Born–Oppenheimer
approximation
This article has multiple issues. Please help
improve it or discuss these issues on the talk more
Learn
In quantum chemistry and molecular
physics, the Born–Oppenheimer (BO)
approximation is the assumption that the
motion of atomic nuclei and electrons in a
molecule can be separated. The approach
is named after Max Born, and J. Robert
Oppenheimer. In mathematical terms, it
allows the wavefunction of a molecule to
be broken into its electronic and nuclear
(vibrational, rotational) components.
Computation of the energy and the
wavefunction of an average-size molecule
is simplified by the approximation. For
example, the benzene molecule consists
of 12 nuclei and 42 electrons. The time
independent Schrödinger equation, which
must be solved to obtain the energy and
wavefunction of this molecule, is a partial
differential eigenvalue equation in 162
variables—the spatial coordinates of the
electrons and the nuclei. The BO
approximation makes it possible to
compute the wavefunction in two less
complicated consecutive steps. This
approximation was proposed in 1927, in
the early period of quantum mechanics, by
Born and Oppenheimer and is still
indispensable in quantum chemistry.
In the first step of the BO approximation
the electronic Schrödinger equation is
solved, yielding the wavefunction
depending on electrons only.
For benzene this wavefunction depends on
126 electronic coordinates. During this
solution the nuclei are fixed in a certain
configuration, very often the equilibrium
configuration. If the effects of the
quantum mechanical nuclear motion are
to be studied, for instance because a
vibrational spectrum is required, this
electronic computation must be in nuclear
coordinates. In the second step of the BO
approximation this function serves as a
potential in a Schrödinger equation
containing only the nuclei—for benzene an
equation in 36 variables.
The success of the BO approximation is
due to the difference between nuclear and
electronic masses. The approximation is
an important tool of quantum chemistry:
all computations of molecular
wavefunctions for large molecules make
use of it, and without it only the lightest
molecule, H2, can be handled. Even in the
cases where the BO approximation breaks
down, it is used as a point of departure for
the computations.
The electronic energies consist of kinetic
energies, interelectronic repulsions,
internuclear repulsions, and electron–
nuclear attractions. In accord with the
Hellmann-Feynman theorem, the nuclear
potential is taken to be an average over
electron configurations of the sum of the
electron–nuclear and internuclear electric
potentials.
In molecular spectroscopy, because the
ratios of the periods of the electronic,
vibrational and rotational energies are
each related to each other on scales in the
order of a thousand, the Born–
Oppenheimer name has also been
attached to the approximation where the
energy components are treated separately.
The nuclear spin energy is so small that it
is normally omitted.
Short description
The Born–Oppenheimer (BO)
approximation is ubiquitous in quantum
chemical calculations of molecular
wavefunctions. It consists of two steps.
In the first step the nuclear kinetic energy
is neglected,[1] that is, the corresponding
operator Tn is subtracted from the total
molecular Hamiltonian. In the remaining
electronic Hamiltonian He the nuclear
positions enter as parameters. The
electron–nucleus interactions are not
removed, and the electrons still "feel" the
Coulomb potential of the nuclei clamped
at certain positions in space. (This first
step of the BO approximation is therefore
often referred to as the clamped-nuclei
approximation.)
The electronic Schrödinger equation
is solved (out of necessity, approximately).
The quantity r stands for all electronic
coordinates and R for all nuclear
coordinates. The electronic energy
eigenvalue Ee depends on the chosen
positions R of the nuclei. Varying these
positions R in small steps and repeatedly
solving the electronic Schrödinger
equation, one obtains Ee as a function of
R. This is the potential energy surface
(PES): Ee(R) . Because this procedure of
recomputing the electronic wave functions
as a function of an infinitesimally
changing nuclear geometry is reminiscent
of the conditions for the adiabatic
theorem, this manner of obtaining a PES is
often referred to as the adiabatic
approximation and the PES itself is called
an adiabatic surface.[2]
In the second step of the BO
approximation the nuclear kinetic energy
Tn (containing partial derivatives with
respect to the components of R) is
reintroduced, and the Schrödinger
equation for the nuclear motion[3]
is solved. This second step of the BO
approximation involves separation of
vibrational, translational, and rotational
motions. This can be achieved by
application of the Eckart conditions. The
eigenvalue E is the total energy of the
molecule, including contributions from
electrons, nuclear vibrations, and overall
rotation and translation of the molecule.
