Skip to main content

Part of the book series: Progress in Theoretical Chemistry and Physics ((PTCP,volume 27))

Abstract

The idea of a Potential Energy Surface (PES) forms the basis of almost all accounts of the mechanisms of chemical reactions, and much of theoretical molecular spectroscopy. It is assumed that, in principle, the PES can be calculated by means of clamped-nuclei electronic structure calculations based upon the Schrödinger Coulomb Hamiltonian. This article is devoted to a discussion of the origin of the idea, its development in the context of the Old Quantum Theory, and its present status in the quantum mechanics of molecules. It is argued that its present status must be regarded as uncertain.

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 169.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 219.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 219.99
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.

    René Marcelin was killed in action fighting for France in September 1914.

  2. 2.

    That a molecule must reach a certain region of space at a suitable angle, that its speed must exceed a certain limit, that its internal structure must correspond to an unstable configuration etc.; …

  3. 3.

    Nevertheless it seems proper to regard Marcelin’s introduction of phase-space variables and a critical reaction surface into chemical dynamics as the beginning of a formulation of the Transition State Theory that was developed by Wigner in the 1930’s [1215]. The 2n phase-space variables q,p were identified with the n nuclei specified in the chemical formula of the participating species, and the Hamiltonian was that for classical nuclear motion on a Potential Energy Surface; this dynamics was assumed to give rise to a critical surface which was such that reaction trajectories cross the surface precisely once. The classical nature of the formalism was quite clear because the Uncertainty Principle precludes the precise specification of position on the critical surface simultaneously with the momentum of the nuclei.

  4. 4.

    This is strictly true only for integrable Hamiltonians [26].

  5. 5.

    The difficulties for action-angle quantization posed by the existence of chaotic motions in non-separable systems [30] were recognized by Einstein at the time the Old Quantum Theory was developed [31].

  6. 6.

    This is the earliest reference we know of where the idea of adiabatic separation of the electrons and the nuclei is proposed explicitly.

  7. 7.

    This also deals with the uninteresting overall translation of the molecule.

  8. 8.

    The rotational and vibrational energies occur together because of the choice of the parameter λ; as is well-known, Born and Oppenheimer later showed that a better choice is to take the quarter power of the mass ratio as this separates the vibrational and rotational energies in the orders of the perturbation expansion [38].

  9. 9.

    The details can be found in the original paper [38], and in various English language presentations, for example [4446].

  10. 10.

    If one sets κ=0…one obtains a differential equation in the x k alone, the X l appearing as parameters:…. Evidently, this represents the electronic motion for stationary nuclei.

  11. 11.

    W(X) in the notation of the above quotation.

  12. 12.

    The reader may find it helpful to refer to the Appendix which summarizes some mathematical notions that are needed here, and illustrates them in a simple model of coupled oscillators with two degrees of freedom.

  13. 13.

    It is always possible to split off the kinetic energy of the centre-of-mass without any approximation; with this choice we retain the separation of the electronic and nuclear kinetic energies as well, as in (1.24). Explicit formulae are given in e.g. [3] where it is shown that the nuclear kinetic energy terms involve reciprocals of the nuclear masses, so that overall, the nuclear kinetic energy is proportional to κ 4.

  14. 14.

    After the elimination of the centre-of-mass variables \(\hat{\mathsf{H}}^{\mathrm{elec}}\) is playing the role of \(\hat{H}_{o}\) in (1.20).

  15. 15.

    We assume that the centre-of-mass contributions are eliminated as usual.

  16. 16.

    The multiminima case can also be treated in this way.

  17. 17.

    The Lorentz Theory of the electron for example [77].

  18. 18.

    A similar requirement must be placed on the denominator in (12) of [85] for the equation to provide a secure definition.

  19. 19.

    In its original form b=b o , the equilibrium configuration, on the right-hand side of (1.50).

  20. 20.

