Skip to main content

Adiabatic, Born-Oppenheimer, and Non-adiabatic Approaches

  • Reference work entry
  • First Online:
Handbook of Computational Chemistry
  • 6616 Accesses

Abstract

A detailed derivation of the adiabatic approximation and the Born-Oppenheimer approximation is presented, the difference between these two approximations is discussed and the circumstances under which the adiabatic approximation collapses are discussed. It is shown that the solution of the Schrödinger equation in the adiabatic approximation can be divided into one representing the motion of electrons in the field of fixed nuclei and another one representing the motion of nuclei in the potential generated by the presence of the electrons. The shapes of the potential energy curves generated by the electrons and the motion of the nuclei in these potentials are also analyzed. Finally, the state-of-the-art highly accurate calculations for diatomic molecules performed without the use of the Born-Oppenheimer approximation is presented.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 1,099.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 1,399.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Precisely, the BO and adiabatic approximations are based on the difference of the time scales of movements of the nuclei and the electrons. For example, in some weakly bound molecular anions, the excess electron is very weakly bound (e.g., in the water anion). Such an electron moves at a speed comparable to the oscillating motion of atomic nuclei. A complete description of the dynamics of such an anion in the framework of the adiabatic approximation is doomed to failure. The failure can be attributed to the fact that the electron may be unbound for some large areas of the configuration space of the nuclei. In these areas the electronic wave function is not square integrable and the adiabatic corrections are divergent.

  2. 2.

    We use the most common definition of the new coordinates. Internal coordinates may be defined in many different ways, see, e.g., Piela (2007).

  3. 3.

    “Frolov’s calculations showed that when one mass increased without limit (the atomic case), any discrete spectrum persisted, but when two masses were allowed to increase without limit (the molecular cause), the Hamiltonian ceased to be well defined, and this failure led to what he called adiabatic divergence in attempts to compute discrete eigenstates of (21). This divergence is discussed in some mathematical detail in the Appendix to Frolov (1999). It does not arise from the choice of a translationally invariant form for the electronic Hamiltonian; rather it is due to the lack of any kinetic energy term to dominate the Coulomb potential. Thus it really is essential to characterize the spectrum of H elec to see whether the traditional approach can be validated.” (Sutcliffe 2003)

  4. 4.

    For the sake of simplicity, we denote ψ k el(r; R) ≡ ψ k el and χ k (R) ≡ χ k .

  5. 5.

    As a result of simple transformations:

  6. 6.

    More inventive faculty trying to explain to students the BO approximation takes an example of a cow (symbolizes the nucleus) and flies flying around it (electrons). Flies almost immediately adapt to the current position of the cow, just because they are lighter and move faster. Therefore, the cow only sees a cloud of flies, while the flies only see a static cow.

  7. 7.

    In fact, there exist rovibrational states, e.g., in the helium dimer, which cannot be properly described by the model of a harmonic oscillator.

  8. 8.

    This model works well for molecules like NH3, CH3Cl, C6H6. In a general case the asymmetric top model should be used (see Haken and Wolf 2004 for details).

  9. 9.

    Crossing the states of the same symmetry is possible if you work within the adiabatic approximation.

  10. 10.

    From now on the symbol “tot” will be used to denote the sum of the total nonrelativistic energy and the corrections up to the certain order in terms of the hyperfine constant α.

  11. 11.

    Optimal values of these parameters in the basis functions (126) and (127) are determined using the gradient method.

  12. 12.

    Unfortunately in the results of Wolniewicz, the values for the v = 22 are missing.

References

  • Bethe, H. A., & Salpeter, E. E. (1957). Quantum mechanics of one- and two-electron systems. Berlin: Springer.

    Book  Google Scholar 

  • Born, M., & Oppenheimer, J. R. (1927). Zur Quantentheorie der Molekeln (On the quantum theory of molecules). Annalen der Physik, 84, 457–484.

    Article  CAS  Google Scholar 

  • Bubin, S., Leonarski, F., Stanke, M., & Adamowicz, L. (2009). Charge asymmetry in pure vibrational states of the HD molecule. The Journal of Chemical Physics, 130, 124120.

