Angular momentum coupling: Difference between revisions

From Citizendium
Jump to navigation Jump to search
imported>Paul Wormer
No edit summary
imported>Paul Wormer
No edit summary
Line 3: Line 3:
Other examples are  the coupling of spin- and orbital-angular momentum of an electron (where we see the spin and the orbital motion as subsystems of a single electron) and the coupling of nucleonic spins in the [[shell model]] of the nucleus.  
Other examples are  the coupling of spin- and orbital-angular momentum of an electron (where we see the spin and the orbital motion as subsystems of a single electron) and the coupling of nucleonic spins in the [[shell model]] of the nucleus.  


==Usefulness==
==Application==
Angular momentum coupling is useful and applicable when two conditions are satisfied. In the first place,  the angular momenta of the subsystems must be [[constant of the motion|constants of the motion]]<ref>A constant of the motion is also referred to as a ''conserved'' property</ref> in the absence of interactions between them. That is, if the interactions between the subsystems are switched off or neglected, each individual angular momentum is a constant of the motion.  When the subsystems are interacting their angular momenta  are in general no longer constants of the motion, but their sum, the total angular momentum, must still be a constant of the motion. This is the second condition.  
Angular momentum coupling is useful and applicable when two conditions are satisfied. In the first place,  the angular momenta of the subsystems must be [[constant of the motion|constants of the motion]]<ref>A constant of the motion is also referred to as a ''conserved'' property. It is represented by a Hermitian operator that commutes with the Hamiltonian of the system.</ref> in the absence of interactions between them. That is, if the interactions between the subsystems are switched off or neglected, each individual angular momentum is a constant of the motion.  When the subsystems are interacting their angular momenta  are in general no longer constants of the motion, but their sum, the total angular momentum, must still be a constant of the motion. This is the second condition.  


These two conditions are surprisingly often fulfilled due to the fact that they almost always follow from [[rotational symmetry]]&mdash;the symmetry of spherical systems and isotropic interactions.
These two conditions are surprisingly often fulfilled due to the fact that they almost always follow from [[rotational symmetry]]&mdash;the symmetry of spherical systems and isotropic interactions.
Line 10: Line 10:


