Heavy fermion material

Last updated

In Materials Science, heavy fermion materials are a specific type of intermetallic compound, containing elements with 4f or 5f electrons in unfilled electron bands. [1] Electrons are one type of fermion, and when they are found in such materials, they are sometimes referred to as heavy electrons. [2] 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.

Contents

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.

Historical overview

Heavy fermion behavior was discovered by K. Andres, J.E. Graebner and H.R. Ott in 1975, who observed enormous magnitudes of the linear specific heat capacity in CeAl3. [3]

While investigations on doped superconductors led to the conclusion that the existence of localized magnetic moments and superconductivity in one material was incompatible, the opposite was shown, when in 1979 Frank Steglich et al. discovered heavy fermion superconductivity in the material CeCu2Si2. [4]

In 1994, the discovery of a quantum critical point and non-Fermi liquid behavior in the phase diagram of heavy fermion compounds by H. von Löhneysen et al. led to a new rise of interest in the research of these compounds. [5] Another experimental breakthrough was the demonstration in 1998 (by the group of Gil Lonzarich) that quantum criticality in heavy fermions can be the reason for unconventional superconductivity. [6]

Heavy fermion materials play an important role in current scientific research, acting as prototypical materials for unconventional superconductivity, non-Fermi liquid behavior and quantum criticality. The actual interaction between localized magnetic moments and conduction electrons in heavy fermion compounds is still not completely understood and a topic of ongoing investigation.[ citation needed ]

Properties

Heavy fermion materials belong to the group of strongly correlated electron systems.

Several members of the group of heavy fermion materials become superconducting below a critical temperature. The superconductivity is unconventional, ie. not covered by BCS theory.

At high temperatures, heavy fermion compounds behave like normal metals and the electrons can be described as a Fermi gas, in which the electrons are assumed to be non-interacting fermions. In this case, the interaction between the f electrons, which present a local magnetic moment and the conduction electrons, can be neglected.

The Fermi liquid theory of Lev Landau provides a good model to describe the properties of most heavy fermion materials at low temperatures. In this theory, the electrons are described by quasiparticles, which have the same quantum numbers and charge, but the interaction of the electrons is taken into account by introducing an effective mass, which differs from the actual mass of a free electron.

Optical properties

Typical frequency-dependent optical conductivity of a heavy fermion compound. Blue line: T > Tcoh. Red line: T < Tcoh. Optical properties heavy fermion.png
Typical frequency-dependent optical conductivity of a heavy fermion compound. Blue line: T > Tcoh. Red line: T < Tcoh.

In order to obtain the optical properties of heavy fermion systems, these materials have been investigated by optical spectroscopy measurements. [7] In these experiments the sample is irradiated by electromagnetic waves with tunable wavelength. Measuring the reflected or transmitted light reveals the characteristic energies of the sample.

Above the characteristic coherence temperature , heavy fermion materials behave like normal metals; i.e. their optical response is described by the Drude model. Compared to a good metal however, heavy fermion compounds at high temperatures have a high scattering rate because of the large density of local magnetic moments (at least one f electron per unit cell), which cause (incoherent) Kondo scattering. Due to the high scattering rate, the conductivity for dc and at low frequencies is rather low. A conductivity roll-off (Drude roll-off) occurs at the frequency that corresponds to the relaxation rate.

Below , the localized f electrons hybridize with the conduction electrons. This leads to the enhanced effective mass, and a hybridization gap develops. In contrast to Kondo insulators, the chemical potential of heavy fermion compounds lies within the conduction band. These changes lead to two important features in the optical response of heavy fermions. [1]

The frequency-dependent conductivity of heavy-fermion materials can be expressed by , containing the effective mass and the renormalized relaxation rate . [8] Due to the large effective mass, the renormalized relaxation time is also enhanced, leading to a narrow Drude roll-off at very low frequencies compared to normal metals. [8] [9] The lowest such Drude relaxation rate observed in heavy fermions so far, in the low GHz range, was found in UPd2Al3. [10]

The gap-like feature in the optical conductivity represents directly the hybridization gap, which opens due to the interaction of localized f electrons and conduction electrons. Since the conductivity does not vanish completely, the observed gap is actually a pseudogap. [11] At even higher frequencies we can observe a local maximum in the optical conductivity due to normal interband excitations. [1]

Heat capacity

Specific heat for normal metals

At low temperature and for normal metals, the specific heat consists of the specific heat of the electrons which depends linearly on temperature and of the specific heat of the crystal lattice vibrations (phonons) which depends cubically on temperature

with proportionality constants and .