    The variables are expressed in dimensionless form for simplicity. The quantum oscillator \(\hat{\mathsf{h}}=\frac{1}{2}(\hat{\mathsf{p}}^{2}+\hat{\mathsf{q}}^{2})\) has eigenvalues \(n+\frac{1}{2}\).

References

  1. Löwdin P-O (1989) Pure Appl Chem 61:2065–2074

    Article  Google Scholar 

  2. Thomson JJ (1899) Philos Mag 48:547–567

    CAS  Google Scholar 

  3. Woolley RG, Sutcliffe BT (2003) In: Brändas EJ, Kryachko ES (eds) Fundamental world of quantum chemistry, vol 1. Kluwer Academic, Dordrecht

    Google Scholar 

  4. Sutcliffe BT, Woolley RG (2012) J Chem Phys 137:22A544

    Article  Google Scholar 

  5. Marcelin R (1914) Contribution à l’étude de la cinétique physico-chimique. Gauthier-Villars, Paris

    Google Scholar 

  6. Marcelin R (1915) Ann Phys 3:120–231

    CAS  Google Scholar 

  7. Marcelin R (1914) C R Hebd Séances Acad Sci 158:116–118

    CAS  Google Scholar 

  8. Marcelin R (1914) C R Hebd Séances Acad Sci 158:407–409

    CAS  Google Scholar 

  9. Gibbs JW (1902) Elementary principles in statistical mechanics. C. Scribner, New York

    Google Scholar 

  10. Navarro L (1998) Arch Hist Exact Sci 53:147–180

    Article  Google Scholar 

  11. Duhem P (1911) Traité d’énergétique, 2 vols. Gauthier-Villars, Paris

    Google Scholar 

  12. Pelzer H, Wigner EP (1932) Z Phys Chem Abt B 15:445–471

    Google Scholar 

  13. Wigner EP (1938) Trans Faraday Soc 34:29–41

    Article  CAS  Google Scholar 

  14. Laidler KJ, King MC (1983) J Phys Chem 87:2657–2664

    Article  CAS  Google Scholar 

  15. Waalkens H, Schubert R, Wiggins S (2008) Nonlinearity 21:R1–R118

    Article  Google Scholar 

  16. Earnshaw S (1842) Trans Camb Philos Soc 7:97–112

    Google Scholar 

  17. Bohr N (1922) The structure of the atom. Nobel Lecture, December 11. Available online as http://www.nobelprize.org/nobel_prizes/physics/laureates/1922/Bohr-lecture.html

  18. Bohr N (1913) Philos Mag 26:1–25

    CAS  Google Scholar 

  19. Schwarzschild K (1916) Sitzber Preuss Akad Wiss 1:548–561

    Google Scholar 

  20. Heurlinger T (1919) Z Phys 20:188–190

    CAS  Google Scholar 

  21. Born M (1925) Vorlesungen über Atommechanik. Springer, Berlin. The mechanics of the atom. Translated by JW Fisher (1927). George Bell and Sons, London

    Google Scholar 

  22. Sommerfeld A (1919) Atombau und Spektrallinien. Verlag F Vieweg und Sohn, Braunschweig

    Google Scholar 

  23. Ehrenfest P (1916) Ann Phys 356:327–352

    Article  Google Scholar 

  24. Bohr N (1918) Kgl Danske Vid Selskab 4:1–36

    CAS  Google Scholar 

  25. Bohr N (1923) Z Phys 13:117–165

    Article  CAS  Google Scholar 

  26. Berry MV (1984) J Phys A, Math Gen 17:1225–1233

    Article  CAS  Google Scholar 

  27. Gutzwiller M (1998) Rev Mod Phys 70:589–639

    Article  Google Scholar 

  28. Percival IC (1973) J Phys B, Atom Mol Phys 6:L229–L232

    Article  Google Scholar 

  29. Thirring W (1987) In: Kilmister CW (ed) Schrödinger, centenary celebrations of a polymath. Cambridge University Press, Cambridge

