HOME

TheInfoList



OR:

In
quantum chemistry Quantum chemistry, also called molecular quantum mechanics, is a branch of physical chemistry focused on the application of quantum mechanics to chemical systems, particularly towards the quantum-mechanical calculation of electronic contributions ...
and molecular physics, the Born–Oppenheimer (BO) approximation is the best-known mathematical approximation in molecular dynamics. Specifically, it is the assumption that the
wave function A wave function in quantum physics is a mathematical description of the quantum state of an isolated quantum system. The wave function is a complex-valued probability amplitude, and the probabilities for the possible results of measurements m ...
s of
atomic nuclei The atomic nucleus is the small, dense region consisting of protons and neutrons at the center of an atom, discovered in 1911 by Ernest Rutherford based on the 1909 Geiger–Marsden gold foil experiment. After the discovery of the neutron i ...
and electrons in a molecule can be treated separately, based on the fact that the nuclei are much heavier than the electrons. Due to the larger relative mass of a nucleus compared to an electron, the coordinates of the nuclei in a system are approximated as fixed, while the coordinates of the electrons are dynamic. The approach is named after Max Born and
J. Robert Oppenheimer J. Robert Oppenheimer (; April 22, 1904 – February 18, 1967) was an American theoretical physicist. A professor of physics at the University of California, Berkeley, Oppenheimer was the wartime head of the Los Alamos Laboratory and is often ...
who proposed it in 1927, in the early period of quantum mechanics. The approximation is widely used in quantum chemistry to speed up the computation of molecular wavefunctions and other properties for large molecules. There are cases where the assumption of separable motion no longer holds, which make the approximation lose validity (it is said to "break down"), but even then the approximation is usually used as a starting point for more refined methods. In molecular
infrared spectroscopy Infrared spectroscopy (IR spectroscopy or vibrational spectroscopy) is the measurement of the interaction of infrared radiation with matter by absorption, emission, or reflection. It is used to study and identify chemical substances or function ...
, using the BO approximation means considering molecular energy as a sum of independent terms, e.g.: E_\text = E_\text + E_\text + E_\text + E_\text. These terms are of different orders of magnitude and the nuclear spin energy is so small that it is often omitted. The electronic energies E_\text consist of kinetic energies, interelectronic repulsions, internuclear repulsions, and electron–nuclear attractions, which are the terms typically included when computing the electronic structure of molecules.


Example

The
benzene Benzene is an organic chemical compound with the molecular formula C6H6. The benzene molecule is composed of six carbon atoms joined in a planar ring with one hydrogen atom attached to each. Because it contains only carbon and hydrogen atoms ...
molecule consists of 12 nuclei and 42 electrons. The Schrödinger equation, which must be solved to obtain the energy levels and wavefunction of this molecule, is a partial differential eigenvalue equation in the three-dimensional coordinates of the nuclei and electrons, giving 3 × 12 + 3 × 42 = 36 nuclear + 126 electronic = 162 variables for the wave function. The
computational complexity In computer science, the computational complexity or simply complexity of an algorithm is the amount of resources required to run it. Particular focus is given to computation time (generally measured by the number of needed elementary operations ...
, i.e., the computational power required to solve an eigenvalue equation, increases faster than the square of the number of coordinates. When applying the BO approximation, two smaller, consecutive steps can be used: For a given position of the nuclei, the ''electronic'' Schrödinger equation is solved, while treating the nuclei as stationary (not "coupled" with the dynamics of the electrons). This corresponding
eigenvalue In linear algebra, an eigenvector () or characteristic vector of a linear transformation is a nonzero vector that changes at most by a scalar factor when that linear transformation is applied to it. The corresponding eigenvalue, often denoted ...
problem then consists only of the 126 electronic coordinates. This electronic computation is then repeated for other possible positions of the nuclei, i.e. deformations of the molecule. For benzene, this could be done using a grid of 36 possible nuclear position coordinates. The electronic energies on this grid are then connected to give a potential energy surface for the nuclei. This potential is then used for a second Schrödinger equation containing only the 36 coordinates of the nuclei. So, taking the most optimistic estimate for the complexity, instead of a large equation requiring at least 162^2 = 26\,244 hypothetical calculation steps, a series of smaller calculations requiring 126^2 N = 15\,876 \,N (with ''N'' being the number of grid points for the potential) and a very small calculation requiring 36^2 = 1296 steps can be performed. In practice, the scaling of the problem is larger than n^2, and more approximations are applied in
computational chemistry Computational chemistry is a branch of chemistry that uses computer simulation to assist in solving chemical problems. It uses methods of theoretical chemistry, incorporated into computer programs, to calculate the structures and properties of m ...
to further reduce the number of variables and dimensions. The slope of the potential energy surface can be used to simulate molecular dynamics, using it to express the mean force on the nuclei caused by the electrons and thereby skipping the calculation of the nuclear Schrödinger equation.