In the temperature range mentioned above, the electronic contribution is the major part of the specific heat. In the free electron model a simple model system that neglects electron interaction or metals that could be described by it, the electronic specific heat is given by

with Boltzmann constant , the electron density and the Fermi energy (the highest single particle energy of occupied electronic states). The proportionality constant is called the Sommerfeld coefficient.

Relation between heat capacity and "thermal effective mass"

For electrons with a quadratic dispersion relation (as for the free-electron gas), the Fermi energy εF is inversely proportional to the particle's mass m:

where stands for the Fermi wave number that depends on the electron density and is the absolute value of the wave number of the highest occupied electron state. Thus, because the Sommerfeld parameter is inversely proportional to , is proportional to the particle's mass and for high values of , the metal behaves as a Fermi gas in which the conduction electrons have a high thermal effective mass.

Example: UBe13 at low temperatures

Experimental results for the specific heat of the heavy fermion compound UBe13 show a peak at a temperature around 0.75 K that goes down to zero with a high slope if the temperature approaches 0 K. Due to this peak, the factor is much higher than the free electron model in this temperature range. In contrast, above 6 K, the specific heat for this heavy fermion compound approaches the value expected from free-electron theory.

Quantum criticality

The presence of local moment and delocalized conduction electrons leads to a competition of the Kondo interaction (which favors a non-magnetic ground state) and the RKKY interaction (which generates magnetically ordered states, typically antiferromagnetic for heavy fermions). By suppressing the Néel temperature of a heavy-fermion antiferromagnet down to zero (e.g. by applying pressure or magnetic field or by changing the material composition), a quantum phase transition can be induced. [12] For several heavy-fermion materials it was shown that such a quantum phase transition can generate very pronounced non-Fermi liquid properties at finite temperatures. Such quantum-critical behavior is also studied in great detail in the context of unconventional superconductivity.

Examples of heavy-fermion materials with well-studied quantum-critical properties are CeCu6−xAu, [13] CeIn3, [6] CePd2Si2, [6] YbRh2Si2, and CeCoIn5. [14] [15]

Some heavy fermion compounds

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.

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

Unconventional superconductors are materials that display superconductivity which does not conform to conventional BCS theory or its extensions.

In solid state physics, a particle's effective mass is the mass that it seems to have when responding to forces, or the mass that it seems to have when interacting with other identical particles in a thermal distribution. One of the results from the band theory of solids is that the movement of particles in a periodic potential, over long distances larger than the lattice spacing, can be very different from their motion in a vacuum. The effective mass is a quantity that is used to simplify band structures by modeling the behavior of a free particle with that mass. For some purposes and some materials, the effective mass can be considered to be a simple constant of a material. In general, however, the value of effective mass depends on the purpose for which it is used, and can vary depending on a number of factors.

<span class="mw-page-title-main">Fermi gas</span> Physical model of gases composed of many non-interacting identical fermions

A Fermi gas is an idealized model, an ensemble of many non-interacting fermions. Fermions are particles that obey Fermi–Dirac statistics, like electrons, protons, and neutrons, and, in general, particles with half-integer spin. These statistics determine the energy distribution of fermions in a Fermi gas in thermal equilibrium, and is characterized by their number density, temperature, and the set of available energy states. The model is named after the Italian physicist Enrico Fermi.

<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">Kondo effect</span> Physical phenomenon due to impurities

In physics, the Kondo effect describes the scattering of conduction electrons in a metal due to magnetic impurities, resulting in a characteristic change i.e. a minimum in electrical resistivity with temperature. The cause of the effect was first explained by Jun Kondo, who applied third-order perturbation theory to the problem to account for scattering of s-orbital conduction electrons off d-orbital electrons localized at impurities. Kondo's calculation predicted that the scattering rate and the resulting part of the resistivity should increase logarithmically as the temperature approaches 0 K. Experiments in the 1960s by Myriam Sarachik at Bell Laboratories provided the first data that confirmed the Kondo effect. Extended to a lattice of magnetic impurities, the Kondo effect likely explains the formation of heavy fermions and Kondo insulators in intermetallic compounds, especially those involving rare earth elements such as cerium, praseodymium, and ytterbium, and actinide elements such as uranium. The Kondo effect has also been observed in quantum dot systems.