    Google Scholar 

  30. Poincaré H (1899) Les méthodes nouvelles de la mécanique céleste. Gauthier-Villars, Paris

    Google Scholar 

  31. Einstein A (1917) Verh Dtsch Phys Ges 19:82–92

    Google Scholar 

  32. Pauli W (1922) Ann Phys 68:177–240

    Article  CAS  Google Scholar 

  33. Nordheim L (1923) Z Phys 19:69–93

    Article  CAS  Google Scholar 

  34. Kemble EC (1926) Molecular spectra of gases. Bull NRC 11(57):303–304

    Google Scholar 

  35. Kragh H (2012) Neils Bohr and the quantum atom: the Bohr model of atomic structure, 1913–1925. Oxford University Press, Oxford, p 239

    Book  Google Scholar 

  36. Born M, Heisenberg W (1924) Ann Phys 74:1–31

    Article  CAS  Google Scholar 

  37. Eckart C (1935) Phys Rev 47:552–558

    Article  CAS  Google Scholar 

  38. Born M, Oppenheimer JR (1927) Ann Phys 84:457–484

    Article  CAS  Google Scholar 

  39. Heitler W, London F (1927) Z Phys 44:455–472

    Article  CAS  Google Scholar 

  40. London F (1928) In: Quantentheorie und Chemie. Hirzel, Leipzig, p 59

    Google Scholar 

  41. London F (1928) In: Probleme der modernen Physik. Hirzel, Leipzig, p 104

    Google Scholar 

  42. Slater JC (1927) Proc Natl Acad Sci 13:423–430

    Article  CAS  Google Scholar 

  43. Born M, Fock V (1928) Z Phys 51:165–180

    Article  CAS  Google Scholar 

  44. Born M, Huang K (1954) Dynamical theory of crystal lattices. Clarendon, Oxford

    Google Scholar 

  45. Lathouwers L, van Leuven P (1982) Adv Chem Phys 49:115–189

    Article  CAS  Google Scholar 

  46. An English language translation of the original paper can be found at www.ulb.ac.be/cpm/people/scientists/bsutclif/main.html

  47. Sutcliffe BT (1990) In: Maksic ZB (ed) Theoretical models of chemical bonding, part 1. Springer, Berlin, pp 1–28

    Google Scholar 

  48. Pauling L, Wilson EB Jr. (1935) Introduction to quantum mechanics. McGraw-Hill, New York. See especially Chap X, p 262

    Google Scholar 

  49. Born M (1951) Nachr. Akad. Wiss. Göttingen. Math-Phys. Klasse IIa. Math Phys Chem Abt, S Art Nr 6:1–3

    Google Scholar 

  50. O’Malley TF (1971) Adv At Mol Phys 7:223–249

    Article  Google Scholar 

  51. Hall GG (1987) Int J Quant Chem XXXI:383–391

    Article  Google Scholar 

  52. Loss M, Miyao T, Spohn H (2007) J Funct Anal 243:353–393

    Article  Google Scholar 

  53. Reed M, Simon B (1978) Analysis of operators. Methods of modern mathematical physics, vol IV. Academic Press, New York

    Google Scholar 

  54. Kato T (1951) Trans Am Math Soc 70:195–218

    Google Scholar 

  55. Frolov AM (1999) Phys Rev A 59:4270–4280

    Article  CAS  Google Scholar 

  56. Messiah A (1961) Quantum mechanics. North-Holland, Amsterdam

    Google Scholar 

  57. Bastiaan JB, Bowman JM (2009) Int Rev Phys Chem 28:577–606

    Article  Google Scholar 

  58. Broeckhove J, Lathouwers L, van Leuven P (1991) J Math Chem 6:207–241

    Article  CAS  Google Scholar 

  59. Mead CA (1992) Rev Mod Phys 64:51–85

    Article  CAS  Google Scholar 

  60. Faure F, Zhilinskii B (2002) Phys Lett A 302:242–252

    Article  CAS  Google Scholar 

  61. Combes J-M (1975) In: International symposium on mathematical physics, Kyoto University, Kyoto. Lecture notes in physics, vol 39, pp 467–471