As an example of angular momentum coupling in a spherical symmetric system, we consider a two-electron atom. First, assume that there is no electron-electron  interaction (or other interactions such as spin-orbit coupling), but only the electron-nucleus Coulomb attraction.  In this simplified model the atomic Hamiltonian is a sum of kinetic energies of the two electrons and the spherical symmetric electron-nucleus interaction. The kinetic energy of the nucleus, which is three to four orders of magnitude smaller than that of the electrons, is neglected. The ''orbital angular momentum'' '''l'''(''i'') (a vector operator) of  electron ''i''  (with ''i'' = 1 or 2) commutes with the total Hamiltonian.  Both operators,  '''l'''(1) and '''l(2)''',  are constant of the motion. It can be shown that the commutation of '''l'''(''i'') with the model Hamiltonian has the consequence  that electron ''i'' can be rotated around the nucleus independently of the other electron; upon rotation nothing happens to the energy of either electron (which is easy to understand because the electron-electron interaction is off and the nucleus is spherical symmetric).  
As an example of angular momentum coupling in a spherical symmetric system, we consider a two-electron atom. First, assume that there is no electron-electron  interaction (or other interactions such as spin-orbit coupling), but only the electron-nucleus Coulomb attraction.  In this simplified model the atomic Hamiltonian is a sum of kinetic energies of the two electrons and the spherical symmetric electron-nucleus interaction. The kinetic energy of the nucleus, which is three to four orders of magnitude smaller than that of the electrons, is neglected. The ''orbital angular momentum'' '''l'''(''i'') (a vector operator) of  electron ''i''  (with ''i'' = 1 or 2) commutes with the total Hamiltonian.  Both operators,  '''l'''(1) and '''l(2)''',  are constant of the motion. It can be shown that the commutation of '''l'''(''i'') with the model Hamiltonian has the consequence  that electron ''i'' can be rotated around the nucleus independently of the other electron; upon rotation nothing happens to the energy of either electron (which is easy to understand because the electron-electron interaction is off and the nucleus is spherical symmetric).  
If we now switch on the electron-electron interaction depending on the distance ''d''(1,2) between the electrons, we get the Hamiltonian ''H'', which is a very good approximation of the exact Hamiltonian. Now only a ''simultaneous and equal rotation'' of the two electrons will leave  ''d''(1,2) invariant. Independent rotation of one electron only will change the distance to the other electron and hence the electron-electron interaction energy.   
Switching on the electron-electron interaction depending on the distance ''d''(1,2) between the electrons, we get the Hamiltonian ''H'', which is a very good approximation of the exact Hamiltonian. Clearly, only a ''simultaneous and equal rotation'' of the two electrons will leave  ''d''(1,2) invariant. Independent rotation of only one electron will change the distance to the other electron and hence the electron-electron interaction energy.   
It can be shown that this implies that neither '''l'''(1) nor '''l'''(2) commute with ''H'', but their sum '''L''' = '''l'''(1) + '''l'''(2) still does. The operator '''L''' commutes with ''H'' if and only if simultaneous rotation of the two electrons leaves ''H'' invariant.
It can be shown that this implies that neither '''l'''(1) nor '''l'''(2) commute with ''H'', but their sum '''L''' = '''l'''(1) + '''l'''(2) still does. The operator '''L''' commutes with ''H'' if and only if simultaneous rotation of the two electrons leaves ''H'' invariant.
Given eigenstates of '''l'''(1) and '''l'''(2), the construction of eigenstates of '''L'''  is the ''coupling of the angular momenta of electron 1 and 2''. It is fairly easy to construct eigenstates of '''L''' by angular momentum coupling. They are labeled by a non-negative
Given eigenstates of '''l'''(1) and '''l'''(2), the construction of eigenstates of '''L'''  is the ''coupling of the angular momenta of electron 1 and 2''. It is fairly easy to construct eigenstates of '''L''' by angular momentum coupling. They are labeled by a non-negative
integer ''l''. It can be shown that eigenstates of different ''l'' do not mix under the total Hamiltonian ''H'' (which includes electron-electron interaction),
integer ''l''. It can be shown that eigenstates of different ''l'' do not mix under the total Hamiltonian ''H'' (which includes electron-electron interaction),
which means that eigenvectors of ''H'' are completely contained in a space of a single definite ''l''. This fact is a great aid in obtaining the eigenvectors of ''H'' (solution of the time-independent Schrödinger equation of the atom).  
which means that eigenvectors of ''H'' are completely contained in a space of a single definite ''l''. This fact is a great aid in obtaining the eigenvectors of ''H'' (solution of the time-independent Schrödinger equation of the atom).  
<!--
For instance, the orbit and spin of a single particle can interact through [[spin-orbit interaction]], in which case it is useful to couple the spin and  orbit angular momentum of the particle. Or two charged particles, each with a well-defined angular momentum, may interact  by [[Electrostatic#Coulomb's law|Coulomb forces]], in which case coupling of the two one-particle angular momenta  to a total angular momentum is a useful step in the solution of the two-particle [[Schrödinger equation]].