Derivation
It will be discussed how the BO
approximation may be derived and under
which conditions it is applicable. At the
same time we will show how the BO
approximation may be improved by
including vibronic coupling. To that end the
second step of the BO approximation is
generalized to a set of coupled eigenvalue
equations depending on nuclear
coordinates only. Off-diagonal elements in
these equations are shown to be nuclear
kinetic energy terms.
It will be shown that the BO approximation
can be trusted whenever the PESs,
obtained from the solution of the
electronic Schrödinger equation, are well
separated:
.
We start from the exact non-relativistic,
time-independent molecular Hamiltonian:
with
The position vectors of the
electrons and the position vectors
of
the nuclei are with respect to a Cartesian
inertial frame. Distances between particles
are written as (distance
between electron i and nucleus A) and
similar definitions hold for and .
We assume that the molecule is in a
homogeneous (no external force) and
isotropic (no external torque) space. The
only interactions are the two-body
Coulomb interactions among the electrons
and nuclei. The Hamiltonian is expressed
in atomic units, so that we do not see
Planck's constant, the dielectric constant
of the vacuum, electronic charge, or
electronic mass in this formula. The only
constants explicitly entering the formula
are ZA and MA – the atomic number and
mass of nucleus A.
It is useful to introduce the total nuclear
momentum and to rewrite the nuclear
kinetic energy operator as follows:
Suppose we have K electronic
eigenfunctions of , that is,
we have solved
The electronic wave functions will be
taken to be real, which is possible when
there are no magnetic or spin interactions.
The parametric dependence of the
functions on the nuclear coordinates is
indicated by the symbol after the
semicolon. This indicates that, although
is a real-valued function of , its
functional form depends on .
For example, in the molecular-orbital-
linear-combination-of-atomic-orbitals
(LCAO-MO) approximation, is a
molecular orbital (MO) given as a linear
expansion of atomic orbitals (AOs). An AO
depends visibly on the coordinates of an
electron, but the nuclear coordinates are
not explicit in the MO. However, upon
change of geometry, i.e., change of , the
LCAO coefficients obtain different values
and we see corresponding changes in the
functional form of the MO .
We will assume that the parametric
dependence is continuous and
differentiable, so that it is meaningful to
consider
which in general will not be zero.
The total wave function is
expanded in terms of :
with
and where the subscript indicates that
the integration, implied by the bra–ket
notation, is over electronic coordinates
only. By definition, the matrix with general
element
is diagonal. After multiplication by the real
function from the left and
integration over the electronic coordinates
the total Schrödinger equation
is turned into a set of K coupled
eigenvalue equations depending on
nuclear coordinates only
The column vector has elements
. The matrix
is diagonal, and the nuclear
Hamilton matrix is non-diagonal; its off-
diagonal (vibronic coupling) terms
are further discussed below.
The vibronic coupling in this approach is
through nuclear kinetic energy terms.
Solution of these coupled equations gives
an approximation for energy and
wavefunction that goes beyond the Born–
Oppenheimer approximation.
Unfortunately, the off-diagonal kinetic
energy terms are usually difficult to
handle. This is why often a diabatic
transformation is applied, which retains
part of the nuclear kinetic energy terms on
the diagonal, removes the kinetic energy
terms from the off-diagonal and creates
coupling terms between the adiabatic
PESs on the off-diagonal.
