Van Vleck paramagnetism

Last updated

In condensed matter and atomic physics, Van Vleck paramagnetism refers to a positive and temperature-independent contribution to the magnetic susceptibility of a material, derived from second order corrections to the Zeeman interaction. The quantum mechanical theory was developed by John Hasbrouck Van Vleck between the 1920s and the 1930s to explain the magnetic response of gaseous nitric oxide (NO) and of rare-earth salts. [1] [2] [3] [4] Alongside other magnetic effects like Paul Langevin's formulas for paramagnetism (Curie's law) and diamagnetism, Van Vleck discovered an additional paramagnetic contribution of the same order as Langevin's diamagnetism. Van Vleck contribution is usually important for systems with one electron short of being half filled and this contribution vanishes for elements with closed shells. [5] [6]

Contents

Description

The magnetization of a material under an external small magnetic field is approximately described by

where is the magnetic susceptibility. When a magnetic field is applied to a paramagnetic material, its magnetization is parallel to the magnetic field and . For a diamagnetic material, the magnetization opposes the field, and .

Experimental measurements show that most non-magnetic materials have a susceptibility that behaves in the following way:

,

where is the absolute temperature; are constant, and , while can be positive, negative or null. Van Vleck paramagnetism often refers to systems where and .

Derivation

The Hamiltonian for an electron in a static homogeneous magnetic field in an atom is usually composed of three terms

where is the vacuum permeability, is the Bohr magneton, is the g-factor, is the elementary charge, is the electron mass, is the orbital angular momentum operator, the spin and is the component of the position operator orthogonal to the magnetic field. The Hamiltonian has three terms, the first one is the unperturbed Hamiltonian without the magnetic field, the second one is proportional to , and the third one is proportional to . In order to obtain the ground state of the system, one can treat exactly, and treat the magnetic field dependent terms using perturbation theory. Note that for strong magnetic fields, Paschen-Back effect dominates.

First order perturbation theory

First order perturbation theory on the second term of the Hamiltonian (proportional to ) for electrons bound to an atom, gives a positive correction to energy given by

where is the ground state, is the Landé g-factor of the ground state and is the total angular momentum operator (see Wigner–Eckart theorem). This correction leads to what is known as Langevin paramagnetism (the quantum theory is sometimes called Brillouin paramagnetism), that leads to a positive magnetic susceptibility. For sufficiently large temperatures, this contribution is described by Curie's law:

,

a susceptibility that is inversely proportional to the temperature , where is the material dependent Curie constant. If the ground state has no total angular momentum there is no Curie contribution and other terms dominate.

The first perturbation theory on the third term of the Hamiltonian (proportional to ), leads to a negative response (magnetization that opposes the magnetic field). Usually known as Larmor or Langenvin diamagnetism:

where is another constant proportional to the number of atoms per unit volume, and is the mean squared radius of the atom. Note that Larmor susceptibility does not depend on the temperature.

Second order: Van Vleck susceptibility

While Curie and Larmor susceptibilities were well understood from experimental measurements, J.H. Van Vleck noticed that the calculation above was incomplete. If is taken as the perturbation parameter, the calculation must include all orders of perturbation up to the same power of . As Larmor diamagnetism comes from first order perturbation of the , one must calculate second order perturbation of the term:

where the sum goes over all excited degenerate states , and are the energies of the excited states and the ground state, respectively, the sum excludes the state , where . Historically, J.H. Van Vleck called this term the "high frequency matrix elements". [4]

In this way, Van Vleck susceptibility comes from the second order energy correction, and can be written as

where is the number density, and and are the projection of the spin and orbital angular momentum in the direction of the magnetic field, respectively.

In this way, , as the signs of Larmor and Van Vleck susceptibilities are opposite, the sign of depends on the specific properties of the material.

General formula and Van Vleck criteria

For a more general system (molecules, complex systems), the paramagnetic susceptibility for an ensemble of independent magnetic moments can be written as

where

,
,

and is the Landé g-factor of state i. Van Vleck summarizes the results of this formula in four cases, depending on the temperature: [3]

  1. if all , where is Boltzmann constant, the susceptibility follows Curie law: ;
  2. if all , the susceptibility is independent of the temperature;
  3. if all is either or , the susceptibility has a mixed behavior and where is a constant;
  4. if all , there is no simple dependence on .