In both cases the  separate angular momenta are no longer [[constants of motion]], but the sum of the two angular momenta usually still is.  Angular momentum coupling in atoms is of importance in  atomic [[spectroscopy]]. Angular momentum coupling of [[electron spin]]s is of importance in [[quantum chemistry]]. Also in the nuclear [[shell model]] angular momentum coupling is ubiquitous.
==General theory and detailed origin==
An example of the first situation is an atom whose  electrons  only feel the Coulomb field of its nucleus. If we ignore the electron-electron interaction (and other small interactions such as spin-orbit coupling),  the ''orbital angular momentum'' '''l''' of each electron commutes with the total Hamiltonian. In this model the atomic Hamiltonian is a sum of kinetic energies of the electrons and the spherical symmetric electron-nucleus interactions. The individual electron angular momenta '''l'''(i) commute with this Hamiltonian. That is, they are conserved properties of this approximate model of the atom.
These two situations originate in classical mechanics. The third kind of conserved angular momentum, associated with [[spin (physics)|spin]], does not have a classical counterpart. However, all rules of angular momentum coupling apply to spin as well.
In general the conservation of angular momentum implies full rotational symmetry
(described by the groups [[SO(3)]] and [[SU(2)]]) and, conversely, spherical symmetry implies conservation of angular momentum. If two or more physical systems have conserved angular momenta, it can be useful to add these momenta to a total angular momentum of the combined system&mdash;a conserved property of the total system.
The building of eigenstates of the total conserved angular momentum from the angular momentum eigenstates of the individual subsystems is referred to as ''angular momentum coupling''.
Application of angular momentum coupling is useful  when there is an interaction between subsystems that, without  interaction, would have conserved angular momentum. By the very interaction the spherical symmetry of the subsystems is broken, but the angular momentum of the total system remains a constant of  motion. Use of the latter fact is helpful in the solution of the Schrödinger equation.
As an example we consider two electrons, 1 and 2, in an atom (say the helium atom). If there is no electron-electron interaction, but only electron nucleus interaction, the two electrons can be rotated around the nucleus independently of each other; nothing happens to their energy. Both operators,  '''l'''(1) and '''l(2)''',  are conserved.
However, if we switch on the electron-electron interaction depending on the distance ''d''(1,2) between the electrons, then only a simultaneous
and equal rotation of the two electrons will leave  ''d''(1,2) invariant. In such a case neither
'''l'''(1) nor '''l'''(2) is a constant of  motion but '''L''' = '''l'''(1) + '''l'''(2)
is. Given eigenstates of '''l'''(1) and '''l'''(2), the construction of eigenstates of '''L''' (which still is conserved)  is the ''coupling of the angular momenta of electron 1 and 2''.
In [[quantum mechanics]], coupling also exists between angular momenta belonging to different [[Hilbert space]]s of a single object, ''e.g.'' its [[spin (physics)|spin]] and its orbital [[angular momentum]].
Reiterating slightly differently the above:  one expands the [[quantum state]]s of composed systems (''i.e.'' made of subunits like two [[hydrogen atom]]s or two [[electron]]s) in [[basis (linear algebra)|basis sets]] which are made of [[direct product]]s of [[quantum state]]s which in turn describe the subsystems individually.  We assume that the states of the subsystems can be chosen as eigenstates of their angular momentum operators (and of their component along any arbitrary ''z'' axis).  The subsystems are therefore correctly described by a set of ''l'', ''m'' [[quantum number]]s (see [[angular momentum]] for details). When there is interaction between the subsystems, the total Hamiltonian contains terms that do not commute with the angular operators acting on the subsystems only. However, these terms ''do'' commute with the ''total'' angular momentum operator. Sometimes one refers to the non-commuting interaction terms in the Hamiltonian as ''angular momentum coupling terms'', because they necessitate the angular momentum coupling.
-->
====Footnote====
====Footnote====
<references />
<references />
==Group theoretical background==
A spherical system has symmetry group [[SO(3)]], the special orthogonal group in three dimensions.
This group consists of orthogonal 3 &times; 3 matrices with unit determinant. The group is closely related to the spin rotation group [[SU(2)]] (special unitary group), which consists of unitary 2 &times; 2 matrices with unit determinant. Both groups are [[Lie groups]] and have isomorphic [[Lie algebras]]. The Lie algebras are 3-dimensional and are generated by the components of angular momentum through the commutation relations
:<math>
[l_x, l_y] = i l_z\quad \hbox{and cyclic permutation of}\; x,\;y,\;z.
</math>
To simplify the discussion we restrict the attention to SO(3), extension to SU(2) is easy.
As is usual, one goes from the Lie algebra to the group by exponentiation:
:<math>
\mathcal{R}(\hat{\mathbf{n}}, \phi) = \exp[-i \phi \hat{\mathbf{n}}\cdot \mathbf{l}]
</math>
where <math>{\scriptstyle \mathcal{R}(\hat{\mathbf{n}}, \phi)}</math> belongs to SO(3) and represents a rotation around the unit vector <math>{\scriptstyle \hat{\mathbf{n}} }</math> over an angle &phi; and  '''l''' &equiv; (''l''<sub>''x''</sub>, ''l''<sub>''y''</sub>, ''l''<sub>''z''</sub>). In the two-electron atom example, '''l''' is the angular momentum of one electron.


