Angular momentum coupling: Difference between revisions

From Citizendium
Jump to navigation Jump to search
imported>Paul Wormer
(New page: 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 su...)
 
imported>Paul Wormer
No edit summary
Line 1: Line 1:
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 is an [[atom]] with ''N'' > 1 electrons, with the electrons being its subsystems. Each electron has its own [[orbital angular momentum]], i.e., is in an eigenstate of its own angular momentum operator. For an atom, angular momentum coupling is the construction of an ''N''-electron eigenstate of the total angular momentum operator out of the individual electronic eigenstates.   
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.  
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==
==Usefulness==
Angular momentum coupling is useful under two conditions. (i) The angular momenta of the subsystems are [[constant of the motion|constants of the motion]] when the interactions between the subsystems are switched off (or neglected). (ii) Upon switching on the interactions, the angular momenta of the subsystems are no longer constants of the motion, but the total angular momentum remains a constant of the motion.  
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.  


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.
First we note that angular momentum is a time-independent and well-defined property of a physical system <ref>A constant of the motion, also referred to as a ''conserved'' property</ref> 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 [[molecule moving]] in a field-free space is an example of the second kind of system.   
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).
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. 
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).
 


<!--
<!--
Line 19: Line 28:
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.
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.


An example of the second situation is a [[rigid rotor]] moving in field-free space. A rigid rotor has a well-defined, time-independent,  angular momentum.
 


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.
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.

Revision as of 11:32, 6 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.

Usefulness

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). 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. 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

See also

Clebsch-Gordan coefficients

(To be continued)