Exchange interaction

Last updated

In chemistry and physics, the exchange interaction is a quantum mechanical constraint on the states of indistinguishable particles. While sometimes called an exchange force, or, in the case of fermions, Pauli repulsion, its consequences cannot always be predicted based on classical ideas of force. [1] Both bosons and fermions can experience the exchange interaction.

Contents

The wave function of indistinguishable particles is subject to exchange symmetry: the wave function either changes sign (for fermions) or remains unchanged (for bosons) when two particles are exchanged. The exchange symmetry alters the expectation value of the distance between two indistinguishable particles when their wave functions overlap. For fermions the expectation value of the distance increases, and for bosons it decreases (compared to distinguishable particles). [2]

The exchange interaction arises from the combination of exchange symmetry and the Coulomb interaction. For an electron in an electron gas, the exchange symmetry creates an "exchange hole" in its vicinity, which other electrons with the same spin tend to avoid due to the Pauli exclusion principle. This decreases the energy associated with the Coulomb interactions between the electrons with same spin. [3] Since two electrons with different spins are distinguishable from each other and not subject to the exchange symmetry, the effect tends to align the spins. Exchange interaction is the main physical effect responsible for ferromagnetism, and has no classical analogue.

For bosons, the exchange symmetry makes them bunch together, and the exchange interaction takes the form of an effective attraction that causes identical particles to be found closer together, as in Bose–Einstein condensation.

Exchange interaction effects were discovered independently by physicists Werner Heisenberg and Paul Dirac in 1926. [4] [5]

Exchange symmetry

Quantum particles are fundamentally indistinguishable. Wolfgang Pauli demonstrated that this is a type of symmetry: states of two particles must be either symmetric or antisymmetric when coordinate labels are exchanged. [6] In a simple one-dimensional system with two identical particles in two states and the system wavefunction can therefore be written two ways:

Exchanging and gives either a symmetric combination of the states ('plus') or an antisymmetric combination ('minus'). Particles that give symmetric combinations are called bosons; those with antisymmetric combinations are called fermions.

The two possible combinations imply different physics. For example the expectation value of the square of the distance between the two particles is: [7] :258

The last term reduces the expected value for bosons and increases the value for fermions but only when the states and physically overlap ().

The physical effect of the exchange symmetry requirement is not a force. Rather it is a significant geometrical constraint, increasing the curvature of wavefunctions to prevent the overlap of the states occupied by indistinguishable fermions. The terms "exchange force" and "Pauli repulsion" for fermions are sometimes used as an intuitive description of the effect but this intuition can give incorrect physical results. [1] [7] :291

Exchange interactions between localized electron magnetic moments

Quantum mechanical particles are classified as bosons or fermions. The spin–statistics theorem of quantum field theory demands that all particles with half-integer spin behave as fermions and all particles with integer spin behave as bosons. Multiple bosons may occupy the same quantum state; however, by the Pauli exclusion principle, no two fermions can occupy the same state. Since electrons have spin 1/2, they are fermions. This means that the overall wave function of a system must be antisymmetric when two electrons are exchanged, i.e. interchanged with respect to both spatial and spin coordinates. First, however, exchange will be explained with the neglect of spin.

Exchange of spatial coordinates

Taking a hydrogen molecule-like system (i.e. one with two electrons), one may attempt to model the state of each electron by first assuming the electrons behave independently, and taking wave functions in position space of for the first electron and for the second electron. The functions and are orthogonal, and each corresponds to an energy eigenstate. Two wave functions for the overall system in position space can be constructed. One uses an antisymmetric combination of the product wave functions in position space:

(1)

The other uses a symmetric combination of the product wave functions in position space:

(2)

To treat the problem of the Hydrogen molecule perturbatively, the overall Hamiltonian is decomposed into a unperturbed Hamiltonian of the non-interacting hydrogen atoms and a perturbing Hamiltonian, which accounts for interactions between the two atoms . The full Hamiltonian is then:

where and

The first two terms of denote the kinetic energy of the electrons. The remaining terms account for attraction between the electrons and their host protons (ra1/b2). The terms in account for the potential energy corresponding to: proton–proton repulsion (Rab), electron–electron repulsion (r12), and electron–proton attraction between the electron of one host atom and the proton of the other (ra2/b1). All quantities are assumed to be real.