<!--
By expanding the exponential operator it follows that
==Spin-orbit coupling==
:<math>
 
[H, l_x] = [H, l_y] = [H, l_z] = 0 \Longleftrightarrow [H, \mathcal{R}(\hat{\mathbf{n}}, \phi)] = 0, \quad \forall\, \hat{\mathbf{n}}\;\;\hbox{and}\;\; \forall \phi.
The behavior of [[atoms]] and smaller [[Subatomic particle|particles]] is well described by the theory of [[quantum mechanics]], in which each particle has an intrinsic angular momentum called [[spin (physics)|spin]] and specific configurations (of e.g. electrons in an atom) are described by a set of [[quantum numbers]].  Collections of particles also have angular momenta and corresponding quantum numbers, and under different circumstances the angular momenta of the parts add in different ways to form the angular momentum of the whole. Angular momentum coupling is a category including some of the ways that subatomic particles can interact with each other. 
</math>
This relation shows clearly the correspondence between rotational symmetry and angular momentum.


In [[atomic physics]], '''spin-orbit coupling''' also known as '''spin-pairing''' describes a weak magnetic interaction, or [[coupling (physics)|coupling]], of the particle [[spin (physics)|spin]] and the [[orbital motion (quantum)|orbital motion]] of this particle, e.g. the [[electron]] spin and its motion around an [[atom]]ic [[atomic nucleus|nucleus]]. One of its effects is to separate the energy of internal states of the atom, e.g. spin-aligned and spin-antialigned that would otherwise be identical in energy.  This interaction is responsible for many of the details of atomic structure.
If the electron-electron interaction is switched off in the two-electron atom, the symmetry group is the [[outer  product group]] SO(3) &times; SO(3). If the interaction is on, the symmetry group becomes the [[inner product group]], which is isomorphic to SO(3) and is the subgroup of SO(3) &times; SO(3) consisting of simultaneous and equal rotations of the two electrons.
 
The inner product group consists of the following elements
In the [[macroscopic]] world of [[astrodynamics|orbital mechanics]], the term ''spin-orbit coupling'' is sometimes used in the same sense as [[spin-orbital resonance]].
:<math>
 
\exp[-i \phi \hat{\mathbf{n}}\cdot \mathbf{l}] \otimes \exp[-i \phi \hat{\mathbf{n}}\cdot \mathbf{l}]=
===LS coupling===
\exp[-i \phi \hat{\mathbf{n}}\cdot \big(\mathbf{l}\otimes 1 + 1\otimes \mathbf{l}\big) ].
 
</math>
In light atoms (generally Z<30), electron spins '''s'''<sub>''i''</sub> interact among themselves so they combine to form a total spin angular momentum '''S'''. The same happens with orbital angular momenta '''l'''<sub>''i''</sub>, forming a single orbital angular momentum '''L'''. The interaction between the quantum numbers '''L''' and '''S''' is called '''Russell-Saunders coupling''' or '''LS coupling'''. Then '''S''' and '''L''' add together and form a total angular momentum '''J''':
Writing
 
:<math>
:<math>\mathbf J = \mathbf L + \mathbf S</math> where <math>\mathbf L = \sum_i \mathbf{l}_i</math> and <math>\mathbf S = \sum_i \mathbf{s}_i</math>
\mathbf{L} \equiv  \mathbf{l}\otimes 1 + 1\otimes \mathbf{l} \equiv \mathbf{l}(1)+  \mathbf{l}(2)
 
</math>
This is an approximation which is good as long as any external magnetic fields are weak.  In larger magnetic fields, these two momenta decouple, giving rise to a different splitting pattern in the energy levels (the '''Paschen-Back effect.'''), and the size of LS coupling term becomes small.
we find the correspondence between total '''L''' and the simultaneous rotation of the two electrons.  
 
Construction of eigenstates of ''L''<sup>2</sup> = ''L''<sub>''x''</sub><sup>2</sup> + ''L''<sub>''y''</sub><sup>2</sup> + ''L''<sub>''z''</sub><sup>2</sup> out of eigenstates of ''l''(1)<sup>2</sup> and ''l''(2)<sup>2</sup>, i.e., angular momentum coupling, is equivalent to the group theoretical subduction
For an extensive example on how LS-coupling is practically applied, see the article on [[Term symbol]]s.
:<math>
 
\mathrm{SO(3)} \times \mathrm{SO(3)} \downarrow \mathrm{SO(3)}.
===jj coupling===
</math>
In heavier atoms the situation is different. In atoms with bigger nuclear charges, spin-orbit interactions are frequently as large or larger than spin-spin interactions or orbit-orbit interactions. In this situation, each orbital angular momentum '''l'''<sub>''i''</sub> tends to combine with each individual spin angular momentum '''s'''<sub>''i''</sub>, originating individual total angular momenta '''j'''<sub>''i''</sub>. These then add up to form the total angular momentum '''J'''
:<math>\mathbf J = \sum_i \mathbf j_i = \sum_i (\mathbf{l}_i + \mathbf{s}_i)</math>  
This description, facilitating calculation of this kind of interaction, is known as '''jj coupling'''.
 