    Article  Google Scholar 

  • Cafiero, M., & Adamowicz, L. (2002). Nonadiabatic calculations of the dipole moments of LiH and LiD. Physical Review Letters, 88, 33002.

    Article  Google Scholar 

  • Dalgarno, A., & McCarroll, R. (1956). Adiabatic coupling between electronic and nuclear motion in molecules. Proceedings of the Royal Society (London), A237, 383.

    Article  Google Scholar 

  • Davydov, A. S. (1965). Quantum mechanics. Oxford: Pergamon Press.

    Google Scholar 

  • Davydov, A. S. (1976). Quantum mechanics (2nd ed.). Oxford: Pergamon Pr.

    Google Scholar 

  • Demtröder, W. (2010). Atoms, molecules and photons: An introduction to atomic-, molecular- and quantum physics (2nd ed., 2010 ed.). Berlin: Springer, Berlin (January 19, 2011).

    Google Scholar 

  • Frolov, A. M. (1999). Bound-state calculations of Coulomb three-body systems. Physical Review A, 59, 4270.

    Article  CAS  Google Scholar 

  • Haken, H., & Wolf, H. C. (2010). Molecular physics and elements of quantum chemistry (2nd Edn., 2004 edition). Berlin: Springer.

    Google Scholar 

  • Handy, N. C., & Lee, A. M. (1986). The adiabatic approximation. Chemical Physics Letters, 252, 425–430.

    Article  Google Scholar 

  • Herzberg, G. (1951). Spectra of diatomic molecules (2nd ed.). D. Van Nostrand, New York.

    Google Scholar 

  • Howells, M. H., & Kennedy, R. A. (1990). Relativistic corrections for the ground and first excited states of H2 +, HD+ and D2 +. Journal of the Chemical Society Faraday Transactions, 86, 3495.

    Article  CAS  Google Scholar 

  • Hulburt, H. M., & Hirschfelder, J. O. (1941). Potential energy functions for diatomic molecules. The Journal of Chemical Physics, 9, 61.

    Article  Google Scholar 

  • Jahn, H. A., & Teller, E. (1937). Stability of polyatomic molecules in degenerate electronic states. I. Orbital degeneracy. Proceedings of the Royal Society of London Series A, 161, 220.

    Google Scholar 

  • Kinghorn, D. B., & Adamowicz, L. (1997). The Journal of Chemical Physics, 106, 4589.

    Article  CAS  Google Scholar 

  • Kołos, W., & Wolniewicz, L. (1963). Nonadiabatic theory for diatomic molecules and its application to the hydrogen molecule. Reviews of Modern Physics, 35, 473.

    Article  Google Scholar 

  • Kołos, W. (1970). Adiabatic approximation and its accuracy. Advanced in Quantum Chemistry, 5, 99–133.

    Article  Google Scholar 

  • Kolos, W., & Sadlej, J. (1998). Atom i czasteczka (in Polish). Warszawa: WNT

    Google Scholar 

  • Krȩglewski, M. (1979). Zadania z chemii kwantowej, Wydawnictwo Naukowe Uniwersytetu im. Adama Mickiewicza w Poznaniu.

    Google Scholar 

  • Landau, L. D., & Lifschitz, E. M. (1981). Quantum mechanics – Non relativistic theory (Course of theoretical physics, Vol. 3, 3rd ed.). Oxford: Pergamon Press. Butterworth-Heinemann.

    Google Scholar 

  • Pachucki, K., & Grotch, H. (1995). Pure recoil corrections to hydrogen energy levels. Physical Review A, 51, 1854.

    Article  CAS  Google Scholar 

  • Pavanello, M., Adamowicz, L., Alijah, A., Zobov, N. F., Mizus, I. I., Polyansky, O. L., Tennyson, J., Szidarovszky, T., Császár, A. G., Berg, M., Petrignani, A., & Wolf, A. (2012). Precision measurements and computations of transition energies in rotationally cold triatomic hydrogen ions up to the midvisible spectral range. Physical Review Letters, 108, 023002.

    Article  Google Scholar 

  • Piela, L. (2007). Ideas of quantum chemistry (1st ed.). Amsterdam: Elsevier Science., Amsterdam.