If we can neglect the off-diagonal
elements the equations will uncouple and
simplify drastically. In order to show when
this neglect is justified, we suppress the
coordinates in the notation and write, by
applying the Leibniz rule for differentiation,
the matrix elements of as
The diagonal ( ) matrix elements
of the operator
vanish, because we assume time-reversal
invariant, so can be chosen to be
always real. The off-diagonal matrix
elements satisfy
The matrix element in the numerator is
The matrix element of the one-electron
operator appearing on the right side is
finite.
When the two surfaces come close,
, the nuclear
momentum coupling term becomes large
and is no longer negligible. This is the
case where the BO approximation breaks
down, and a coupled set of nuclear motion
equations must be considered instead of
the one equation appearing in the second
step of the BO approximation.
Conversely, if all surfaces are well
separated, all off-diagonal terms can be
neglected, and hence the whole matrix of
is effectively zero. The third term on
the right side of the expression for the
matrix element of Tn (the Born–
Oppenheimer diagonal correction) can
approximately be written as the matrix of
squared and, accordingly, is then
negligible also. Only the first (diagonal)
kinetic energy term in this equation
survives in the case of well separated
surfaces, and a diagonal, uncoupled, set of
nuclear motion equations results:
which are the normal second step of the
BO equations discussed above.
We reiterate that when two or more
potential energy surfaces approach each
other, or even cross, the Born–
Oppenheimer approximation breaks down,
and one must fall back on the coupled
equations. Usually one invokes then the
diabatic approximation.
The Born–Oppenheimer
approximation with the
correct symmetry
To include the correct symmetry within the
Born–Oppenheimer (BO)
approximation,[4][5] a molecular system
presented in terms of (mass-dependent)
nuclear coordinates and formed by the
two lowest BO adiabatic potential energy
surfaces (PES) and is
considered. To ensure the validity of the
BO approximation, the energy E of the
system is assumed to be low enough so
that becomes a closed PES in the
region of interest, with the exception of
sporadic infinitesimal sites surrounding
degeneracy points formed by and
(designated as (1, 2) degeneracy
points).
The starting point is the nuclear adiabatic
BO (matrix) equation written in the form[6]
where is a column vector containing
the unknown nuclear wave functions
, is a diagonal matrix
containing the corresponding adiabatic
potential energy surfaces , m is the
reduced mass of the nuclei, E is the total
energy of the system, is the gradient
operator with respect to the nuclear
coordinates , and is a matrix
containing the vectorial non-adiabatic
coupling terms (NACT):
Here are eigenfunctions of the
electronic Hamiltonian assumed to form a
complete Hilbert space in the given region
in configuration space.
To study the scattering process taking
place on the two lowest surfaces, one
extracts from the above BO equation the
two corresponding equations:
where
(k = 1, 2), and is the
(vectorial) NACT responsible for the
coupling between and .
Next a new function is introduced:[7]
and the corresponding rearrangements are
made:
1. Multiplying the second equation by i and
combining it with the first equation yields
the (complex) equation
2. The last term in this equation can be
deleted for the following reasons: At those
points where is classically closed,
by definition, and at those
points where becomes classically
allowed (which happens at the vicinity of
the (1, 2) degeneracy points) this implies
that: , or
. Consequently, the
last term is, indeed, negligibly small at
every point in the region of interest, and
the equation simplifies to become
In order for this equation to yield a solution
with the correct symmetry, it is suggested
to apply a perturbation approach based on
an elastic potential , which
coincides with at the asymptotic
region.
The equation with an elastic potential can
be solved, in a straightforward manner, by
substitution. Thus, if is the solution of
this equation, it is presented as
where is an arbitrary contour, and the
exponential function contains the relevant
symmetry as created while moving along
.
The function can be shown to be a
solution of the (unperturbed/elastic)
equation
Having , the full solution of the
above decoupled equation takes the form
where satisfies the resulting
inhomogeneous equation:
In this equation the inhomogeneity
ensures the symmetry for the perturbed
part of the solution along any contour and
therefore for the solution in the required
region in configuration space.