    Google Scholar 

  62. Combes J-M (1977) Acta Phys Austriaca Suppl XVII:139–159

    Google Scholar 

  63. Combes J-M, Seiler R (1980) Quantum dynamics of molecules. In: Woolley RG (ed) NATO ASI B57. Plenum, New York

    Google Scholar 

  64. Combes J-M, Duclos P, Seiler R (1981) In: Velo G, Wightman A (eds) Rigorous atomic and molecular physics. Plenum, New York

    Google Scholar 

  65. Löwdin P-O (1966) Perturbation theory and its application in quantum mechanics. In: Wilcox CH (ed) Proceedings of Madison symposium. Wiley, New York

    Google Scholar 

  66. Zhislin G (1960) Tr Mosc Mat Obsc 9:81–128

    Google Scholar 

  67. van Winter C (1964) Kgl Danske Vid Selskab 1:1–60

    Google Scholar 

  68. Hunziker W (1966) Helv Phys Acta 39:451–462

    CAS  Google Scholar 

  69. Hagedorn GA, Joye A (2007) In: Gesztezy F, Deift P, Galvez C, Perry P, Schlag GW (eds) Spectral theory and mathematical physics. A festschrift in honor of Barry Simon’s 60th birthday. Oxford University Press, London, p 203

    Chapter  Google Scholar 

  70. Klein M, Martinez A, Seiler R, Wang XP (1992) Commun Math Phys 143:607–639

    Article  Google Scholar 

  71. Feffermann CL (1983) Bull Am Math Soc 9:129–206

    Article  Google Scholar 

  72. Hagedorn GA, Joye A (1999) Rev Math Phys 11:41–101

    Article  Google Scholar 

  73. Primas H (1983) Chemistry, quantum mechanics and reductionism, 2nd edn. Springer, Berlin

    Book  Google Scholar 

  74. Woolley RG (1980) Isr J Chem 19:30–46

    CAS  Google Scholar 

  75. Woolley RG (1982) Struct Bond (Springer) 52:1–35

    Article  CAS  Google Scholar 

  76. Anderson PW (1984) Basic notions of condensed matter physics. Benjamin-Cummings, Redwood City

    Google Scholar 

  77. Lorentz HA (1909) The theory of electrons and its applications to the phenomena of light and radiant heat. Teubner, Leipzig