    Google Scholar 

  • Ruch, E., & Schönhofer, A. (1965). Ein Beweis des Jahn-Teller-Theorems mit Hilfe eines Satzes über die Induktion von Darstellungen endlicher Gruppen. Theoretica Chimica Acta, 3, 291–304.

    Article  CAS  Google Scholar 

  • Stanke, M., & Adamowicz, L. (2013). Molecular relativistic corrections determined in the framework where the Born-Oppenheimer approximation is not assumed. The Journal of Physical Chemistry A, 117 (39), 10129–10137.

    Article  CAS  Google Scholar 

  • Stanke, M., Kȩdziera, D., Molski, M., Bubin, S., Barysz, M., & Adamowicz, L. (2006). Convergence of experiment and theory on the pure vibrational spectrum of HeH+. Physical Review Letters, 96, 233002.

    Article  Google Scholar 

  • Stanke, M., Kȩdziera, D., Bubin, S., & Adamowicz, L. (2007a). Lowest excitation energy of9Be. Physical Review Letters, 99, 043001.

    Article  Google Scholar 

  • Stanke, M., Kȩdziera, D., Bubin, S., & Adamowicz, L. (2007b). Ionization potential of9Be calculated including nuclear motion and relativistic corrections. Physical Review A, 75, 052510.

    Article  Google Scholar 

  • Stanke, M., Kȩdziera, D., Bubin, S., & Adamowicz, L. (2008). Complete α 2 relativistic corrections to the pure vibrational non-Born-Oppenheimer energies of HeH+. Physical Review A, 77, 022506.

    Article  Google Scholar 

  • Stanke, M., Bubin, S., & Adamowicz, L. (2009a). Fundamental vibrational transitions of the3He4He+ and7LiH+ ions calculated without assuming the Born-Oppenheimer approximation and with leading relativistic corrections. Physical Review A, 79, 060501(R).

    Google Scholar 

  • Stanke, M., Komasa, J., Bubin, S., & Adamowicz, L. (2009b). Five lowest1S states of the Be atom calculated with a finite-nuclear-mass approach and with relativistic and QED corrections. Physical Review A, 80, 022514.

    Article  Google Scholar 

  • Sutcliffe, B. T. (2003). Approximate separation of electronic and nuclear motion, Part 6. In S. Wilson, P. F. Bernath, & R. McWeeny (Eds.), Handbook of molecular physics and quantum chemistry (Vol. 1, p. 475). Chichester: Wiley.

    Google Scholar 

  • Wolniewicz, L. (2011). Private consultations.

    Google Scholar 

  • Wolniewicz, L. (1995). Nonadiabatic energies of the ground state of the hydrogen molecule. The Journal of Chemical Physics, 103, 1792.

    Article  CAS  Google Scholar 

  • Wolniewicz, L., & Orlikowski, T. (1991). The 1sσ g and 2pσ u states of the H2 +, D2 + and HD+ ions. Molecular Physics, 74, 103–111.

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work has been supported by the Polish National Science Centre, grant DEC-2013/10/E/ST4/00033. I am grateful to Professor Krzysztof Strasburger, the reviewer, for his valuable comments and to Professors Jacek Karwowski and Ludwik Adamowicz for useful discussions. Thanks are extended to Ms. Ewa Palikot, MSc. for drawing the figures.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Monika Stanke .

Editor information

Editors and Affiliations

Appendices

Appendix A

Detailed derivation of the Laplace operator in the new coordinate (7), (8), (9),and (10)

  • First derivatives of the nuclear coordinates:

    • nucleus a

      $$ \displaystyle\begin{array}{rcl} \frac{\partial } {\partial X_{a}^{{\prime}}}& =& \frac{\partial X_{\mathrm{CM}}} {\partial X_{a}^{{\prime}}} \frac{\partial } {\partial X_{\mathrm{CM}}}+\frac{\partial Y _{\mathrm{CM}}} {\partial X_{a}^{{\prime}}} \frac{\partial } {\partial Y _{\mathrm{CM}}}+\frac{\partial Z_{\mathrm{CM}}} {\partial X_{a}^{{\prime}}} \frac{\partial } {\partial Z_{\mathrm{CM}}}+ \frac{\partial R_{x}} {\partial X_{a}^{{\prime}}} \frac{\partial } {\partial R_{x}}+ \frac{\partial R_{y}} {\partial X_{a}^{{\prime}}} \frac{\partial } {\partial R_{y}}+ \frac{\partial R_{z}} {\partial X_{a}^{{\prime}}} \frac{\partial } {\partial R_{z}}+ \\ & +& \sum _{i=1}^{N_{e} }\left ( \frac{\partial x_{i}} {\partial X_{a}^{{\prime}}} \frac{\partial } {\partial x_{i}} + \frac{\partial y_{i}} {\partial X_{a}^{{\prime}}} \frac{\partial } {\partial y_{i}} + \frac{\partial z_{i}} {\partial X_{a}^{{\prime}}} \frac{\partial } {\partial z_{i}}\right )\; =\; \frac{M_{a}} {M} \frac{\partial } {\partial X_{\mathrm{CM}}} + \frac{\partial } {\partial R_{x}} -\frac{1} {2}\sum _{i=1}^{N_{e} } \frac{\partial } {\partial x_{i}} {}\end{array}$$
      (128)

      analogously:

      $$\displaystyle\begin{array}{rcl} \frac{\partial } {\partial Y _{a}^{{\prime}}}& =& \frac{M_{a}} {M} \frac{\partial } {\partial Y _{\mathrm{CM}}} + \frac{\partial } {\partial R_{y}} -\frac{1} {2}\sum _{i=1}^{N_{e} } \frac{\partial } {\partial y_{i}}, \\ \frac{\partial } {\partial Z_{a}^{{\prime}}}& =& \frac{M_{a}} {M} \frac{\partial } {\partial Z_{\mathrm{CM}}} + \frac{\partial } {\partial R_{z}} -\frac{1} {2}\sum _{i=1}^{N_{e} } \frac{\partial } {\partial z_{i}} {}\end{array}$$
      (129)
    • nucleus b - analogous:

      $$\displaystyle\begin{array}{rcl} \frac{\partial } {\partial X_{b}^{{\prime}}}& =& =\; \frac{M_{a}} {M} \frac{\partial } {\partial X_{\mathrm{CM}}} - \frac{\partial } {\partial R_{x}} -\frac{1} {2}\sum _{i=1}^{N_{e} } \frac{\partial } {\partial x_{i}} {}\end{array}$$
      (130)

    and similarly for the nuclear coordinates \(\tilde{y}_{b}\) and \(\tilde{z}_{b}\)

  • First derivatives of the electron coordinates:

    $$\displaystyle\begin{array}{rcl} \frac{\partial } {\partial x^{{\prime}}_{i}}& =& \frac{X_{\mathrm{CM}}} {\partial x^{{\prime}}_{i}} \frac{\partial } {X_{\mathrm{CM}}}+\frac{Y _{\mathrm{CM}}} {\partial x^{{\prime}}_{i}} \frac{\partial } {Y _{\mathrm{CM}}}+\frac{Z_{\mathrm{CM}}} {\partial x^{{\prime}}_{i}} \frac{\partial } {Z_{\mathrm{CM}}}+ \frac{R_{x}} {\partial x^{{\prime}}_{i}} \frac{\partial } {R_{x}} + \frac{R_{y}} {\partial x^{{\prime}}_{i}} \frac{\partial } {R_{y}} + \frac{R_{z}} {\partial x^{{\prime}}_{i}} \frac{\partial } {R_{z}} + \\ & & +\!\sum _{j=1}^{N_{e} }\left ( \frac{\partial x_{j}} {\partial x^{{\prime}}_{i}} \frac{\partial } {\partial x_{j}}+ \frac{\partial y_{j}} {\partial x^{{\prime}}_{i}} \frac{\partial } {\partial y_{j}}+ \frac{\partial z_{j}} {\partial x^{{\prime}}_{i}} \frac{\partial } {\partial z_{j}}\right )= \frac{m} {M} \frac{\partial } {\partial X_{\mathrm{CM}}}+\!\sum _{j=1}^{N_{e} }\left (\delta _{ij} \frac{\partial } {\partial x_{j}}\right )= \\ & =& \frac{m} {M} \frac{\partial } {\partial X_{\mathrm{CM}}} + \frac{\partial } {\partial x_{i}} {}\end{array}$$
    (131)

    and similarly for the electron coordinates y i and z i

  • Second derivatives of the nuclear coordinates (we use the formula \(\left [a+(b+c)\right ]^{2} = a^{2} + b^{2} + c^{2} + 2ab + 2bc + 2ac\)):

