Hund's cases

Last updated

In rotational-vibrational and electronic spectroscopy of diatomic molecules, Hund's coupling cases are idealized descriptions of rotational states in which specific terms in the molecular Hamiltonian and involving couplings between angular momenta are assumed to dominate over all other terms. There are five cases, proposed by Friedrich Hund in 1926-27 [1] and traditionally denoted by the letters (a) through (e). Most diatomic molecules are somewhere between the idealized cases (a) and (b). [2]

Contents

Angular momenta

To describe the Hund's coupling cases, we use the following angular momenta (where boldface letters indicate vector quantities):

These vector quantities depend on corresponding quantum numbers whose values are shown in molecular term symbols used to identify the states. For example, the term symbol 2Π3/2 denotes a state with S = 1/2, Λ = 1 and J = 3/2.

Choosing the applicable Hund's case

Hund's coupling cases are idealizations. The appropriate case for a given situation can be found by comparing three strengths: the electrostatic coupling of to the internuclear axis, the spin-orbit coupling, and the rotational coupling of and to the total angular momentum .

For 1Σ states the orbital and spin angular momenta are zero and the total angular momentum is just the nuclear rotational angular momentum. [3] For other states, Hund proposed five possible idealized modes of coupling. [4]

Hund's caseElectrostaticSpin-orbitRotational
(a)strongintermediateweak
(b)strongweakintermediate
(c)intermediatestrongweak
(d)intermediateweakstrong
(e)weakintermediatestrong
strongintermediate

The last two rows are degenerate because they have the same good quantum numbers. [5]

In practice there are also many molecular states which are intermediate between the above limiting cases. [3]

Case (a)

The most common [6] case is case (a) in which is electrostatically coupled to the internuclear axis, and is coupled to by spin-orbit coupling. Then both and have well-defined axial components, and respectively. As they are written with the same Greek symbol, the spin component should not be confused with states, which are states with orbital angular component equal to zero. defines a vector of magnitude pointing along the internuclear axis. Combined with the rotational angular momentum of the nuclei , we have . In this case, the precession of and around the nuclear axis is assumed to be much faster than the nutation of and around .

The good quantum numbers in case (a) are , , , and . However is not a good quantum number because the vector is strongly coupled to the electrostatic field and therefore precesses rapidly around the internuclear axis with an undefined magnitude. [6] We express the rotational energy operator as , where is a rotational constant. There are, ideally, fine-structure states, each with rotational levels having relative energies starting with . [2] For example, a 2Π state has a 2Π1/2 term (or fine structure state) with rotational levels = 1/2, 3/2, 5/2, 7/2, ... and a 2Π3/2 term with levels = 3/2, 5/2, 7/2, 9/2.... [4] Case (a) requires > 0 and so does not apply to any Σ states, and also > 0 so that it does not apply to any singlet states. [7]

The selection rules for allowed spectroscopic transitions depend on which quantum numbers are good. For Hund's case (a), the allowed transitions must have and and and and . [8] In addition, symmetrical diatomic molecules have even (g) or odd (u) parity and obey the Laporte rule that only transitions between states of opposite parity are allowed.

Case (b)

In case (b), the spin-orbit coupling is weak or non-existent (in the case ). In this case, we take and and assume precesses quickly around the internuclear axis.

The good quantum numbers in case (b) are , , , and . We express the rotational energy operator as , where is a rotational constant. The rotational levels therefore have relative energies starting with . [2] For example, a 2Σ state has rotational levels = 0, 1, 2, 3, 4, ..., and each level is divided by spin-orbit coupling into two levels = ± 1/2 (except for = 0 which corresponds only to = 1/2 because cannot be negative). [9]

Another example is the 3Σ ground state of dioxygen, which has two unpaired electrons with parallel spins. The coupling type is Hund's case b), and each rotational level N is divided into three levels = , , . [10]

For case b) the selection rules for quantum numbers , , and and for parity are the same as for case a). However for the rotational levels, the rule for quantum number does not apply and is replaced by the rule . [11]

Case (c)

In case (c), the spin-orbit coupling is stronger than the coupling to the internuclear axis, and and from case (a) cannot be defined. Instead and combine to form , which has a projection along the internuclear axis of magnitude . Then , as in case (a).

The good quantum numbers in case (c) are , , and . [2] Since is undefined for this case, the states cannot be described as , or . [12] An example of Hund's case (c) is the lowest 3Πu state of diiodine (I2), which approximates more closely to case (c) than to case (a). [6]

The selection rules for , and parity are valid as for cases (a) and (b), but there are no rules for and since these are not good quantum numbers for case (c). [6]

Case (d)

In case (d), the rotational coupling between and is much stronger than the electrostatic coupling of to the internuclear axis. Thus we form by coupling and and the form by coupling and .