    Google Scholar 

  78. Lieb EH, Loss M (2003) Adv Theor Math Phys 7:667–710

    Article  Google Scholar 

  79. Lieb EH, Loss M (2005) Commun Math Phys 258:675–695

    Article  Google Scholar 

  80. Hasler D, Herbst I (2008) Commun Math Phys 279:769–787

    Article  Google Scholar 

  81. Loss M, Miyao T, Spohn H (2009) Lett Math Phys 89:21–31

    Article  CAS  Google Scholar 

  82. Lieb EH, Seiringer R (2010) The stability of matter in quantum mechanics. Cambridge University Press, Cambridge

    Google Scholar 

  83. Hunter G (1975) Int J Quant Chem 9:237–242

    Article  CAS  Google Scholar 

  84. Abedi A, Matra NT, Gross EKU (2010) Phys Rev Lett 105:123002

    Article  Google Scholar 

  85. Kutzelnigg W (2007) Mol Phys 105:2627–2647

    Article  CAS  Google Scholar 

  86. Hunter G (1981) Int J Quant Chem 19:755–761

    Article  CAS  Google Scholar 

  87. Czub J, Wolniewicz L (1978) Mol Phys 36:1301–1305

    Article  CAS  Google Scholar 

  88. Wilson EB Jr. (1979) Int J Quantum Chem, Symp 13:5–14

    CAS  Google Scholar 

  89. Cassam-Chenai P (2006) Chem Phys Lett 420:354–357

    Article  CAS  Google Scholar 

  90. Pachuki K, Komasa J (2009) J Chem Phys 130:164113

    Article  Google Scholar 

  91. Henderson JR, Tennyson J, Sutcliffe BT (1993) J Chem Phys 98:7191–7203

    Article  CAS  Google Scholar 

  92. Miller S, Tennyson J (1988) Chem Phys Lett 145:117–120

    Article  CAS  Google Scholar 

  93. Sutcliffe BT, Woolley RG (2005) Phys Chem Chem Phys 7:3664–3676

    Article  CAS  Google Scholar 

  94. Dixmier J (1981) Von Neumann algebras. Elsevier/North-Holland, Amsterdam

    Google Scholar 

  95. Ballentine LE (1990) Quantum mechanics. Prentice-Hall, Englewood Cliffs

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Brian Sutcliffe .

Editor information

Editors and Affiliations

Appendix

Appendix

In this appendix we give an heuristic account of the mathematical notion of the direct integral of Hilbert spaces, and then study a simple model problem to illustrate the general ideas discussed in the paper.

Consider a self-adjoint operator T that depends on a parameter X, so T=T(X). The parameter X=−∞≤X≤+∞, covers the whole real line , and T(X) is assumed to be defined for all X. T(X) is an operator on a Hilbert space, which is denoted ; its eigenfunctions {ϕ} defined by

$$T(X) \phi(X)_k=\lambda_k(X) \phi(X)_k, $$

form a complete orthogonal set for the space . The scalar product is

$$ \bigl\langle\phi(X)_{k} \big|\phi(X)_{j} \bigr \rangle_{X}=\int\phi(X,x)_{k}^{*}\phi (X,x)_{j} dx=f(X)_{kj} \equiv f(X) \delta_{kj},\quad f(X)<\infty $$
(1.51)

T may have discrete eigenvalues below a continuous spectrum that starts at Λ, that is

$$\sigma(X)=\sigma\bigl(T(X)\bigr)=\bigl[\lambda_0(X), \lambda_1(X),\ldots ,\lambda_k(X) \bigr) \bigcup \bigl[ \varLambda(X),\infty \bigr). $$

Now let’s introduce a ‘big’ Hilbert space as a direct integral over the {},

(1.52)

and correspondingly the operator acting on defined by

The scalar product on the big space is defined explicitly in terms of (1.51) by

(1.53)

In (1.51) one can always chose the functions {ϕ k } to be orthonormalized independently of X,

$$f(X)=1. $$

However this choice is not consistent with (1.53), which requires f(X) to decrease sufficiently fast as |X|→∞ for the integral to exist. The mathematical motivation for this construction is this: the cartesian product of the spaces {},

is a field of Hilbert spaces over which has a natural vector space structure. In modern geometric language, the Hilbert space is a fibre over a point X in a fibre bundle; is the vector space of sections of this bundle. The subspace of consisting of square integrable sections is the direct integral (1.52). The direct integral is the generalization to the continuous infinite dimensional case of the notion of the direct sum of finite dimensional vector spaces.

Purely as a heuristic explanation suppose initially that the parameter X has only two discrete values {X 1,X 2}. There are then two eigenvalue equations to consider, and two associated Hilbert spaces. In the direct sum space we have

The eigenfunctions of are then two-dimensional column vectors, and so

The spectrum of is the union of the spectra of T(X 1) and T(X 2). This discussion is trivially extended to n points {X k :k=1,…,n}, with the spectrum of given by the union of the n operators T(X 1),…,T(X n ). The limit n→∞ is not trivial since it brings in important notions from topology and integration (measure theory) which we gloss over [94]. When these are taken into consideration however the result is that the spectrum σ of is purely continuous since its direct integral representation implies that its spectrum is the union of the spectra of the infinite set of T(X) operators,

(1.54)

where L 0 is the minimum value of λ 0(X). The eigenvalue equation for is,

with Λ a continuous index for the {Φ}. Even if T(X) is self-adjoint, it doesn’t follow that its direct integral is self-adjoint; that depends on specifics and has to be investigated. So the {Φ} cannot be assumed to be complete.

