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]

Angular momenta edit

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

  •  , the electronic orbital angular momentum
  •  , the electronic spin angular momentum
  •  , the total electronic angular momentum
  •  , the rotational angular momentum of the nuclei
  •  , the total angular momentum of the system (exclusive of nuclear spin)
  •  , the total angular momentum exclusive of electron (and nuclear) spin

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 edit

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 case Electrostatic Spin-orbit Rotational
(a) strong intermediate weak
(b) strong weak intermediate
(c) intermediate strong weak
(d) intermediate weak strong
(e) weak intermediate strong
strong intermediate

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) edit

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) edit

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) edit

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) edit

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) edit

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]

References edit

  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. ^ a b c d e f Brown, John M.; Carrington, Alan (2003). Rotational Spectroscopy of Diatomic Molecules. Cambridge University Press. ISBN 0521530784.
  3. ^ a b 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. ^ a b 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. ^ a b c d 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.