Detailed description

The BO approximation recognizes the large difference between the electron mass and the masses of atomic nuclei, and correspondingly the time scales of their motion. Given the same amount of momentum, the nuclei move much more slowly than the electrons. In mathematical terms, the BO approximation consists of expressing the wavefunction (\Psi_\mathrm) of a molecule as the product of an electronic wavefunction and a nuclear ( vibrational,
rotational Rotation, or spin, is the circular movement of an object around a '' central axis''. A two-dimensional rotating object has only one possible central axis and can rotate in either a clockwise or counterclockwise direction. A three-dimensional ...
) wavefunction. \Psi_\mathrm = \psi_\mathrm \psi_\mathrm . This enables a separation of the Hamiltonian operator into electronic and nuclear terms, where cross-terms between electrons and nuclei are neglected, so that the two smaller and decoupled systems can be solved more efficiently. In the first step the nuclear kinetic energy is neglected,Authors often justify this step 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 ''m''n ≫ ''m''e implies ''p''2/(2''m''n) ≪ ''p''2/(2''m''e). 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. The corresponding operators do not contain mass and the molecule can be treated as a box containing the electrons and nuclei. Since the kinetic energy is ''p''2/(2''m''), 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). that is, the corresponding operator ''T''n is subtracted from the total
molecular Hamiltonian In atomic, molecular, and optical physics and quantum chemistry, the molecular Hamiltonian is the Hamiltonian operator representing the energy of the electrons and nuclei in a molecule. This operator and the associated Schrödinger equation pl ...
. In the remaining electronic Hamiltonian ''H''e the nuclear positions are no longer variable, but are constant parameters (they enter the equation "parametrically"). The electron–nucleus interactions are ''not'' removed, i.e., the electrons still "feel" the
Coulomb potential The electric potential (also called the ''electric field potential'', potential drop, the electrostatic potential) is defined as the amount of work energy needed to move a unit of electric charge from a reference point to the specific point in ...
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 : H_\text(\mathbf r, \mathbf R) \chi(\mathbf r, \mathbf R) = E_\text \chi(\mathbf r, \mathbf R) is solved approximatelyTypically, the Schrödinger equation for molecules cannot be solved exactly. Approximation methods include the Hartree-Fock method The quantity r stands for all electronic coordinates and R for all nuclear coordinates. The electronic energy
eigenvalue In linear algebra, an eigenvector () or characteristic vector of a linear transformation is a nonzero vector that changes at most by a scalar factor when that linear transformation is applied to it. The corresponding eigenvalue, often denoted ...
''E''e 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 ''E''e as a function of R. This is the potential energy surface (PES): ''E''e(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 The adiabatic theorem is a concept in quantum mechanics. Its original form, due to Max Born and Vladimir Fock (1928), was stated as follows: :''A physical system remains in its instantaneous eigenstate if a given perturbation is acting on it s ...
, this manner of obtaining a PES is often referred to as the ''adiabatic approximation'' and the PES itself is called an ''adiabatic surface''.It is assumed, in accordance with the
adiabatic theorem The adiabatic theorem is a concept in quantum mechanics. Its original form, due to Max Born and Vladimir Fock (1928), was stated as follows: :''A physical system remains in its instantaneous eigenstate if a given perturbation is acting on it s ...
, 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.
In the second step of the BO approximation the nuclear kinetic energy ''T''n (containing partial derivatives with respect to the components of R) is reintroduced, and the Schrödinger equation for the nuclear motionThis 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. : _\text + E_\text(\mathbf R)\phi(\mathbf R) = E \phi(\mathbf R) 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 Eckart conditions, named after Carl Eckart, simplify the nuclear motion (rovibrational) Hamiltonian that arises in the second step of the Born–Oppenheimer approximation. They make it possible to approximately separate rotation from vibrat ...
. The eigenvalue ''E'' is the total energy of the molecule, including contributions from electrons, nuclear vibrations, and overall rotation and translation of the molecule. In accord with the
Hellmann–Feynman theorem In quantum mechanics, the Hellmann–Feynman theorem relates the derivative of the total energy with respect to a parameter, to the expectation value of the derivative of the Hamiltonian with respect to that same parameter. According to the theore ...
, the nuclear potential is taken to be an average over electron configurations of the sum of the electron–nuclear and internuclear electric potentials.