While molecular oxygen O
2
and nitric oxide NO are similar paramagnetic gases, O
2
follows Curie law as in case (a), while NO, deviates slightly from it. In 1927, Van Vleck considered NO to be in case (d) and obtained a more precise prediction of its susceptibility using the formula above. [2] [4]

Systems of interest

The standard example of Van Vleck paramagnetism are europium(III) oxide (Eu
2
O
3
) salts where there are six 4f electrons in trivalent europium ions. The ground state of Eu3+
that has a total azimuthal quantum number and Curie's contribution () vanishes, the first excited state with is very close to the ground state at 330 K and contributes through second order corrections as showed by Van Vleck. A similar effect is observed in samarium salts (Sm3+
ions). [7] [6] In the actinides, Van Vleck paramagnetism is also important in Bk5+
and Cm4+
which have a localized 5f6 configuration. [7]

Related Research Articles

<span class="mw-page-title-main">Diamagnetism</span> Magnetic property of ordinary materials

Diamagnetism is the property of materials that are repelled by a magnetic field; an applied magnetic field creates an induced magnetic field in them in the opposite direction, causing a repulsive force. In contrast, paramagnetic and ferromagnetic materials are attracted by a magnetic field. Diamagnetism is a quantum mechanical effect that occurs in all materials; when it is the only contribution to the magnetism, the material is called diamagnetic. In paramagnetic and ferromagnetic substances, the weak diamagnetic force is overcome by the attractive force of magnetic dipoles in the material. The magnetic permeability of diamagnetic materials is less than the permeability of vacuum, μ0. In most materials, diamagnetism is a weak effect which can be detected only by sensitive laboratory instruments, but a superconductor acts as a strong diamagnet because it entirely expels any magnetic field from its interior.

<span class="mw-page-title-main">Paramagnetism</span> Weak, attractive magnetism possessed by most elements and some compounds

Paramagnetism is a form of magnetism whereby some materials are weakly attracted by an externally applied magnetic field, and form internal, induced magnetic fields in the direction of the applied magnetic field. In contrast with this behavior, diamagnetic materials are repelled by magnetic fields and form induced magnetic fields in the direction opposite to that of the applied magnetic field. Paramagnetic materials include most chemical elements and some compounds; they have a relative magnetic permeability slightly greater than 1 and hence are attracted to magnetic fields. The magnetic moment induced by the applied field is linear in the field strength and rather weak. It typically requires a sensitive analytical balance to detect the effect and modern measurements on paramagnetic materials are often conducted with a SQUID magnetometer.

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

Superparamagnetism is a form of magnetism which appears in small ferromagnetic or ferrimagnetic nanoparticles. In sufficiently small nanoparticles, magnetization can randomly flip direction under the influence of temperature. The typical time between two flips is called the Néel relaxation time. In the absence of an external magnetic field, when the time used to measure the magnetization of the nanoparticles is much longer than the Néel relaxation time, their magnetization appears to be in average zero; they are said to be in the superparamagnetic state. In this state, an external magnetic field is able to magnetize the nanoparticles, similarly to a paramagnet. However, their magnetic susceptibility is much larger than that of paramagnets.

In electromagnetism, the magnetic susceptibility is a measure of how much a material will become magnetized in an applied magnetic field. It is the ratio of magnetization M to the applied magnetizing field intensity H. This allows a simple classification, into two categories, of most materials' responses to an applied magnetic field: an alignment with the magnetic field, χ > 0, called paramagnetism, or an alignment against the field, χ < 0, called diamagnetism.

The fluctuation–dissipation theorem (FDT) or fluctuation–dissipation relation (FDR) is a powerful tool in statistical physics for predicting the behavior of systems that obey detailed balance. Given that a system obeys detailed balance, the theorem is a proof that thermodynamic fluctuations in a physical variable predict the response quantified by the admittance or impedance of the same physical variable, and vice versa. The fluctuation–dissipation theorem applies both to classical and quantum mechanical systems.

<span class="mw-page-title-main">Permeability (electromagnetism)</span> Ability of magnetization

In electromagnetism, permeability is the measure of magnetization produced in a material in response to an applied magnetic field. Permeability is typically represented by the (italicized) Greek letter μ. It is the ratio of the magnetic induction to the magnetizing field as a function of the field in a material. The term was coined by William Thomson, 1st Baron Kelvin in 1872, and used alongside permittivity by Oliver Heaviside in 1885. The reciprocal of permeability is magnetic reluctivity.

The classical XY model is a lattice model of statistical mechanics. In general, the XY model can be seen as a specialization of Stanley's n-vector model for n = 2.

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">Gaussian units</span> Variant of the centimetre–gram–second unit system

Gaussian units constitute a metric system of physical units. This system is the most common of the several electromagnetic unit systems based on cgs (centimetre–gram–second) units. It is also called the Gaussian unit system, Gaussian-cgs units, or often just cgs units. The term "cgs units" is ambiguous and therefore to be avoided if possible: there are several variants of cgs with conflicting definitions of electromagnetic quantities and units.

In physics, the gyromagnetic ratio of a particle or system is the ratio of its magnetic moment to its angular momentum, and it is often denoted by the symbol γ, gamma. Its SI unit is the radian per second per tesla (rad⋅s−1⋅T−1) or, equivalently, the coulomb per kilogram (C⋅kg−1).