In the application of this mathematics to the Born-Oppenheimer approximation, the role of x is taken by the electronic coordinates t e, and X is to be identified with definite choices of the nuclear coordinates b. If there are M nuclei the parameters b are elements of . The operator T(X) is the clamped-nuclei Hamiltonian \(\hat{\mathsf{K}}(\mathbf{b},\hat{\mathbf{t}}^{\mathrm{e}})_{o}=\hat{\mathsf{K}}_{o}\). With the conventional normalization of clamped-nuclei electronic eigenfunctions independent of the nuclear positions b, the formal eigenvectors, (1.36), [56] of \(\hat{\mathsf{H}}^{\mathrm{elec}}\) do not belong to the Hilbert space ; this simply reflects the loose use of the Dirac delta function for the position operator eigenfunctions.

We now consider a concrete model consisting of coupled harmonic oscillators with two degrees of freedom; we try to mimic the steps taken in the usual Born-Oppenheimer discussion. Consider the following Hamiltonian where κ and a are dimensionless constantsFootnote 20

$$ \hat{\mathsf{H}}=\hat{\mathsf{p}}^2+\kappa^4\hat{ \mathsf{P}}^2+\hat{\mathsf{x}}^2+\hat{ \mathsf{X}}^2+a\hat{\mathsf{x}}\hat{\mathsf{X}}. $$
(1.55)

The only non-zero commutators of the operators are

$$[\hat{\mathsf{x}},\hat{\mathsf{p}}]=i ,\qquad [\hat{\mathsf{X}},\hat{ \mathsf{P}}]=i. $$

Following the conventional discussion of electron-nuclear separation outlined in Sects. 1.3.1, 1.3.2, define

$$\begin{aligned} \hat{\mathsf{H}}_o =&\hat{\mathsf{p}}^2+\hat{ \mathsf{x}}^2+a\hat{\mathsf{x}}\hat{\mathsf{X}}+\hat{ \mathsf{X}}^2 \end{aligned}$$
(1.56)
$$\begin{aligned} \hat{\mathsf{H}}_1 =&\hat{\mathsf{P}}^2 \end{aligned}$$
(1.57)

so that

$$ \hat{\mathsf{H}}=\hat{\mathsf{H}}_o+\kappa^4 \hat{ \mathsf{H}}_1 $$
(1.58)

with Schrödinger equation

$$ (\hat{\mathsf{H}}-E )\varPsi=0. $$
(1.59)

We note that

$$ [\hat{\mathsf{H}}_o,\hat{\mathsf{H}}_1] \neq 0 $$
(1.60)

so the two parts cannot be simultaneously diagonalized. A principal axis transformation of the whole expression \(\hat{\mathsf{H}}\) brings it to separable form, but we do not need to pursue explicitly the full solution here.

On the other hand

$$ [\hat{\mathsf{H}}_o,\hat{\mathsf{X}}]=0 $$
(1.61)

so these two operators may be simultaneously diagonalized, and consider \(\hat{\mathsf{H}}_{o}\) at a definite eigenvalue of \(\hat{\mathsf{X}}\), say X

$$ \hat{\mathsf{K}}_o=\hat{\mathsf{p}}^2+\hat{ \mathsf{x}}^2+a\hat{\mathsf{x}} {X}+{X}^2. $$
(1.62)

This is the Hamiltonian (in the variables \(\hat{\mathsf{x}},\hat{\mathsf{p}}\)) of a displaced oscillator in which X is a (c-number) parameter, with Schrödinger equation in position representation

$$ \hat{\mathsf{K}}_o \varphi(x,X)_n=E_n(X) \varphi(x,X)_n $$
(1.63)

\(\hat{\mathsf{K}}_{o}\) is the analogue in this model of the ‘clamped-nuclei’ electronic Hamiltonian.