==Spin-spin coupling==
''See also: [[J-coupling]] and [[Dipolar coupling]] in NMR  spectroscopy''
 
'''Spin-spin coupling''' is the coupling of the [[spin (physics)|spin]] the intrinsic angular momentum (spin) of different particles.
Such coupling between pairs of nuclear spins is an important feature of [[Nuclear Magnetic Resonance]] spectroscopy as it can
provide detailed information about the structure and conformation of molecules. 
Spin-spin coupling between nuclear spin and electronic spin is responsible for [[hyperfine structure]] in atomic spectra.
 
== Term symbols ==
[[Term symbol]]s are used to represent the states and spectral transitions of atoms, they are found from  coupling of angular momenta mentioned above. When the state of an atom has been specified with a term symbol, the allowed transitions can be found through [[selection rule]]s by considering which transitions would conserve [[angular momentum]]. A [[photon]] has spin 1, and when there is a transition with emission or absorption of a photon the atom will need to change state to conserve angular momentum. The term symbol selection rules are. ΔS=0, ΔL=0,±1, Δl=±1, ΔJ=0,±1
 
== Relativistic effects ==
 
In very heavy atoms, relativistic shifting of the energies of the electron energy levels accentuates spin-orbit coupling effect.  Thus, for example, uranium molecular orbital diagrams must directly incorporate relativistic symbols when considering interactions with other atoms.
 
== Nuclear coupling ==
 
In atomic nuclei, the spin-orbit interaction is much stronger than for atomic electrons, and is incorporated directly into the nuclear shell model.  In addition, unlike atomic-electron term symbols, the lowest energy state is not L - S, but rather, l + s. All nuclear levels whose l value (orbital angular momentum) is greater than zero are thus split in the shell model to create states designated by l + s and l - s.  Due to the nature of the [[shell model]], which assumes an average potential rather than a central Coulombic potential, the nucleons that go into the l + s and l - s nuclear states are considered degenerate within each orbital (e.g. The 2p3/2 contains four nucleons, all of the same energy. Higher in energy is the 2p1/2 which contains two equal-energy nucleons).
-->
== See also ==
== See also ==
[[Clebsch-Gordan coefficients]]
[[Clebsch-Gordan coefficients]]
'''(To be continued)'''

Revision as of 07:02, 7 October 2007

In quantum mechanics, angular momentum coupling is the procedure of constructing eigenstates of a system's angular momentum out of angular momentum eigenstates of its subsystems. The historic example of a system to which angular momentum coupling is applied, is an atom with N > 1 electrons (the subsystems of the atom). Each electron has its own orbital angular momentum, i.e., is in an eigenstate of its own angular momentum operator. Angular momentum coupling is the construction of an N-electron eigenstate of the total atomic angular momentum operator out of the N individual electronic angular momentum eigenstates.

Other examples are the coupling of spin- and orbital-angular momentum of an electron (where we see the spin and the orbital motion as subsystems of a single electron) and the coupling of nucleonic spins in the shell model of the nucleus.

Application

Angular momentum coupling is useful and applicable when two conditions are satisfied. In the first place, the angular momenta of the subsystems must be constants of the motion[1] in the absence of interactions between them. That is, if the interactions between the subsystems are switched off or neglected, each individual angular momentum is a constant of the motion. When the subsystems are interacting their angular momenta are in general no longer constants of the motion, but their sum, the total angular momentum, must still be a constant of the motion. This is the second condition.