The relevance of the present approach
was demonstrated while studying a two-
arrangement-channel model (containing
one inelastic channel and one reactive
channel) for which the two adiabatic
states were coupled by a Jahn–Teller
conical intersection.[8][9] A nice fit between
the symmetry-preserved single-state
treatment and the corresponding two-state
treatment was obtained. This applies in
particular to the reactive state-to-state
probabilities (see Table III in Ref. 5a and
Table III in Ref. 5b), for which the ordinary
BO approximation led to erroneous results,
whereas the symmetry-preserving BO
approximation produced the accurate
results, as they followed from solving the
two coupled equations.
See also
Adiabatic ionization
Adiabatic process (quantum
mechanics)
Avoided crossing
Born–Huang approximation
Franck–Condon principle
References
1. This step is often justified by stating
that "the heavy nuclei move more
slowly than the light electrons".
Classically this statement makes
sense only if the momentum p of
electrons and nuclei is of the same
order of magnitude. In that case mn ≫
me implies p2/(2mn) ≪ p2/(2me). It is
easy to show that for two bodies in
circular orbits around their center of
mass (regardless of individual
masses), the momenta of the two
bodies are equal and opposite, and
that for any collection of particles in
the center-of-mass frame, the net
momentum is zero. Given that the
center-of-mass frame is the lab frame
(where the molecule is stationary), the
momentum of the nuclei must be
equal and opposite to that of the
electrons. A hand-waving justification
can be derived from quantum
mechanics as well. Recall that the
corresponding operators do not
contain mass and think of the
molecule as a box containing the
electrons and nuclei and see particle
in a box. Since the kinetic energy is
p2/(2m), it follows that, indeed, the
kinetic energy of the nuclei in a
molecule is usually much smaller than
the kinetic energy of the electrons, the
mass ratio being on the order of 104).
2. It is assumed, in accordance with the
adiabatic theorem, that the same
electronic state (for instance, the
electronic ground state) is obtained
upon small changes of the nuclear
geometry. The method would give a
discontinuity (jump) in the PES if
electronic state switching would occur.
3. This equation is time-independent, and
stationary wavefunctions for the nuclei
are obtained; nevertheless, it is
traditional to use the word "motion" in
this context, although classically
motion implies time dependence.
4. Max Born; J. Robert Oppenheimer
(1927). "Zur Quantentheorie der
Molekeln" [On the Quantum Theory of
Molecules]. Annalen der Physik (in
German). 389 (20): 457–484.
Bibcode:1927AnP...389..457B .
doi:10.1002/andp.19273892002 .
5. M. Born and K. Huang, Dynamical
Theory of Crystal Lattices, 1954
(Oxford University Press, New York),
Chapter IV.
6. M. Baer, Beyond Born–Oppenheimer:
Electronic non-Adiabatic Coupling
Terms and Conical Intersections, 2006
(Wiley and Sons, Inc., Hoboken, N.J.),
Chapter 2.
7. M. Baer and R. Englman, Chem. Phys.
Lett. 265, 105 (1997).
8. (a) R. Baer, D. M. Charutz, R. Kosloff
and M. Baer, J. Chem. Phys. 111, 9141
(1996); (b) S. Adhikari and G. D. Billing,
J. Chem. Phys. 111, 40 (1999).
9. D. M. Charutz, R. Baer and M. Baer,
Chem. Phys. Lett. 265, 629 (1996).
External links
Resources related to the Born–
Oppenheimer approximation:
The original article (in German)
Translation by S. M. Blinder
The Born–Oppenheimer approximation ,
a section from Peter Haynes' doctoral
thesis
Retrieved from
"https://en.wikipedia.org/w/index.php?title=Born–
Oppenheimer_approximation&oldid=878560794"
Last edited 4 months ago by an ano…
Content is available under CC BY-SA 3.0 unless
otherwise noted.