The solution is immediate; we make a unitary transformation by introducing a displaced coordinate involving X

$$\begin{aligned} \hat{\mathsf{x}}' =&\hat{\mathsf{x}}-\frac{1}{2}aX,\qquad \hat{\mathsf{p}}'=\hat{\mathsf{p}} \end{aligned}$$
(1.64)
$$\begin{aligned} \hat{\mathsf{U}} =&e^{iaX\hat{\mathsf{p}}/2},\qquad \hat{\mathsf{K}}_{o}^{\prime}= \hat{\mathsf{U}}^{-1}\hat{\mathsf{K}}_0 \hat{\mathsf{U}} \end{aligned}$$
(1.65)

so that the transformed \(\hat{\mathsf{K}}_{o}\) in the new variables is

$$\hat{\mathsf{K}}'_{o}=\hat{\mathsf{p}}'{}^{2}+ \hat{\mathsf{x}}'{}^{2} +\biggl(1-\frac{a^2}{4} \biggr)X^2 $$

which has oscillator eigenvalues and eigenfunctions

$$ E_n(X)=E_n^0+\biggl(1-\frac{a^2}{4} \biggr)X^2,\qquad \varphi\bigl(x'\bigr)_n $$
(1.66)

where \(x'=x-\frac{aX}{2}\), the {φ n } are the usual harmonic oscillator eigenfunctions and \(E_{n}^{0}\) is the energy of the free oscillator \(2(n+\frac{1}{2})\). For fixed n the spectrum σ(X) is discrete and, as a function of the X parameters, would be conventionally interpreted as a ‘potential energy curve’. As far as (1.58) is concerned, \(\hat{\mathsf{K}}_{o}\), evaluated at some point X 0 cannot be regarded as an ‘approximation’ to \(\hat{\mathsf{H}}_{o}\), since obviously

$$\bigl[\hat{\mathsf{K}}_{o}(X_0),\hat{ \mathsf{H}}_1\bigr]=0 $$

so they can be simultaneously diagonalized, and \(\hat{\mathsf{H}}_{1}\) has purely continuous spectrum (free motion). So we have to consider X in the full problem in its operator form, \(\hat{\mathsf{X}}\).

We make the same unitary transformation of the \(\hat{\mathsf{x}},\hat{\mathsf{p}}\) variables in \(\hat{\mathsf{H}}_{o}\) as before, and it is still brought to diagonal form; however \(\hat{\mathsf{H}}_{1}\) will be modified because \(\hat{\mathsf{P}}\) is also translated by the operator \(\hat{\mathsf{U}}\) in (1.65) that generates the coordinate displacement (cf. (1.60)); thus

$$\hat{\mathsf{U}}^{-1} \hat{\mathsf{P}} \hat{\mathsf{U}}=\hat{ \mathsf{P}}+\frac{1}{2}a\hat{\mathsf{p}} $$

so the transform of \(\hat{\mathsf{H}}_{1}\) contains linear and quadratic terms in \(\hat{\mathsf{p}}\).