    • nucleus a

      $$\displaystyle\begin{array}{rcl} \frac{\partial ^{2}} {\partial X^{{\prime}}_{a}{}^{2}}& =& \!\left (\frac{M_{a}} {M} \frac{\partial } {\partial X_{\mathrm{CM}}}+ \frac{\partial } {\partial R_{x}}-\frac{1} {2}\sum _{i=1}^{N_{e} } \frac{\partial } {\partial x_{i}}\right )\left (\frac{M_{a}} {M} \frac{\partial } {\partial X_{\mathrm{CM}}}+ \frac{\partial } {\partial R_{x}}-\frac{1} {2}\sum _{i=1}^{N_{e} } \frac{\partial } {\partial x_{i}}\right )\!= \\ & =& \left (\frac{M_{a}} {M} \right )^{2} \frac{\partial ^{2}} {\partial X_{\mathrm{CM}}^{2}} + \frac{\partial ^{2}} {\partial R_{x}^{2}} + \frac{1} {4}\sum _{i=1}^{N_{e} }\left ( \frac{\partial } {\partial x_{i}}\right ) + \\ & & +2\frac{M_{a}} {M} \frac{\partial ^{2}} {\partial X_{\mathrm{CM}}\partial R_{x}} - \frac{\partial } {\partial X_{\mathrm{CM}}}\sum _{i=1}^{N_{e} } \frac{\partial } {\partial x_{i}} -\frac{M_{a}} {M} \frac{\partial } {\partial R_{x}}\sum _{i=1}^{N_{e} } \frac{\partial } {\partial x_{i}} {}\end{array}$$
      (132)
    • nucleus b

      $$\displaystyle\begin{array}{rcl} \frac{\partial ^{2}} {\partial X^{{\prime}}_{b}{}^{2}}& =& \left (\frac{M_{b}} {M} \frac{\partial } {\partial X_{\mathrm{CM}}}- \frac{\partial } {\partial R_{x}}-\frac{1} {2}\sum _{i=1}^{N_{e} } \frac{\partial } {\partial x_{i}}\right )\left (\frac{M_{b}} {M} \frac{\partial } {\partial X_{\mathrm{CM}}}- \frac{\partial } {\partial R_{x}}-\frac{1} {2}\sum _{i=1}^{N_{e} } \frac{\partial } {\partial x_{i}}\right )= \\ & =& \left (\frac{M_{b}} {M} \right )^{2} \frac{\partial ^{2}} {\partial X_{\mathrm{CM}}^{2}} + \frac{\partial ^{2}} {\partial R_{x}^{2}} + \frac{1} {4}\sum _{i=1}^{N_{e} }\left ( \frac{\partial } {\partial x_{i}}\right ) + \\ & & -2\frac{M_{b}} {M} \frac{\partial ^{2}} {\partial X_{\mathrm{CM}}\partial R_{x}}+ \frac{\partial } {\partial X_{\mathrm{CM}}}\sum _{i=1}^{N_{e} } \frac{\partial } {\partial x_{i}}-\frac{M_{b}} {M} \frac{\partial } {\partial R_{x}}\sum _{i=1}^{N_{e} } \frac{\partial } {\partial x_{i}} {}\end{array}$$
      (133)
  • Second derivatives of the electron coordinates:

    $$\displaystyle\begin{array}{rcl} \frac{\partial ^{2}} {\partial x^{{\prime}}_{i}{}^{2}}& =& \left ( \frac{m} {M} \frac{\partial } {\partial X_{\mathrm{CM}}} + \frac{\partial } {\partial x_{i}}\right )\left ( \frac{m} {M} \frac{\partial } {\partial X_{\mathrm{CM}}} + \frac{\partial } {\partial x_{i}}\right ) \\ & =& \left ( \frac{m} {M}\right )^{2} \frac{\partial ^{2}} {\partial X_{\mathrm{CM}}^{2}} + \frac{\partial ^{2}} {\partial x_{i}^{2}} + 2 \frac{m} {M} \frac{\partial ^{2}} {\partial X_{\mathrm{CM}}\partial x_{i}} {}\end{array}$$
    (134)

    Similarly, the remaining coordinates.