The good quantum numbers in case (d) are , , , , and . Because is a good quantum number, the rotational energy is simply . [2]

Case (e)

In case (e), we first form and then form by coupling and . This case is rare but has been observed. [13] Rydberg states which converge to ionic states with spin–orbit coupling (such as 2Π) are best described as case (e). [14]

The good quantum numbers in case (e) are , , and . Because is once again a good quantum number, the rotational energy is . [2]

Related Research Articles

<span class="mw-page-title-main">Angular momentum</span> Conserved physical quantity; rotational analogue of linear momentum

In physics, angular momentum is the rotational analog of linear momentum. It is an important physical quantity because it is a conserved quantity – the total angular momentum of a closed system remains constant. Angular momentum has both a direction and a magnitude, and both are conserved. Bicycles and motorcycles, frisbees, rifled bullets, and gyroscopes owe their useful properties to conservation of angular momentum. Conservation of angular momentum is also why hurricanes form spirals and neutron stars have high rotational rates. In general, conservation limits the possible motion of a system, but it does not uniquely determine it.

<span class="mw-page-title-main">Diatomic molecule</span> Molecule composed of any two atoms

Diatomic molecules are molecules composed of only two atoms, of the same or different chemical elements. If a diatomic molecule consists of two atoms of the same element, such as hydrogen or oxygen, then it is said to be homonuclear. Otherwise, if a diatomic molecule consists of two different atoms, such as carbon monoxide or nitric oxide, the molecule is said to be heteronuclear. The bond in a homonuclear diatomic molecule is non-polar.

In particle physics, the Dirac equation is a relativistic wave equation derived by British physicist Paul Dirac in 1928. In its free form, or including electromagnetic interactions, it describes all spin-12 massive particles, called "Dirac particles", such as electrons and quarks for which parity is a symmetry. It is consistent with both the principles of quantum mechanics and the theory of special relativity, and was the first theory to account fully for special relativity in the context of quantum mechanics. It was validated by accounting for the fine structure of the hydrogen spectrum in a completely rigorous way.

<span class="mw-page-title-main">Energy level</span> Different states of quantum systems

A quantum mechanical system or particle that is bound—that is, confined spatially—can only take on certain discrete values of energy, called energy levels. This contrasts with classical particles, which can have any amount of energy. The term is commonly used for the energy levels of the electrons in atoms, ions, or molecules, which are bound by the electric field of the nucleus, but can also refer to energy levels of nuclei or vibrational or rotational energy levels in molecules. The energy spectrum of a system with such discrete energy levels is said to be quantized.

<span class="mw-page-title-main">Morse potential</span> Model for the potential energy of a diatomic molecule

The Morse potential, named after physicist Philip M. Morse, is a convenient interatomic interaction model for the potential energy of a diatomic molecule. It is a better approximation for the vibrational structure of the molecule than the quantum harmonic oscillator because it explicitly includes the effects of bond breaking, such as the existence of unbound states. It also accounts for the anharmonicity of real bonds and the non-zero transition probability for overtone and combination bands. The Morse potential can also be used to model other interactions such as the interaction between an atom and a surface. Due to its simplicity, it is not used in modern spectroscopy. However, its mathematical form inspired the MLR (Morse/Long-range) potential, which is the most popular potential energy function used for fitting spectroscopic data.

In quantum mechanics, a rotational transition is an abrupt change in angular momentum. Like all other properties of a quantum particle, angular momentum is quantized, meaning it can only equal certain discrete values, which correspond to different rotational energy states. When a particle loses angular momentum, it is said to have transitioned to a lower rotational energy state. Likewise, when a particle gains angular momentum, a positive rotational transition is said to have occurred.

In rotordynamics, the rigid rotor is a mechanical model of rotating systems. An arbitrary rigid rotor is a 3-dimensional rigid object, such as a top. To orient such an object in space requires three angles, known as Euler angles. A special rigid rotor is the linear rotor requiring only two angles to describe, for example of a diatomic molecule. More general molecules are 3-dimensional, such as water, ammonia, or methane.

In physics and chemistry, a selection rule, or transition rule, formally constrains the possible transitions of a system from one quantum state to another. Selection rules have been derived for electromagnetic transitions in molecules, in atoms, in atomic nuclei, and so on. The selection rules may differ according to the technique used to observe the transition. The selection rule also plays a role in chemical reactions, where some are formally spin-forbidden reactions, that is, reactions where the spin state changes at least once from reactants to products.