<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.

In solid-state physics, the free electron model is a quantum mechanical model for the behaviour of charge carriers in a metallic solid. It was developed in 1927, principally by Arnold Sommerfeld, who combined the classical Drude model with quantum mechanical Fermi–Dirac statistics and hence it is also known as the Drude–Sommerfeld model.

<span class="mw-page-title-main">Magnon</span> Spin 1 quasiparticle; quantum of a spin wave

A magnon is a quasiparticle, a collective excitation of the spin structure of an electron in a crystal lattice. In the equivalent wave picture of quantum mechanics, a magnon can be viewed as a quantized spin wave. Magnons carry a fixed amount of energy and lattice momentum, and are spin-1, indicating they obey boson behavior.

A quantum critical point is a point in the phase diagram of a material where a continuous phase transition takes place at absolute zero. A quantum critical point is typically achieved by a continuous suppression of a nonzero temperature phase transition to zero temperature by the application of a pressure, field, or through doping. Conventional phase transitions occur at nonzero temperature when the growth of random thermal fluctuations leads to a change in the physical state of a system. Condensed matter physics research over the past few decades has revealed a new class of phase transitions called quantum phase transitions which take place at absolute zero. In the absence of the thermal fluctuations which trigger conventional phase transitions, quantum phase transitions are driven by the zero point quantum fluctuations associated with Heisenberg's uncertainty principle.

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

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.

<span class="mw-page-title-main">Piers Coleman</span> British-American physicist

Piers Coleman is a British-born theoretical physicist, working in the field of theoretical condensed matter physics. Coleman is professor of physics at Rutgers University in New Jersey and at Royal Holloway, University of London.

Heavy fermion superconductors are a type of unconventional superconductor.

The Fulde–Ferrell–Larkin–Ovchinnikov (FFLO) phase can arise in a superconductor in large magnetic field. Among its characteristics are Cooper pairs with nonzero total momentum and a spatially non-uniform order parameter, leading to normal conducting areas in the superconductor.

In condensed matter physics, a quantum spin liquid is a phase of matter that can be formed by interacting quantum spins in certain magnetic materials. Quantum spin liquids (QSL) are generally characterized by their long-range quantum entanglement, fractionalized excitations, and absence of ordinary magnetic order.

In quantum mechanics, orbital magnetization, Morb, refers to the magnetization induced by orbital motion of charged particles, usually electrons in solids. The term "orbital" distinguishes it from the contribution of spin degrees of freedom, Mspin, to the total magnetization. A nonzero orbital magnetization requires broken time-reversal symmetry, which can occur spontaneously in ferromagnetic and ferrimagnetic materials, or can be induced in a non-magnetic material by an applied magnetic field.

CeCoIn5 ("Cerium-Cobalt-Indium 5") is a heavy-fermion superconductor with a layered crystal structure, with somewhat two-dimensional electronic transport properties. The critical temperature of 2.3 K is the highest among all of the Ce-based heavy-fermion superconductors.

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.