Appendix B

Flat rotor – a system of two particles with masses m 1 and m 2 at a constant distance R from each other. In this system there are no external forces, but it can rotate around its center of mass. Hamiltonian for this system has a form:

$$\displaystyle{ \hat{H}= - \frac{1} {2m_{1}}\nabla _{\mathbf{r}_{1}}^{2} - \frac{1} {2m_{2}}\nabla _{\mathbf{r}_{2}}^{2}, }$$
(135)

where r 1 and r 2 are vectors that describe the positions of the two bodies in the laboratory reference frame. Alternatively one can describe the same system, in the center-of-mass reference frame. The coordinates of the center of mass in the laboratory frame are given by

$$\displaystyle{ \mathbf{R} = \frac{m_{1}\mathbf{r}_{1} + m_{2}\mathbf{r}_{2}} {m_{1} + m_{2}} \; =\; \left [X,Y,Z\right ] }$$
(136)

If one wants to express the Hamiltonian in the center-of-mass coordinates, one has to express the second derivatives in the new variables. Proceeding as in Appendix A one gets

$$\displaystyle\begin{array}{rcl} \frac{\partial } {\partial x_{1}}& =& \frac{\partial X} {\partial x_{1}}\; \frac{\partial } {\partial X} + \frac{\partial Y } {\partial x_{1}}\; \frac{\partial } {\partial Y } + \frac{\partial Z} {\partial x_{1}}\; \frac{\partial } {\partial Z} + \frac{\partial x} {\partial x_{1}}\; \frac{\partial } {\partial x} + \frac{\partial y} {\partial x_{1}}\; \frac{\partial } {\partial y} + \frac{\partial z} {\partial x_{1}}\; \frac{\partial } {\partial z} \\ & =& \frac{\partial X} {\partial x_{1}}\; \frac{\partial } {\partial X} + \frac{\partial x} {\partial x_{1}}\; \frac{\partial } {\partial x} \\ & =& \frac{m_{1}} {m_{1} + m_{2}}\; \frac{\partial } {\partial X} + \frac{\partial } {\partial x}, {}\end{array}$$
(137)
$$\displaystyle\begin{array}{rcl} \frac{\partial } {\partial x_{2}}& =& \frac{\partial X} {\partial x_{2}}\; \frac{\partial } {\partial X} + \frac{\partial Y } {\partial x_{2}}\; \frac{\partial } {\partial Y } + \frac{\partial Z} {\partial x_{2}}\; \frac{\partial } {\partial Z} + \frac{\partial x} {\partial x_{2}}\; \frac{\partial } {\partial x} + \frac{\partial y} {\partial x_{2}}\; \frac{\partial } {\partial y} + \frac{\partial z} {\partial x_{2}}\; \frac{\partial } {\partial z} \\ & =& \frac{\partial X} {\partial x_{2}}\; \frac{\partial } {\partial X} + \frac{\partial x} {\partial x_{2}}\; \frac{\partial } {\partial x} \\ & =& \frac{m_{2}} {m_{1} + m_{2}}\; \frac{\partial } {\partial X} - \frac{\partial } {\partial x}. {}\end{array}$$
(138)