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: : E_0(\mathbf) \ll E_1(\mathbf) \ll E_2(\mathbf) \ll \cdots \text\mathbf. We start from the ''exact'' non-relativistic, time-independent molecular Hamiltonian: : H = H_\text + T_\text with : H_\text = -\sum_ - \sum_ + \sum_ + \sum_ \quad\text\quad T_\text = -\sum_. The position vectors \mathbf \equiv \ of the electrons and the position vectors \mathbf \equiv \ of the nuclei are with respect to a Cartesian inertial frame. Distances between particles are written as r_ \equiv , \mathbf_i - \mathbf_A, (distance between electron ''i'' and nucleus ''A'') and similar definitions hold for r_ and R_. 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 The Hartree atomic units are a system of natural units of measurement which is especially convenient for atomic physics and computational chemistry calculations. They are named after the physicist Douglas Hartree. By definition, the following four ...
, 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: : T_\text = \sum_ \sum_ \frac \quad\text\quad P_ = -i \frac. Suppose we have ''K'' electronic eigenfunctions \chi_k (\mathbf; \mathbf) of H_\text, that is, we have solved : H_\text \chi_k(\mathbf; \mathbf) = E_k(\mathbf) \chi_k(\mathbf; \mathbf) \quad\text\quad k = 1, \ldots, K. The electronic wave functions \chi_k will be taken to be real, which is possible when there are no magnetic or spin interactions. The ''parametric dependence'' of the functions \chi_k on the nuclear coordinates is indicated by the symbol after the semicolon. This indicates that, although \chi_k is a real-valued function of \mathbf, its functional form depends on \mathbf. For example, in the molecular-orbital-linear-combination-of-atomic-orbitals (LCAO-MO) approximation, \chi_k 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 \mathbf, the LCAO coefficients obtain different values and we see corresponding changes in the functional form of the MO \chi_k. We will assume that the parametric dependence is continuous and differentiable, so that it is meaningful to consider : P_\chi_k(\mathbf; \mathbf) = -i \frac \quad \text\quad \alpha = x,y,z, which in general will not be zero. The total wave function \Psi(\mathbf, \mathbf) is expanded in terms of \chi_k(\mathbf; \mathbf): : \Psi(\mathbf, \mathbf) = \sum_^K \chi_k(\mathbf; \mathbf) \phi_k(\mathbf), with : \langle \chi_(\mathbf; \mathbf) , \chi_k(\mathbf; \mathbf) \rangle_ = \delta_, and where the subscript (\mathbf) indicates that the integration, implied by the
bra–ket notation In quantum mechanics, bra–ket notation, or Dirac notation, is used ubiquitously to denote quantum states. The notation uses angle brackets, and , and a vertical bar , to construct "bras" and "kets". A ket is of the form , v \rangle. Mathemat ...
, is over electronic coordinates only. By definition, the matrix with general element : \big(\mathbb_\text(\mathbf)\big)_ \equiv \langle \chi_(\mathbf; \mathbf) , H_\text , \chi_k(\mathbf; \mathbf) \rangle_ = \delta_ E_k(\mathbf) is diagonal. After multiplication by the real function \chi_(\mathbf; \mathbf) from the left and integration over the electronic coordinates \mathbf the total Schrödinger equation : H \Psi(\mathbf, \mathbf) = E \Psi(\mathbf, \mathbf) is turned into a set of ''K'' coupled eigenvalue equations depending on nuclear coordinates only : mathbb_\text(\mathbf) + \mathbb_\text(\mathbf)\boldsymbol(\mathbf) = E \boldsymbol(\mathbf). The column vector \boldsymbol(\mathbf) has elements \phi_k(\mathbf),\ k = 1, \ldots, K. The matrix \mathbb_\text(\mathbf) is diagonal, and the nuclear Hamilton matrix is non-diagonal; its off-diagonal (''vibronic coupling'') terms \big(\mathbb_\text(\mathbf)\big)_ 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 One of the guiding principles in modern chemical dynamics and spectroscopy is that the motion of the nuclei in a molecule is slow compared to that of its electrons. This is justified by the large disparity between the mass of an electron and th ...
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 Leibniz's rule (named after Gottfried Wilhelm Leibniz) may refer to one of the following: * Product rule in differential calculus * General Leibniz rule, a generalization of the product rule * Leibniz integral rule * The alternating series test, al ...
for differentiation, the matrix elements of T_\text as : T_\text(\mathbf)_ \equiv \big(\mathbb_\text(\mathbf)\big)_ = \delta_ T_\text - \sum_\frac \langle\chi_, P_, \chi_k\rangle_ P_ + \langle\chi_, T_\text, \chi_k\rangle_. The diagonal (k' = k) matrix elements \langle\chi_, P_, \chi_k\rangle_ of the operator P_ vanish, because we assume time-reversal invariant, so \chi_k can be chosen to be always real. The off-diagonal matrix elements satisfy : \langle\chi_, P_, \chi_k\rangle_ = \frac . The matrix element in the numerator is : \langle\chi_, _, H_\mathrm, \chi_k\rangle_ = iZ_A\sum_i \left\langle\chi_\left, \frac\\chi_k\right\rangle_ \quad\text\quad \mathbf_ \equiv \mathbf_i - \mathbf_A. The matrix element of the one-electron operator appearing on the right side is finite. When the two surfaces come close, E_(\mathbf) \approx E_(\mathbf), 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 P^A_\alpha is effectively zero. The third term on the right side of the expression for the matrix element of ''T''n (the ''Born–Oppenheimer diagonal correction'') can approximately be written as the matrix of P^A_\alpha 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: : _\text + E_k(\mathbf)\phi_k(\mathbf) = E \phi_k(\mathbf) \quad\text\quad k = 1, \ldots, K, 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 One of the guiding principles in modern chemical dynamics and spectroscopy is that the motion of the nuclei in a molecule is slow compared to that of its electrons. This is justified by the large disparity between the mass of an electron and th ...
approximation.