References

  1. 1 2 3 P. Coleman (2007). "Heavy Fermions: Electrons at the Edge of Magnetism. Handbook of Magnetism and Advanced Magnetic Materials". In Helmut Kronmuller; Stuart Parkin (eds.). Handbook of Magnetism and Advanced Magnetic Materials. Vol. 1. pp. 95–148. arXiv: cond-mat/0612006 .
  2. "First images of heavy electrons in action". physorg.com. June 2, 2010.
  3. K. Andres; J.E. Graebner; H.R. Ott (1975). "4f-Virtual-Bound-State Formation in CeAl3 at Low Temperatures". Physical Review Letters. 35 (26): 1779–1782. Bibcode:1975PhRvL..35.1779A. doi:10.1103/PhysRevLett.35.1779.
  4. Steglich, F.; Aarts, J.; Bredl, C. D.; Lieke, W.; Meschede, D.; Franz, W.; Schäfer, H. (1979-12-17). "Superconductivity in the Presence of Strong Pauli Paramagnetism: CeCu2Si2". Physical Review Letters. 43 (25): 1892–1896. Bibcode:1979PhRvL..43.1892S. doi:10.1103/PhysRevLett.43.1892. hdl: 1887/81461 .
  5. Löhneysen, H. v.; Pietrus, T.; Portisch, G.; Schlager, H. G.; Schröder, A.; Sieck, M.; Trappmann, T. (1994-05-16). "Non-Fermi-liquid behavior in a heavy-fermion alloy at a magnetic instability". Physical Review Letters. 72 (20): 3262–3265. Bibcode:1994PhRvL..72.3262L. doi:10.1103/PhysRevLett.72.3262. PMID   10056148.
  6. 1 2 3 Mathur, N.D.; Grosche, F.M.; Julian, S.R.; Walker, I.R.; Freye, D.M.; Haselwimmer, R.K.W.; Lonzarich, G.G. (1998). "Magnetically mediated superconductivity in heavy fermion compounds". Nature. 394 (6688): 39–43. Bibcode:1998Natur.394...39M. doi:10.1038/27838. S2CID   52837444.
  7. L. Degiorgi (1999). "The electrodynamic response of heavy-electron compounds". Reviews of Modern Physics. 71 (3): 687–734. Bibcode:1999RvMP...71..687D. doi:10.1103/RevModPhys.71.687.
  8. 1 2 A.J. Millis; P.A. Lee (1987). "Large-orbital-degeneracy expansion for the lattice Anderson model". Physical Review B. 35 (7): 3394–3414. Bibcode:1987PhRvB..35.3394M. doi:10.1103/PhysRevB.35.3394. PMID   9941843.
  9. M. Scheffler; K. Schlegel; C. Clauss; D. Hafner; C. Fella; M. Dressel; M. Jourdan; J. Sichelschmidt; C. Krellner; C. Geibel; F. Steglich (2013). "Microwave spectroscopy on heavy-fermion systems: Probing the dynamics of charges and magnetic moments". Physica Status Solidi B. 250 (3): 439–449. arXiv: 1303.5011 . Bibcode:2013PSSBR.250..439S. doi:10.1002/pssb.201200925. S2CID   59067473.
  10. M. Scheffler; M. Dressel; M. Jourdan; H. Adrian (2005). "Extremely slow Drude relaxation of correlated electrons". Nature . 438 (7071): 1135–1137. Bibcode:2005Natur.438.1135S. doi:10.1038/nature04232. PMID   16372004. S2CID   4391917.
  11. S. Donovan; A. Schwartz; G. Grüner (1997). "Observation of an Optical Pseudogap in UPt3". Physical Review Letters. 79 (7): 1401–1404. Bibcode:1997PhRvL..79.1401D. doi:10.1103/PhysRevLett.79.1401.
  12. Hilbert v. Löhneysen; et al. (2007). "Fermi-liquid instabilities at magnetic quantum phase transitions". Reviews of Modern Physics. 79 (3): 1015–1075. arXiv: cond-mat/0606317 . Bibcode:2007RvMP...79.1015L. doi:10.1103/RevModPhys.79.1015. S2CID   119512333.
  13. H.v. Löhneysen; et al. (1994). "Non-Fermi-liquid behavior in a heavy-fermion alloy at a magnetic instability". Physical Review Letters. 72 (20): 3262–3265. Bibcode:1994PhRvL..72.3262L. doi:10.1103/PhysRevLett.72.3262. PMID   10056148.
  14. J. Paglione; et al. (2003). "Field-Induced Quantum Critical Point in CeCoIn5". Physical Review Letters. 91 (24): 246405. arXiv: cond-mat/0212502 . Bibcode:2003PhRvL..91x6405P. doi:10.1103/PhysRevLett.91.246405. PMID   14683139. S2CID   15129138.
  15. A. Bianchi; et al. (2003). "Avoided Antiferromagnetic Order and Quantum Critical Point in CeCoIn5". Physical Review Letters. 91 (25): 257001. arXiv: cond-mat/0302226 . Bibcode:2003PhRvL..91y7001B. doi:10.1103/PhysRevLett.91.257001. PMID   14754138. S2CID   7562124.

Further reading