In magnetism, the Curie–Weiss law describes the magnetic susceptibility χ of a ferromagnet in the paramagnetic region above the Curie temperature:

In quantum physics, the spin–orbit interaction is a relativistic interaction of a particle's spin with its motion inside a potential. A key example of this phenomenon is the spin–orbit interaction leading to shifts in an electron's atomic energy levels, due to electromagnetic interaction between the electron's magnetic dipole, its orbital motion, and the electrostatic field of the positively charged nucleus. This phenomenon is detectable as a splitting of spectral lines, which can be thought of as a Zeeman effect product of two relativistic effects: the apparent magnetic field seen from the electron perspective and the magnetic moment of the electron associated with its intrinsic spin. A similar effect, due to the relationship between angular momentum and the strong nuclear force, occurs for protons and neutrons moving inside the nucleus, leading to a shift in their energy levels in the nucleus shell model. In the field of spintronics, spin–orbit effects for electrons in semiconductors and other materials are explored for technological applications. The spin–orbit interaction is at the origin of magnetocrystalline anisotropy and the spin Hall effect.

<span class="mw-page-title-main">Magnetization</span> Physical quantity, density of magnetic moment per volume

In classical electromagnetism, magnetization is the vector field that expresses the density of permanent or induced magnetic dipole moments in a magnetic material. Accordingly, physicists and engineers usually define magnetization as the quantity of magnetic moment per unit volume. It is represented by a pseudovector M. Magnetization can be compared to electric polarization, which is the measure of the corresponding response of a material to an electric field in electrostatics.

In quantum mechanics, the position operator is the operator that corresponds to the position observable of a particle.

In quantum mechanics, the Pauli equation or Schrödinger–Pauli equation is the formulation of the Schrödinger equation for spin-½ particles, which takes into account the interaction of the particle's spin with an external electromagnetic field. It is the non-relativistic limit of the Dirac equation and can be used where particles are moving at speeds much less than the speed of light, so that relativistic effects can be neglected. It was formulated by Wolfgang Pauli in 1927. In its linearized form it is known as Lévy-Leblond equation.

The Bohr–Van Leeuwen theorem states that when statistical mechanics and classical mechanics are applied consistently, the thermal average of the magnetization is always zero. This makes magnetism in solids solely a quantum mechanical effect and means that classical physics cannot account for paramagnetism, diamagnetism and ferromagnetism. Inability of classical physics to explain triboelectricity also stems from the Bohr–Van Leeuwen theorem.

For many paramagnetic materials, the magnetization of the material is directly proportional to an applied magnetic field, for sufficiently high temperatures and small fields. However, if the material is heated, this proportionality is reduced. For a fixed value of the field, the magnetic susceptibility is inversely proportional to temperature, that is

Magnetochemistry is concerned with the magnetic properties of chemical compounds. Magnetic properties arise from the spin and orbital angular momentum of the electrons contained in a compound. Compounds are diamagnetic when they contain no unpaired electrons. Molecular compounds that contain one or more unpaired electrons are paramagnetic. The magnitude of the paramagnetism is expressed as an effective magnetic moment, μeff. For first-row transition metals the magnitude of μeff is, to a first approximation, a simple function of the number of unpaired electrons, the spin-only formula. In general, spin–orbit coupling causes μeff to deviate from the spin-only formula. For the heavier transition metals, lanthanides and actinides, spin–orbit coupling cannot be ignored. Exchange interaction can occur in clusters and infinite lattices, resulting in ferromagnetism, antiferromagnetism or ferrimagnetism depending on the relative orientations of the individual spins.

<span class="mw-page-title-main">Weyl equation</span> Relativistic wave equation describing massless fermions

In physics, particularly in quantum field theory, the Weyl equation is a relativistic wave equation for describing massless spin-1/2 particles called Weyl fermions. The equation is named after Hermann Weyl. The Weyl fermions are one of the three possible types of elementary fermions, the other two being the Dirac and the Majorana fermions.

The optical metric was defined by German theoretical physicist Walter Gordon in 1923 to study the geometrical optics in curved space-time filled with moving dielectric materials.

References

  1. Van Vleck, John Hasbrouck (1932). The Theory of Electric and Magnetic Susceptibilities. Clarendon Press.
  2. 1 2 Van Vleck, J. H. (1928-04-01). "On Dielectric Constants and Magnetic Susceptibilities in the New Quantum Mechanics Part III—Application to Dia- and Paramagnetism". Physical Review. 31 (4): 587–613. Bibcode:1928PhRv...31..587V. doi:10.1103/PhysRev.31.587. ISSN   0031-899X.
  3. 1 2 van Vleck, John H. (1977). "John H. van Vleck Nobel Lecture". Nobel Prize. Retrieved 2020-10-18.
  4. 1 2 3 Anderson, Philip W. (1987). John Hasbrouck Van Vleck (PDF). Washington D.C: National Academy of Sciences.
  5. Marder, Michael P. (2010-11-17). Condensed Matter Physics. John Wiley & Sons. ISBN   978-0-470-94994-8.
  6. 1 2 Nolting, Wolfgang; Ramakanth, Anupuru (2009-10-03). Quantum Theory of Magnetism. Springer Science & Business Media. ISBN   978-3-540-85416-6.
  7. 1 2 Coey, J. M. D. (2010). Magnetism and Magnetic Materials. Cambridge: Cambridge University Press. ISBN   978-0-521-81614-4.