Two eigenvalues for the system energy are found:

(3)

where the E+ is the spatially symmetric solution and E is the spatially antisymmetric solution, corresponding to and respectively. A variational calculation yields similar results. can be diagonalized by using the position–space functions given by Eqs. (1) and (2). In Eq. (3), C is the two-site two-electron Coulomb integral (It may be interpreted as the repulsive potential for electron-one at a particular point in an electric field created by electron-two distributed over the space with the probability density , [lower-alpha 1] is the overlap integral, and Jex is the exchange integral, which is similar to the two-site Coulomb integral but includes exchange of the two electrons. It has no simple physical interpretation, but it can be shown to arise entirely due to the anti-symmetry requirement. These integrals are given by:

(4)

(5)

(6)

Although in the hydrogen molecule the exchange integral, Eq. (6), is negative, Heisenberg first suggested that it changes sign at some critical ratio of internuclear distance to mean radial extension of the atomic orbital. [8] [9] [10]

Inclusion of spin

The symmetric and antisymmetric combinations in Equations (1) and (2) did not include the spin variables (α = spin-up; β = spin-down); there are also antisymmetric and symmetric combinations of the spin variables:

(7)

To obtain the overall wave function, these spin combinations have to be coupled with Eqs. (1) and (2). The resulting overall wave functions, called spin-orbitals, are written as Slater determinants. When the orbital wave function is symmetrical the spin one must be anti-symmetrical and vice versa. Accordingly, E+ above corresponds to the spatially symmetric/spin-singlet solution and E to the spatially antisymmetric/spin-triplet solution.

J. H. Van Vleck presented the following analysis: [11]

The potential energy of the interaction between the two electrons in orthogonal orbitals can be represented by a matrix,sayEex. From Eq. (3), the characteristic values of this matrix areC ± Jex. The characteristic values of a matrix are its diagonal elements after it is converted to a diagonal matrix. Now, the characteristic values of the square of the magnitude of the resultant spin, is . The characteristic values of the matricesandare eachand. The characteristic values of the scalar productareand, corresponding to both the spin-singlet (S = 0) and spin-triplet (S = 1) states, respectively.
From Eq. (3) and the aforementioned relations, the matrix Eexis seen to have the characteristic valueC + Jexwhenhas the characteristic value −3/4 (i.e. when S = 0; the spatially symmetric/spin-singlet state). Alternatively, it has the characteristic value C − Jexwhen has the characteristic value +1/4 (i.e. when S = 1; the spatially antisymmetric/spin-triplet state). Therefore,

(8)
and, hence,

(9)
where the spin momenta are given asand.

Dirac pointed out that the critical features of the exchange interaction could be obtained in an elementary way by neglecting the first two terms on the right-hand side of Eq. (9), thereby considering the two electrons as simply having their spins coupled by a potential of the form:

(10)

It follows that the exchange interaction Hamiltonian between two electrons in orbitals Φa and Φb can be written in terms of their spin momenta and . This interaction is named the Heisenberg exchange Hamiltonian or the Heisenberg–Dirac Hamiltonian in the older literature:

(11)

Jab is not the same as the quantity labeled Jex in Eq. (6). Rather, Jab, which is termed the exchange constant, is a function of Eqs. (4), (5), and (6), namely,

(12)

However, with orthogonal orbitals (in which = 0), for example with different orbitals in the same atom, Jab = Jex.

Effects of exchange

If Jab is positive the exchange energy favors electrons with parallel spins; this is a primary cause of ferromagnetism in materials in which the electrons are considered localized in the Heitler–London model of chemical bonding, but this model of ferromagnetism has severe limitations in solids (see below). If Jab is negative, the interaction favors electrons with antiparallel spins, potentially causing antiferromagnetism. The sign of Jab is essentially determined by the relative sizes of Jex and the product of . This sign can be deduced from the expression for the difference between the energies of the triplet and singlet states, EE+:

(13)

Although these consequences of the exchange interaction are magnetic in nature, the cause is not; it is due primarily to electric repulsion and the Pauli exclusion principle. In general, the direct magnetic interaction between a pair of electrons (due to their electron magnetic moments) is negligibly small compared to this electric interaction.

Exchange energy splittings are very elusive to calculate for molecular systems at large internuclear distances. However, analytical formulae have been worked out for the hydrogen molecular ion (see references herein).

Normally, exchange interactions are very short-ranged, confined to electrons in orbitals on the same atom (intra-atomic exchange) or nearest neighbor atoms (direct exchange) but longer-ranged interactions can occur via intermediary atoms and this is termed superexchange.

Direct exchange interactions in solids

In a crystal, generalization of the Heisenberg Hamiltonian in which the sum is taken over the exchange Hamiltonians for all the (i,j) pairs of atoms of the many-electron system gives:.