In molecular physics, the molecular term symbol is a shorthand expression of the group representation and angular momenta that characterize the state of a molecule, i.e. its electronic quantum state which is an eigenstate of the electronic molecular Hamiltonian. It is the equivalent of the term symbol for the atomic case. However, the following presentation is restricted to the case of homonuclear diatomic molecules, or other symmetric molecules with an inversion centre. For heteronuclear diatomic molecules, the u/g symbol does not correspond to any exact symmetry of the electronic molecular Hamiltonian. In the case of less symmetric molecules the molecular term symbol contains the symbol of the group representation to which the molecular electronic state belongs.

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.

In physics, Larmor precession is the precession of the magnetic moment of an object about an external magnetic field. The phenomenon is conceptually similar to the precession of a tilted classical gyroscope in an external torque-exerting gravitational field. Objects with a magnetic moment also have angular momentum and effective internal electric current proportional to their angular momentum; these include electrons, protons, other fermions, many atomic and nuclear systems, as well as classical macroscopic systems. The external magnetic field exerts a torque on the magnetic moment,

<span class="mw-page-title-main">Jaynes–Cummings model</span> Model in quantum optics

The Jaynes–Cummings model is a theoretical model in quantum optics. It describes the system of a two-level atom interacting with a quantized mode of an optical cavity, with or without the presence of light. It was originally developed to study the interaction of atoms with the quantized electromagnetic field in order to investigate the phenomena of spontaneous emission and absorption of photons in a cavity.

The theoretical and experimental justification for the Schrödinger equation motivates the discovery of the Schrödinger equation, the equation that describes the dynamics of nonrelativistic particles. The motivation uses photons, which are relativistic particles with dynamics described by Maxwell's equations, as an analogue for all types of particles.

In mathematical physics, the Belinfante–Rosenfeld tensor is a modification of the energy–momentum tensor that is constructed from the canonical energy–momentum tensor and the spin current so as to be symmetric yet still conserved.

<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.

<span class="mw-page-title-main">Relativistic angular momentum</span> Angular momentum in special and general relativity

In physics, relativistic angular momentum refers to the mathematical formalisms and physical concepts that define angular momentum in special relativity (SR) and general relativity (GR). The relativistic quantity is subtly different from the three-dimensional quantity in classical mechanics.

<span class="mw-page-title-main">Symmetry in quantum mechanics</span> Properties underlying modern physics

Symmetries in quantum mechanics describe features of spacetime and particles which are unchanged under some transformation, in the context of quantum mechanics, relativistic quantum mechanics and quantum field theory, and with applications in the mathematical formulation of the standard model and condensed matter physics. In general, symmetry in physics, invariance, and conservation laws, are fundamentally important constraints for formulating physical theories and models. In practice, they are powerful methods for solving problems and predicting what can happen. While conservation laws do not always give the answer to the problem directly, they form the correct constraints and the first steps to solving a multitude of problems.

Ramsey interferometry, also known as the separated oscillating fields method, is a form of particle interferometry that uses the phenomenon of magnetic resonance to measure transition frequencies of particles. It was developed in 1949 by Norman Ramsey, who built upon the ideas of his mentor, Isidor Isaac Rabi, who initially developed a technique for measuring particle transition frequencies. Ramsey's method is used today in atomic clocks and in the S.I. definition of the second. Most precision atomic measurements, such as modern atom interferometers and quantum logic gates, have a Ramsey-type configuration. A more modern method, known as Ramsey–Bordé interferometry uses a Ramsey configuration and was developed by French physicist Christian Bordé and is known as the Ramsey–Bordé interferometer. Bordé's main idea was to use atomic recoil to create a beam splitter of different geometries for an atom-wave. The Ramsey–Bordé interferometer specifically uses two pairs of counter-propagating interaction waves, and another method named the "photon-echo" uses two co-propagating pairs of interaction waves.

Molecular symmetry in physics and chemistry describes the symmetry present in molecules and the classification of molecules according to their symmetry. Molecular symmetry is a fundamental concept in the application of Quantum Mechanics in physics and chemistry, for example it can be used to predict or explain many of a molecule's properties, such as its dipole moment and its allowed spectroscopic transitions, without doing the exact rigorous calculations. To do this it is necessary to classify the states of the molecule using the irreducible representations from the character table of the symmetry group of the molecule. Among all the molecular symmetries, diatomic molecules show some distinct features and they are relatively easier to analyze.