These two conditions are surprisingly often fulfilled due to the fact that they almost always follow from rotational symmetry—the symmetry of spherical systems and isotropic interactions. First we note that angular momentum is a constant of the motion, i.e., a time-independent and well-defined property of a physical system, in either of two situations: (i) The system is spherical symmetric, or (ii) the system moves (in quantum mechanical sense) in isotropic space. It can be shown that in both cases the angular momentum operator of the system commutes with its Hamiltonian. By Heisenberg's uncertainty relation this means that the angular momentum of the system can assume a sharp value simultaneously with the energy (eigenvalue of the Hamiltonian) of the system. The standard example of a spherical symmetric system is an atom, while a (rigid) molecule moving in a field-free space is an example of the second kind of system. A rigid (non-vibrating) molecule can be seen as a rigid rotor, which moving in field-free space, has a conserved angular momentum.

As an example of angular momentum coupling in a spherical symmetric system, we consider a two-electron atom. First, assume that there is no electron-electron interaction (or other interactions such as spin-orbit coupling), but only the electron-nucleus Coulomb attraction. In this simplified model the atomic Hamiltonian is a sum of kinetic energies of the two electrons and the spherical symmetric electron-nucleus interaction. The kinetic energy of the nucleus, which is three to four orders of magnitude smaller than that of the electrons, is neglected. The orbital angular momentum l(i) (a vector operator) of electron i (with i = 1 or 2) commutes with the total Hamiltonian. Both operators, l(1) and l(2), are constant of the motion. It can be shown that the commutation of l(i) with the model Hamiltonian has the consequence that electron i can be rotated around the nucleus independently of the other electron; upon rotation nothing happens to the energy of either electron (which is easy to understand because the electron-electron interaction is off and the nucleus is spherical symmetric). Switching on the electron-electron interaction depending on the distance d(1,2) between the electrons, we get the Hamiltonian H, which is a very good approximation of the exact Hamiltonian. Clearly, only a simultaneous and equal rotation of the two electrons will leave d(1,2) invariant. Independent rotation of only one electron will change the distance to the other electron and hence the electron-electron interaction energy. It can be shown that this implies that neither l(1) nor l(2) commute with H, but their sum L = l(1) + l(2) still does. The operator L commutes with H if and only if simultaneous rotation of the two electrons leaves H invariant. Given eigenstates of l(1) and l(2), the construction of eigenstates of L is the coupling of the angular momenta of electron 1 and 2. It is fairly easy to construct eigenstates of L by angular momentum coupling. They are labeled by a non-negative integer l. It can be shown that eigenstates of different l do not mix under the total Hamiltonian H (which includes electron-electron interaction), which means that eigenvectors of H are completely contained in a space of a single definite l. This fact is a great aid in obtaining the eigenvectors of H (solution of the time-independent Schrödinger equation of the atom).

Footnote

  1. A constant of the motion is also referred to as a conserved property. It is represented by a Hermitian operator that commutes with the Hamiltonian of the system.

Group theoretical background

A spherical system has symmetry group SO(3), the special orthogonal group in three dimensions. This group consists of orthogonal 3 × 3 matrices with unit determinant. The group is closely related to the spin rotation group SU(2) (special unitary group), which consists of unitary 2 × 2 matrices with unit determinant. Both groups are Lie groups and have isomorphic Lie algebras. The Lie algebras are 3-dimensional and are generated by the components of angular momentum through the commutation relations

To simplify the discussion we restrict the attention to SO(3), extension to SU(2) is easy. As is usual, one goes from the Lie algebra to the group by exponentiation:

where belongs to SO(3) and represents a rotation around the unit vector over an angle φ and l ≡ (lx, ly, lz). In the two-electron atom example, l is the angular momentum of one electron.

By expanding the exponential operator it follows that

This relation shows clearly the correspondence between rotational symmetry and angular momentum.

If the electron-electron interaction is switched off in the two-electron atom, the symmetry group is the outer product group SO(3) × SO(3). If the interaction is on, the symmetry group becomes the inner product group, which is isomorphic to SO(3) and is the subgroup of SO(3) × SO(3) consisting of simultaneous and equal rotations of the two electrons. The inner product group consists of the following elements

Writing

we find the correspondence between total L and the simultaneous rotation of the two electrons. Construction of eigenstates of L2 = Lx2 + Ly2 + Lz2 out of eigenstates of l(1)2 and l(2)2, i.e., angular momentum coupling, is equivalent to the group theoretical subduction

See also

Clebsch-Gordan coefficients