The Born–Oppenheimer approximation with correct symmetry

To include the correct symmetry within the Born–Oppenheimer (BO) approximation, a molecular system presented in terms of (mass-dependent) nuclear coordinates \mathbf and formed by the two lowest BO adiabatic potential energy surfaces (PES) u_1(\mathbf) and u_2 (\mathbf) is considered. To ensure the validity of the BO approximation, the energy ''E'' of the system is assumed to be low enough so that u_2 (\mathbf) becomes a closed PES in the region of interest, with the exception of sporadic infinitesimal sites surrounding degeneracy points formed by u_1(\mathbf) and u_2(\mathbf) (designated as (1, 2) degeneracy points). The starting point is the nuclear adiabatic BO (matrix) equation written in the form : -\frac (\nabla + \tau)^2 \Psi + (\mathbf - E)\Psi = 0, where \Psi(\mathbf) is a column vector containing the unknown nuclear wave functions \psi_k(\mathbf), \mathbf(\mathbf) is a diagonal matrix containing the corresponding adiabatic potential energy surfaces u_k(\mathbf), ''m'' is the reduced mass of the nuclei, ''E'' is the total energy of the system, \nabla is the gradient operator with respect to the nuclear coordinates \mathbf, and \mathbf(\mathbf) is a matrix containing the vectorial non-adiabatic coupling terms (NACT): :\mathbf_ = \langle \zeta_j , \nabla\zeta_k \rangle. Here , \zeta_n\rangle 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: :-\frac \nabla^2\psi_1 + (\tilde_1 - E)\psi_1 - \frac \mathbf_\nabla + \nabla\mathbf_psi_2 = 0, :-\frac \nabla^2\psi_2 + (\tilde_2 - E)\psi_2 + \frac \mathbf_\nabla + \nabla\mathbf_psi_1 = 0, where \tilde_k(\mathbf) = u_k(\mathbf) + (\hbar^/2m)\tau_^2 (''k'' = 1, 2), and \mathbf\tau_ = \mathbf\tau_(\mathbf) is the (vectorial) NACT responsible for the coupling between u_1(\mathbf) and u_2(\mathbf). Next a new function is introduced: : \chi = \psi_1 + i\psi_2, and the corresponding rearrangements are made: # Multiplying the second equation by ''i'' and combining it with the first equation yields the (complex) equation -\frac \nabla^\chi + (\tilde_1 - E)\chi + i\frac \mathbf_\nabla + \nabla\mathbf_chi + i(u_1 - u_2)\psi_2 = 0. # The last term in this equation can be deleted for the following reasons: At those points where u_2(\mathbf) is classically closed, \psi_(\mathbf) \sim 0 by definition, and at those points where u_2(\mathbf) becomes classically allowed (which happens at the vicinity of the (1, 2) degeneracy points) this implies that: u_1(\mathbf) \sim u_2(\mathbf), or u_1(\mathbf) - u_2(\mathbf) \sim 0. Consequently, the last term is, indeed, negligibly small at every point in the region of interest, and the equation simplifies to become -\frac \nabla^\chi + (\tilde_1 - E)\chi + i\frac \mathbf_\nabla + \nabla\mathbf_chi = 0. 