Magnetic resonance is a quantum mechanical resonant effect that can appear when a magnetic dipole is exposed to a static magnetic field and perturbed with another, oscillating electromagnetic field. Due to the static field, the dipole can assume a number of discrete energy eigenstates, depending on the value of its angular momentum quantum number. The oscillating field can then make the dipole transit between its energy states with a certain probability and at a certain rate. The overall transition probability will depend on the field's frequency and the rate will depend on its amplitude. When the frequency of that field leads to the maximum possible transition probability between two states, a magnetic resonance has been achieved. In that case, the energy of the photons composing the oscillating field matches the energy difference between said states. If the dipole is tickled with a field oscillating far from resonance, it is unlikely to transition. That is analogous to other resonant effects, such as with the forced harmonic oscillator. The periodic transition between the different states is called Rabi cycle and the rate at which that happens is called Rabi frequency. The Rabi frequency should not be confused with the field's own frequency. Since many atomic nuclei species can behave as a magnetic dipole, this resonance technique is the basis of nuclear magnetic resonance, including nuclear magnetic resonance imaging and nuclear magnetic resonance spectroscopy.

References

  1. Aquilanti, V.; Cavalli, S.; Grossi, G. (1996). "Hund's cases for rotating diatomic molecules and for atomic collisions: angular momentum coupling schemes and orbital alignment". Zeitschrift für Physik D. 36 (3–4): 215–219. Bibcode:1996ZPhyD..36..215A. doi:10.1007/BF01426406. S2CID   121444836.
  2. 1 2 3 4 5 6 Brown, John M.; Carrington, Alan (2003). Rotational Spectroscopy of Diatomic Molecules. Cambridge University Press. ISBN   0521530784.
  3. 1 2 Straughan, B. P.; Walker, S. (1976). "Chap.1 Molecular Quantum Numbers of Diatomic Molecules". Spectroscopy vol.3. Chapman and Hall. p. 9. ISBN   0-412-13390-3.
  4. 1 2 Herzberg, Gerhard (1950). Molecular Spectra and Molecular Structure, Vol I.Spectra of Diatomic Molecules (2nd ed.). van Nostrand Reinhold. pp. 219–220. Reprint 2nd ed. with corrections (1989): Krieger Publishing Company. ISBN   0-89464-268-5
  5. Nikitin, E. E.; Zare, R. N. (1994). "Correlation diagrams for Hund's coupling cases in diatomic molecules with high rotational angular momentum". Molecular Physics. 82 (1): 85–100. Bibcode:1994MolPh..82...85N. doi:10.1080/00268979400100074.
  6. 1 2 3 4 Hollas, J. Michael (1996). Modern Spectroscopy (3rd ed.). John Wiley & Sons. pp. 205–8. ISBN   0-471-96523-5.
  7. Straughan, B. P.; Walker, S. (1976). "Chap.1 Molecular Quantum Numbers of Diatomic Molecules". Spectroscopy vol.3. Chapman and Hall. p. 11. ISBN   0-412-13390-3.
  8. Straughan, B. P.; Walker, S. (1976). "Chap.1 Molecular Quantum Numbers of Diatomic Molecules". Spectroscopy vol.3. Chapman and Hall. pp. 14–15. ISBN   0-412-13390-3.
  9. Herzberg p.222. In this source is denoted as .
  10. Straughan, B. P.; Walker, S. (1976). Spectroscopy vol.2. Chapman and Hall. p. 88. ISBN   0-412-13370-9.
  11. Straughan and Walker p.14-15. In this source is denoted as .
  12. Straughan, B. P.; Walker, S. (1976). "Chap.1 Molecular Quantum Numbers of Diatomic Molecules". Spectroscopy vol.3. Chapman and Hall. p. 14. ISBN   0-412-13390-3.
  13. Carrington, A.; Pyne, C. H.; Shaw, A. M.; Taylor, S. M.; Hutson, J. M.; Law, M. M. (1996). "Microwave spectroscopy and interaction potential of the long-range He⋯Kr+ ion: An example of Hund's case (e)". The Journal of Chemical Physics. 105 (19): 8602. Bibcode:1996JChPh.105.8602C. doi:10.1063/1.472999.
  14. Lefebvre-Brion, H. (1990). "Hund's case (e): Application to Rydberg states with a 2Π ionic core". Journal of Chemical Physics. 93 (8): 5898. Bibcode:1990JChPh..93.5898L. doi:10.1063/1.459499.