Similarly, for the remaining coordinates. The second derivatives are given by

$$\displaystyle\begin{array}{rcl} \frac{\partial ^{2}} {\partial x_{1}^{2}}& =& \left ( \frac{m_{1}} {m_{1} + m_{2}}\; \frac{\partial } {\partial X} + \frac{\partial } {\partial x}\right ) \cdot \left ( \frac{m_{1}} {m_{1} + m_{2}}\; \frac{\partial } {\partial X} + \frac{\partial } {\partial x}\right ) \\ & =& \left ( \frac{m_{1}} {m_{1} + m_{2}}\right )^{2} \frac{\partial ^{2}} {\partial X^{2}} + \frac{2m_{1}} {m_{1} + m_{2}}\; \frac{\partial ^{2}} {\partial X\;\partial x} + \frac{\partial ^{2}} {\partial x^{2}}{}\end{array}$$
(139)
$$\displaystyle\begin{array}{rcl} \frac{\partial ^{2}} {\partial x_{2}^{2}}& =& \left ( \frac{m_{2}} {m_{1} + m_{2}}\; \frac{\partial } {\partial X} - \frac{\partial } {\partial x}\right ) \cdot \left ( \frac{m_{2}} {m_{1} + m_{2}}\; \frac{\partial } {\partial X} - \frac{\partial } {\partial x}\right ) \\ & =& \left ( \frac{m_{2}} {m_{1} + m_{2}}\right )^{2} \frac{\partial ^{2}} {\partial X^{2}} - \frac{2m_{2}} {m_{1} + m_{2}}\; \frac{\partial ^{2}} {\partial X\;\partial x} + \frac{\partial ^{2}} {\partial x^{2}}{}\end{array}$$
(140)

Thus, after substituting the second derivatives, one obtains the Hamiltonian operator (135) in the new variables

$$\displaystyle\begin{array}{rcl} \hat{H}& =& -\frac{1} {2}\left [ \frac{m_{1}} {(m_{1} + m_{2})^{2}} + \frac{m_{2}} {(m_{1} + m_{2})^{2}}\right ]\left ( \frac{\partial ^{2}} {\partial X^{2}} + \frac{\partial ^{2}} {\partial Y ^{2}} + \frac{\partial ^{2}} {\partial Z^{2}}\right ) \\ & -& \frac{1} {2}\left [ \frac{1} {m_{1}} + \frac{1} {m_{2}}\right ]\left ( \frac{\partial ^{2}} {\partial x^{2}} + \frac{\partial ^{2}} {\partial y^{2}} + \frac{\partial ^{2}} {\partial z^{2}}\right ) \\ & =& - \frac{1} {2M}\;\Delta _{\mathbf{R}} -\frac{1} {2\mu }\;\Delta _{\mathbf{R}} \equiv \hat{H} _{\mathbf{R}} + \hat{H} _{\mathbf{R}}, {}\end{array}$$
(141)

where

$$\displaystyle\begin{array}{rcl} M& =& m_{1} + m_{2},\qquad \frac{1} {\mu } = \frac{1} {m_{1}} + \frac{1} {m_{2}},\qquad \Delta _{\mathbf{R}} = \frac{\partial ^{2}} {\partial X^{2}} + \frac{\partial ^{2}} {\partial Y ^{2}} + \frac{\partial ^{2}} {\partial Z^{2}}, \\ \Delta _{\mathbf{R}}& =& \frac{\partial ^{2}} {\partial x^{2}} + \frac{\partial ^{2}} {\partial y^{2}} + \frac{\partial ^{2}} {\partial z^{2}}, {}\end{array}$$
(142)

and μ is the reduced mass. For the two particles moving freely in the space, that is assuming that V = 0, the eigenvalue equation of (141) can be written as \(\hat{H}_{\mathbf{R}}\) and \(\hat{H}_{\mathbf{R}}\):

$$\displaystyle{ (\hat{H} _{\mathbf{R}} + \hat{H} _{\mathbf{R}})\psi (\mathbf{R},\mathbf{r}) = E\psi (\mathbf{R},\mathbf{r}) }$$
(143)

for

$$\displaystyle{\psi (\mathbf{R},\;\mathbf{r}) =\psi _{\mathbf{R}}(\mathbf{R})\;\psi _{\mathbf{R}}(\mathbf{r}),}$$

where

$$\displaystyle{\hat{H} _{\mathbf{R}}\;\psi _{\mathbf{R}}(\mathbf{R}) = E_{\mathbf{R}}\;\psi _{\mathbf{R}}(\mathbf{R})\qquad \mbox{ and}\qquad \hat{H} _{\mathbf{R}}\;\psi _{\mathbf{R}}(\mathbf{r}) = E_{\mathbf{R}}\;\psi _{\mathbf{R}}(\mathbf{r}).}$$