(14)

The 1/2 factor is introduced because the interaction between the same two atoms is counted twice in performing the sums. Note that the J in Eq.(14) is the exchange constant Jab above not the exchange integral Jex. The exchange integral Jex is related to yet another quantity, called the exchange stiffness constant (A) which serves as a characteristic of a ferromagnetic material. The relationship is dependent on the crystal structure. For a simple cubic lattice with lattice parameter ,

(15)

For a body-centered cubic lattice,

(16)

and for a face-centered cubic lattice,

(17)

The form of Eq. (14) corresponds identically to the Ising model of ferromagnetism except that in the Ising model, the dot product of the two spin angular momenta is replaced by the scalar product SijSji. The Ising model was invented by Wilhelm Lenz in 1920 and solved for the one-dimensional case by his doctoral student Ernst Ising in 1925. The energy of the Ising model is defined to be:

(18)

Limitations of the Heisenberg Hamiltonian and the localized electron model in solids

Because the Heisenberg Hamiltonian presumes the electrons involved in the exchange coupling are localized in the context of the Heitler–London, or valence bond (VB), theory of chemical bonding, it is an adequate model for explaining the magnetic properties of electrically insulating narrow-band ionic and covalent non-molecular solids where this picture of the bonding is reasonable. Nevertheless, theoretical evaluations of the exchange integral for non-molecular solids that display metallic conductivity in which the electrons responsible for the ferromagnetism are itinerant (e.g. iron, nickel, and cobalt) have historically been either of the wrong sign or much too small in magnitude to account for the experimentally determined exchange constant (e.g. as estimated from the Curie temperatures via TC ≈ 2⟨J⟩/3kB where ⟨J⟩ is the exchange interaction averaged over all sites).

The Heisenberg model thus cannot explain the observed ferromagnetism in these materials. [12] In these cases, a delocalized, or Hund–Mulliken–Bloch (molecular orbital/band) description, for the electron wave functions is more realistic. Accordingly, the Stoner model of ferromagnetism is more applicable.

In the Stoner model, the spin-only magnetic moment (in Bohr magnetons) per atom in a ferromagnet is given by the difference between the number of electrons per atom in the majority spin and minority spin states. The Stoner model thus permits non-integral values for the spin-only magnetic moment per atom. However, with ferromagnets (g = 2.0023 ≈ 2) tends to overestimate the total spin-only magnetic moment per atom.

For example, a net magnetic moment of 0.54 μB per atom for Nickel metal is predicted by the Stoner model, which is very close to the 0.61 Bohr magnetons calculated based on the metal's observed saturation magnetic induction, its density, and its atomic weight. [13] By contrast, an isolated Ni atom (electron configuration = 3d84s2) in a cubic crystal field will have two unpaired electrons of the same spin (hence, ) and would thus be expected to have in the localized electron model a total spin magnetic moment of (but the measured spin-only magnetic moment along one axis, the physical observable, will be given by ).

Generally, valence s and p electrons are best considered delocalized, while 4f electrons are localized and 5f and 3d/4d electrons are intermediate, depending on the particular internuclear distances. [14] In the case of substances where both delocalized and localized electrons contribute to the magnetic properties (e.g. rare-earth systems), the Ruderman–Kittel–Kasuya–Yosida (RKKY) model is the currently accepted mechanism.

See also

Notes

  1. Not to be confused with the total spin, .

Related Research Articles

<span class="mw-page-title-main">Antiparticle</span> Particle with opposite charges

In particle physics, every type of particle of "ordinary" matter is associated with an antiparticle with the same mass but with opposite physical charges. For example, the antiparticle of the electron is the positron. While the electron has a negative electric charge, the positron has a positive electric charge, and is produced naturally in certain types of radioactive decay. The opposite is also true: the antiparticle of the positron is the electron.

In quantum mechanics, indistinguishable particles are particles that cannot be distinguished from one another, even in principle. Species of identical particles include, but are not limited to, elementary particles, composite subatomic particles, as well as atoms and molecules. Quasiparticles also behave in this way. Although all known indistinguishable particles only exist at the quantum scale, there is no exhaustive list of all possible sorts of particles nor a clear-cut limit of applicability, as explored in quantum statistics. They were first discussed by Werner Heisenberg and Paul Dirac in 1926.