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 u_0(\mathbf), which coincides with u_1(\mathbf) at the asymptotic region. The equation with an elastic potential can be solved, in a straightforward manner, by substitution. Thus, if \chi_0 is the solution of this equation, it is presented as :\chi_0(\mathbf, \Gamma) = \xi_(\mathbf) \exp\left \Gamma)\right where \Gamma is an arbitrary contour, and the exponential function contains the relevant symmetry as created while moving along \Gamma. The function \xi_0(\mathbf) can be shown to be a solution of the (unperturbed/elastic) equation :-\frac \nabla^\xi_0 + (u_0 - E) \xi_0 = 0. Having \chi_0(\mathbf, \Gamma), the full solution of the above decoupled equation takes the form :\chi(\mathbf, \Gamma) = \chi_0(\mathbf, \Gamma) + \eta(\mathbf, \Gamma), where \eta(\mathbf, \Gamma) satisfies the resulting inhomogeneous equation: :-\frac \nabla^\eta + (\tilde_1 - E)\eta + i\frac \mathbf_\nabla + \nabla\mathbf_eta = (u_1 - u_0)\chi_0. 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 In quantum chemistry, a conical intersection of two or more potential energy surfaces is the set of molecular geometry points where the potential energy surfaces are degenerate (intersect) and the non-adiabatic couplings between these states are ...
. 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 Ionization, or Ionisation is the process by which an atom or a molecule acquires a negative or positive charge by gaining or losing electrons, often in conjunction with other chemical changes. The resulting electrically charged atom or molecule i ...
*
Adiabatic process (quantum mechanics) The adiabatic theorem is a concept in quantum mechanics. Its original form, due to Max Born and Vladimir Fock (1928), was stated as follows: :''A physical system remains in its instantaneous eigenstate if a given perturbation is acting on it s ...
*
Avoided crossing In quantum physics and quantum chemistry, an avoided crossing (sometimes called intended crossing, ''non-crossing'' or anticrossing) is the phenomenon where two eigenvalues of an Hermitian matrix representing a quantum observable and depending on ' ...
*
Born–Huang approximation The Born–Huang approximation (named after Max Born and Huang Kun) is an approximation closely related to the Born–Oppenheimer approximation. It takes into account diagonal nonadiabatic effects in the electronic Hamiltonian better than t ...
*
Franck–Condon principle The Franck–Condon principle (named for James Franck and Edward Condon) is a rule in spectroscopy and quantum chemistry that explains the intensity of vibronic transitions (the simultaneous changes in electronic and vibrational energy levels of ...
* Kohn anomaly


Notes


References


External links

Resources related to the Born–Oppenheimer approximation:
The original article
(in German)
Translation by S. M. Blinder


a section from Peter Haynes' doctoral thesis {{DEFAULTSORT:Born-Oppenheimer approximation Quantum chemistry Approximations Max Born