Nevertheless (1.61) is still valid, and formally we may write \(\hat{\mathsf{H}}_{0}\) as a direct integral

$$ \hat{\mathsf{H}}_o=\int^{\oplus}_{X} \hat{\mathsf{K}}(\hat{\mathsf{x}},\hat{\mathsf{p}},X)_o dX. $$
(1.67)

The Schrödinger equation for \(\hat{\mathsf{H}}_{o}\) in position representation is now an equation involving functions of x and X

$$ \hat{\mathsf{H}}_o \varPhi_{\varepsilon}\bigl(x',X \bigr)=\varepsilon \varPhi_{\varepsilon}\bigl(x',X\bigr). $$
(1.68)

Just as before (see (1.54)) the direct integral decomposition (1.67) implies that the spectrum is purely continuous, explicitly

$$ \sigma(\hat{\mathsf{H}}_o)=\bigcup_{X} \sigma(X)= [1,\infty ). $$
(1.69)

ε in (1.68) takes all positive values ≥1, where 1 is the minimum value of the oscillator eigenvalue \(2(n+\frac{1}{2})\). The associated continuum eigenfunctions {Φ} may formally be written as products of oscillator eigenfunctions (in x′), and delta functions (in X). They don’t lie in Hilbert space of course and one needs the Gel’fand construction of a rigged space to make sense of the formal calculation [95]. If one returns to the x variable, the {φ n } are functions of x and X, since x′=x′(X).

One can’t do anything very useful with the direct integral expression (1.67) for \(\hat{\mathsf{H}}_{o}\) apart from adding it onto the κ 4 term, which just returns us to the full problem. The full wavefunction in (1.59) can be expanded as

(1.70)
$$\begin{aligned} =&\sum_n c(X)_n \varphi \bigl(x'(X) \bigr)_n \end{aligned}$$
(1.71)

which obviously leads towards a variational approach [56]; such expansions rely on the completeness of the states employed. In this simple problem there is no difficulty, but as noted earlier, in realistic Coulomb systems it is much less clear that a complete set of states is available.

However that may be, let us rehearse again the argument due to Born summarized in Sect. 1.3.2. We substitute (1.71) in (1.59), left multiply by \(\varphi_{m}^{*}\) and integrate out the x′ variables to leave an equation for the coefficients {c(X) n },

$$\begin{aligned} &\int dx' \varphi \bigl(x'(X) \bigr)_{m}^{*} \bigl[\hat{\mathsf{H}}_o+\kappa^4 \hat{ \mathsf{H}}_1 \bigr] \sum_{n}c(X)_n \varphi \bigl(x'(X) \bigr)_n \\ &\quad= E \int dx' \varphi \bigl(x'(X) \bigr)_{m}^{*} \sum_{n}c(X)_n \varphi \bigl(x'(X) \bigr)_n. \end{aligned}$$
(1.72)

At this point in the conventional account, \(\hat{\mathsf{H}}_{o}\) is replaced by \(\hat{\mathsf{K}}_{o}\), (1.62), and then the action of \(\hat{\mathsf{K}}_{o}\) on the functions {φ} in (1.72) can be evaluated using (1.63) in the well-known way,

$$\hat{\mathsf{H}}_o \varphi \bigl(x'(X) \bigr)_n \rightarrow \hat{\mathsf{K}}_o \varphi \bigl(x'(X) \bigr)_n=\varepsilon(X)_n \varphi \bigl(x'(X) \bigr)_n. $$

From the foregoing discussion it is clear that the substitution of \(\hat{\mathsf{H}}_{o}\) by \(\hat{\mathsf{K}}_{o}\) makes a qualitative change in the theory. This change does seem to be the ‘right’ thing to do, but so far there is no explanation as to why this is so.

Rights and permissions

Reprints and permissions

Copyright information

© 2013 Springer International Publishing Switzerland

About this paper

Cite this paper

Sutcliffe, B., Woolley, R.G. (2013). The Potential Energy Surface in Molecular Quantum Mechanics. In: Hotokka, M., Brändas, E., Maruani, J., Delgado-Barrio, G. (eds) Advances in Quantum Methods and Applications in Chemistry, Physics, and Biology. Progress in Theoretical Chemistry and Physics, vol 27. Springer, Cham. https://doi.org/10.1007/978-3-319-01529-3_1

Download citation

Publish with us

Policies and ethics