The mathematical formulations of quantum mechanics are those mathematical formalisms that permit a rigorous description of quantum mechanics. This mathematical formalism uses mainly a part of functional analysis, especially Hilbert spaces, which are a kind of linear space. Such are distinguished from mathematical formalisms for physics theories developed prior to the early 1900s by the use of abstract mathematical structures, such as infinite-dimensional Hilbert spaces, and operators on these spaces. In brief, values of physical observables such as energy and momentum were no longer considered as values of functions on phase space, but as eigenvalues; more precisely as spectral values of linear operators in Hilbert space.

<span class="mw-page-title-main">Pauli exclusion principle</span> Quantum mechanics rule: identical fermions cannot occupy the same quantum state simultaneously

In quantum mechanics, the Pauli exclusion principle states that two or more identical particles with half-integer spins cannot simultaneously occupy the same quantum state within a system that obeys the laws of quantum mechanics. This principle was formulated by Austrian physicist Wolfgang Pauli in 1925 for electrons, and later extended to all fermions with his spin–statistics theorem of 1940.

<span class="mw-page-title-main">Zeeman effect</span> Spectral line splitting in magnetic field

The Zeeman effect is the effect of splitting of a spectral line into several components in the presence of a static magnetic field. It is named after the Dutch physicist Pieter Zeeman, who discovered it in 1896 and received a Nobel prize for this discovery. It is analogous to the Stark effect, the splitting of a spectral line into several components in the presence of an electric field. Also similar to the Stark effect, transitions between different components have, in general, different intensities, with some being entirely forbidden, as governed by the selection rules.

In quantum mechanics, a Fock state or number state is a quantum state that is an element of a Fock space with a well-defined number of particles. These states are named after the Soviet physicist Vladimir Fock. Fock states play an important role in the second quantization formulation of quantum mechanics.

In quantum mechanics, a Slater determinant is an expression that describes the wave function of a multi-fermionic system. It satisfies anti-symmetry requirements, and consequently the Pauli principle, by changing sign upon exchange of two electrons. Only a small subset of all possible fermionic wave functions can be written as a single Slater determinant, but those form an important and useful subset because of their simplicity.

<span class="mw-page-title-main">Fine structure</span> Details in the emission spectrum of an atom

In atomic physics, the fine structure describes the splitting of the spectral lines of atoms due to electron spin and relativistic corrections to the non-relativistic Schrödinger equation. It was first measured precisely for the hydrogen atom by Albert A. Michelson and Edward W. Morley in 1887, laying the basis for the theoretical treatment by Arnold Sommerfeld, introducing the fine-structure constant.

In physics, a parity transformation is the flip in the sign of one spatial coordinate. In three dimensions, it can also refer to the simultaneous flip in the sign of all three spatial coordinates :

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

<span class="mw-page-title-main">Orbital motion (quantum)</span> Quantum mechanical property

Quantum orbital motion involves the quantum mechanical motion of rigid particles about some other mass, or about themselves. In classical mechanics, an object's orbital motion is characterized by its orbital angular momentum and spin angular momentum, which is the object's angular momentum about its own center of mass. In quantum mechanics there are analogous orbital and spin angular momenta which describe the orbital motion of a particle, represented as quantum mechanical operators instead of vectors.

In quantum field theory, a fermionic field is a quantum field whose quanta are fermions; that is, they obey Fermi–Dirac statistics. Fermionic fields obey canonical anticommutation relations rather than the canonical commutation relations of bosonic fields.

<span class="mw-page-title-main">Degenerate energy levels</span> Energy level of a quantum system that corresponds to two or more different measurable states

In quantum mechanics, an energy level is degenerate if it corresponds to two or more different measurable states of a quantum system. Conversely, two or more different states of a quantum mechanical system are said to be degenerate if they give the same value of energy upon measurement. The number of different states corresponding to a particular energy level is known as the degree of degeneracy of the level. It is represented mathematically by the Hamiltonian for the system having more than one linearly independent eigenstate with the same energy eigenvalue. When this is the case, energy alone is not enough to characterize what state the system is in, and other quantum numbers are needed to characterize the exact state when distinction is desired. In classical mechanics, this can be understood in terms of different possible trajectories corresponding to the same energy.

In quantum mechanics, the exchange operator, also known as permutation operator, is a quantum mechanical operator that acts on states in Fock space. The exchange operator acts by switching the labels on any two identical particles described by the joint position quantum state . Since the particles are identical, the notion of exchange symmetry requires that the exchange operator be unitary.

The Bose–Hubbard model gives a description of the physics of interacting spinless bosons on a lattice. It is closely related to the Hubbard model that originated in solid-state physics as an approximate description of superconducting systems and the motion of electrons between the atoms of a crystalline solid. The model was introduced by Gersch and Knollman in 1963 in the context of granular superconductors. The model rose to prominence in the 1980s after it was found to capture the essence of the superfluid-insulator transition in a way that was much more mathematically tractable than fermionic metal-insulator models.

