  4. W. A. Little (1964). "Possibility of Synthesizing an Organic Superconductor". Physical Review . 134 (6A): A1416–A1424. Bibcode:1964PhRv..134.1416L. doi:10.1103/PhysRev.134.A1416.
  5. P. W. Anderson; P. A. Lee; M. Saitoh (1973). "Remarks on giant conductivity in TTF-TCNQ". Solid State Communications . 13 (5): 595–598. Bibcode:1973SSCom..13..595A. doi:10.1016/S0038-1098(73)80020-1.
  6. Thorne, Robert (May 1996). "Charge-Density-Wave Conductors" (PDF). Physics Today .
  7. S. D. Liang; Y. H. Bai; B. Beng (2006). "Peierls instability and persistent current in mesoscopic conducting polymer rings". Physical Review B . 74 (11): 113304. Bibcode:2006PhRvB..74k3304L. doi:10.1103/PhysRevB.74.113304.
  8. D. Jacquemin; A. Femenias; H. Chermette; I. Ciofini; C. Adamo; J. M. Andr; E. A. Perpte (2006). "Assessment of Several Hybrid DFT Functionals for the Evaluation of Bond Length Alternation of Increasingly Long Oligomers". Journal of Physical Chemistry A . 110 (17): 5952–5959. Bibcode:2006JPCA..110.5952J. doi:10.1021/jp060541w. PMID   16640395.
  9. J. R. Ahn; P. G. Kang; K. D. Ryang; H.W. Yeom (2005). "Coexistence of Two Different Peierls Distortions within an Atomic Scale Wire: Si(553)-Au". Physical Review Letters . 95 (19): 196402. Bibcode:2005PhRvL..95s6402A. doi:10.1103/PhysRevLett.95.196402. PMID   16384001.
  10. Voit, Johannes (5 May 2000). "A brief introduction to Luttinger liquids". AIP Conference Proceedings. 544: 309–318. arXiv: cond-mat/0005114 . Bibcode:2000AIPC..544..309V. doi:10.1063/1.1342524. S2CID   117040555.
  11. C. A. M. dos Santos; M. S. da Luz; Yi-Kuo Yu; J. J. Neumeier; J. Moreno; B. D. White (2008). "Electrical transport in single-crystalline Li0.9Mo6O17: A two-band Luttinger liquid exhibiting Bose metal behavior". Physical Review B . 77 (19): 193106. Bibcode:2008PhRvB..77s3106D. doi:10.1103/PhysRevB.77.193106.
  12. F. Wang; J.V. Alvarez; S.-K. Mo; J. W. Allen; G.-H. Gweon; J. He; R. Jin; D. Mandrus; H. Höchst (2006). "New Luttinger-Liquid Physics from Photoemission on Li0.9Mo6O17". Physical Review Letters . 96 (19): 196403. arXiv: cond-mat/0604503 . Bibcode:2006PhRvL..96s6403W. doi:10.1103/PhysRevLett.96.196403. PMID   16803117. S2CID   10365828.
  13. O. M. Auslaender; A. Yacoby; R. de Picciotto; K.W. Baldwin; L. N. Pfeiffer; K.W. West (2000). "Experimental evidence for resonant tunneling in a Luttinger liquid". Physical Review Letters . 84 (8): 1764–1767. arXiv: cond-mat/9909138 . Bibcode:2000PhRvL..84.1764A. doi:10.1103/PhysRevLett.84.1764. PMID   11017620. S2CID   11317080.