We focus only on the relative motion Hamiltonian \(\hat{H}_{\mathbf{R}}\). The eigenfunction ψ R (r) of this Hamiltonian depends on three variables. However, due to the fact that r is constant (r 2 = x 2 + y 2 + z 2 = const), only two angular variables are independent. In the spherical coordinates, with variables r, θ, ϕ, where r = const:

$$\displaystyle\begin{array}{rcl} x& =& x(r,\theta,\phi ) = r\sin \theta \cos \phi \, {}\\ y& =& y(r,\theta,\phi ) = r\sin \theta \sin \phi \, {}\\ z& =& z(r,\theta,\phi ) = r\cos \theta. {}\\ \end{array}$$

the Laplacian can be written as

$$\displaystyle{\Delta _{\mathbf{R}} = \frac{1} {r^{2}}\left [ \frac{\partial } {\partial r}\left (r^{2} \frac{\partial } {\partial r}\right ) + \frac{1} {\sin \theta } \frac{\partial } {\partial \theta }\left (\sin \theta \frac{\partial } {\partial \theta }\right ) + \frac{1} {\sin ^{2}\theta } \frac{\partial ^{2}} {\partial \phi ^{2}}\right ].}$$

Consequently, the rigid rotor Hamiltonian becomes

$$\displaystyle{\hat{H}= -\frac{1} {2I}\left [\frac{1} {\sin \theta } \frac{\partial } {\partial \theta }\left (\sin \theta \frac{\partial } {\partial \theta }\right ) + \frac{1} {\sin ^{2}\theta } \frac{\partial ^{2}} {\partial \phi ^{2}}\right ],}$$

where

$$\displaystyle{I =\mu R^{2}}$$

is the moment of inertia. The eigenvalue problem of the rigid rotor is, thus, reduced to the equation

$$\displaystyle{-\frac{1} {2I}\left [\frac{1} {\sin \theta } \frac{\partial } {\partial \theta }\left (\sin \theta \frac{\partial } {\partial \theta }\right ) + \frac{1} {\sin ^{2}\theta } \frac{\partial ^{2}} {\partial \phi ^{2}}\right ]Y (\theta,\phi ) = EY (\theta,\phi ),}$$

analogous to the problem of the angular momentum

$$\displaystyle{-1\left [\frac{1} {\sin \theta } \frac{\partial } {\partial \theta }\left (\sin \theta \frac{\partial } {\partial \theta }\right ) + \frac{1} {\sin ^{2}\theta } \frac{\partial ^{2}} {\partial \phi ^{2}}\right ]Y (\theta,\phi ) = 1l(l + 1)Y (\theta,\phi ).}$$

Thus, we can immediately conclude that the rigid rotor eigenfunctions are equal to the spherical harmonics

$$\displaystyle{Y _{J}^{M}(\theta,\phi ) = \frac{1} {\sqrt{2\pi }}N_{J,\vert M\vert }P_{J}^{\vert M\vert }(\cos \theta )e^{iM\phi },}$$

where

$$\displaystyle{P_{l}^{m}(x) = (1 - x^{2})^{m/2} \frac{d^{m}} {dx^{m}}P_{l}(x)}$$

is the associate Legendre polynomial. The eigenvalues are given by

$$\displaystyle{E_{J} = \frac{1} {2I}J(J + 1),}$$

where 1∕2I is the rotational constant.

Rights and permissions

Reprints and permissions

Copyright information

© 2017 Springer International Publishing Switzerland

About this entry

Cite this entry

Stanke, M. (2017). Adiabatic, Born-Oppenheimer, and Non-adiabatic Approaches. In: Leszczynski, J., Kaczmarek-Kedziera, A., Puzyn, T., G. Papadopoulos, M., Reis, H., K. Shukla, M. (eds) Handbook of Computational Chemistry. Springer, Cham. https://doi.org/10.1007/978-3-319-27282-5_41

Download citation

Publish with us

Policies and ethics