In quantum mechanics, a raising or lowering operator is an operator that increases or decreases the eigenvalue of another operator. In quantum mechanics, the raising operator is sometimes called the creation operator, and the lowering operator the annihilation operator. Well-known applications of ladder operators in quantum mechanics are in the formalisms of the quantum harmonic oscillator and angular momentum.

In quantum mechanics and quantum field theory, a Schrödinger field, named after Erwin Schrödinger, is a quantum field which obeys the Schrödinger equation. While any situation described by a Schrödinger field can also be described by a many-body Schrödinger equation for identical particles, the field theory is more suitable for situations where the particle number changes.

Spin is an intrinsic form of angular momentum carried by elementary particles, and thus by composite particles such as hadrons, atomic nuclei, and atoms. Spin is quantized, and accurate models for the interaction with spin require relativistic quantum mechanics or quantum field theory.

<span class="mw-page-title-main">Helium atom</span> Atom of helium

A helium atom is an atom of the chemical element helium. Helium is composed of two electrons bound by the electromagnetic force to a nucleus containing two protons along with two neutrons, depending on the isotope, held together by the strong force. Unlike for hydrogen, a closed-form solution to the Schrödinger equation for the helium atom has not been found. However, various approximations, such as the Hartree–Fock method, can be used to estimate the ground state energy and wavefunction of the atom. Historically, the first such helium spectrum calculation was done by Albrecht Unsöld in 1927. Its success was considered to be one of the earliest signs of validity of Schrödinger's wave mechanics.

This is a glossary for the terminology often encountered in undergraduate quantum mechanics courses.

References

  1. 1 2 Mullin, W. J.; Blaylock, G. (2003-11-11). "Quantum statistics: Is there an effective fermion repulsion or boson attraction?". American Journal of Physics. 71 (12): 1223–1231. arXiv: physics/0304067 . doi:10.1119/1.1590658. ISSN   0002-9505.
  2. David J. Griffiths: Introduction to Quantum Mechanics, Second Edition, pp. 207–210
  3. Girvin, Steven M.; Yang, Kun (2019). Modern condensed matter physics. Cambridge New York: Cambridge university press. p. 384. ISBN   978-1-107-13739-4.
  4. Mehrkörperproblem und Resonanz in der Quantenmechanik, W. Heisenberg, Zeitschrift für Physik38, #6–7 (June 1926), pp. 411–426. DOI 10.1007/BF01397160.
  5. Dirac, P. A. M. (1926-10-01). "On the Theory of Quantum Mechanics". Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences. 112 (762). The Royal Society: 661–677. Bibcode:1926RSPSA.112..661D. doi: 10.1098/rspa.1926.0133 . ISSN   1364-5021. JSTOR   94692.
  6. Blum, Alexander (2014-12-01). "From the necessary to the possible: the genesis of the spin-statistics theorem". The European Physical Journal H. 39 (5): 543–574. doi:10.1140/epjh/e2014-50022-5. ISSN   2102-6467.
  7. 1 2 Griffiths, David J.; Schroeter, Darrell F. (2018-08-16). Introduction to Quantum Mechanics. doi:10.1017/9781316995433.
  8. Derivation of the Heisenberg Hamiltonian Archived 2021-10-21 at the Wayback Machine , Rebecca Hihinashvili, accessed on line October 2, 2007.
  9. Quantum Theory of Magnetism: Magnetic Properties of Materials, Robert M. White, 3rd rev. ed., Berlin: Springer-Verlag, 2007, section 2.2.7. ISBN   3-540-65116-0.
  10. The Theory of Electric and Magnetic Susceptibilities, J. H. van Vleck, London: Oxford University Press, 1932, chapter XII, section 76.
  11. Van Vleck, J. H.: Electric and Magnetic Susceptibilities, Oxford, Clarendon Press, p. 318 (1932).
  12. Stuart, R.; Marshall, W. (1960-10-15). "Direct Exchange in Ferromagnets". Physical Review. 120 (2). American Physical Society (APS): 353–357. Bibcode:1960PhRv..120..353S. doi:10.1103/physrev.120.353. ISSN   0031-899X.
  13. Elliot, S. R.: The Physics and Chemistry of Solids, John Wiley & Sons, New York, p. 615 (1998)
  14. J. B. Goodenough: Magnetism and the Chemical Bond, Interscience Publishers, New York, pp. 5–17 (1966).

Further reading