fbpx
Wikipedia

Born–Oppenheimer approximation

In quantum chemistry 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 functions of atomic nuclei 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.[1] The approach is named after Max Born and his 23-year-old graduate student J. Robert Oppenheimer, the latter of whom proposed it in 1927 during a period of intense ferment in the development of quantum mechanics.[2][3]

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 spectroscopy, using the BO approximation means considering molecular energy as a sum of independent terms, e.g.:

These terms are of different orders of magnitude and the nuclear spin energy is so small that it is often omitted. The electronic energies 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 edit

The benzene 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 = 36 nuclear + 3 × 42 = 126 electronic = 162 variables for the wave function. The computational complexity, i.e., the computational power required to solve an eigenvalue equation, increases faster than the square of the number of coordinates.[4]

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 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   hypothetical calculation steps, a series of smaller calculations requiring   (with N being the number of grid points for the potential) and a very small calculation requiring   steps can be performed. In practice, the scaling of the problem is larger than  , and more approximations are applied in computational chemistry 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 edit

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 ( ) of a molecule as the product of an electronic wavefunction and a nuclear (vibrational, rotational) wavefunction.  . 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,[note 1] that is, the corresponding operator Tn is subtracted from the total molecular Hamiltonian. In the remaining electronic Hamiltonian He 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 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

 

where   is the electronic wavefunction for given positions of nuclei (fixed R), is solved approximately.[note 2] 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):  . 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.[note 3]

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[note 4]

 

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.[clarification needed] 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.

Derivation edit

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 correct symmetry edit

To include the correct symmetry within the Born–Oppenheimer (BO) approximation,[2][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][10] 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 edit

Notes edit

  1. ^ 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 mnme 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. 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 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.[citation needed]
  2. ^ Typically, the Schrödinger equation for molecules cannot be solved exactly. Approximation methods include the Hartree-Fock method
  3. ^ 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.[citation needed]
  4. ^ 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.[citation needed]

References edit

  1. ^ Hanson, David. "The Born-Oppenheimer Approximation". Chemistry Libretexts. Chemical Education Digital Library. Retrieved 2 August 2022.
  2. ^ a b 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.
  3. ^ Bird, Kai; Sherwin, Martin K. (2006). American Prometheus: The Triumph and Tragedy of J. Robert Oppenheimer (1st ed.). Vintage Books. pp. 65–66. ISBN 978-0375726262.
  4. ^ T. H. Cormen, C. E. Leiserson, R. L. Rivest, C. Stein, Introduction to Algorithms, 3rd ed., MIT Press, Cambridge, MA, 2009, § 28.2.
  5. ^ Born, M.; Huang, K. (1954). "IV". Dynamical Theory of Crystal Lattices. New York: Oxford University Press.
  6. ^ "Born-Oppenheimer Approach: Diabatization and Topological Matrix". Beyond Born-Oppenheimer: Electronic Nonadiabatic Coupling Terms and Conical Intersections. Hoboken, NJ, USA: John Wiley & Sons, Inc. 28 March 2006. pp. 26–57. doi:10.1002/0471780081.ch2. ISBN 978-0-471-78008-3.
  7. ^ Baer, Michael; Englman, Robert (1997). "A modified Born-Oppenheimer equation: application to conical intersections and other types of singularities". Chemical Physics Letters. Elsevier BV. 265 (1–2): 105–108. Bibcode:1997CPL...265..105B. doi:10.1016/s0009-2614(96)01411-x. ISSN 0009-2614.
  8. ^ Baer, Roi; Charutz, David M.; Kosloff, Ronnie; Baer, Michael (22 November 1996). "A study of conical intersection effects on scattering processes: The validity of adiabatic single‐surface approximations within a quasi‐Jahn–Teller model". The Journal of Chemical Physics. AIP Publishing. 105 (20): 9141–9152. Bibcode:1996JChPh.105.9141B. doi:10.1063/1.472748. ISSN 0021-9606.
  9. ^ Adhikari, Satrajit; Billing, Gert D. (1999). "The conical intersection effects and adiabatic single-surface approximations on scattering processes: A time-dependent wave packet approach". The Journal of Chemical Physics. AIP Publishing. 111 (1): 40–47. Bibcode:1999JChPh.111...40A. doi:10.1063/1.479360. ISSN 0021-9606.
  10. ^ Charutz, David M.; Baer, Roi; Baer, Michael (1997). "A study of degenerate vibronic coupling effects on scattering processes: are resonances affected by degenerate vibronic coupling?". Chemical Physics Letters. Elsevier BV. 265 (6): 629–637. Bibcode:1997CPL...265..629C. doi:10.1016/s0009-2614(96)01494-7. ISSN 0009-2614.

External links edit

Resources related to the Born–Oppenheimer approximation:

  • The original article (in German)
  • Translation by S. M. Blinder
  • Another version of the same translation by S. M. Blinder
  • The Born–Oppenheimer approximation, a section from Peter Haynes' doctoral thesis

born, oppenheimer, approximation, confused, with, born, approximation, quantum, chemistry, molecular, physics, born, oppenheimer, approximation, best, known, mathematical, approximation, molecular, dynamics, specifically, assumption, that, wave, functions, ato. Not to be confused with the Born approximation In quantum chemistry 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 functions of atomic nuclei 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 1 The approach is named after Max Born and his 23 year old graduate student J Robert Oppenheimer the latter of whom proposed it in 1927 during a period of intense ferment in the development of quantum mechanics 2 3 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 spectroscopy using the BO approximation means considering molecular energy as a sum of independent terms e g E total E electronic E vibrational E rotational E nuclear spin displaystyle E text total E text electronic E text vibrational E text rotational E text nuclear spin 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 electronic displaystyle E text electronic 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 Contents 1 Example 2 Detailed description 3 Derivation 4 The Born Oppenheimer approximation with correct symmetry 5 See also 6 Notes 7 References 8 External linksExample editThe benzene molecule consists of 12 nuclei and 42 electrons The Schrodinger 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 36 nuclear 3 42 126 electronic 162 variables for the wave function The computational complexity i e the computational power required to solve an eigenvalue equation increases faster than the square of the number of coordinates 4 When applying the BO approximation two smaller consecutive steps can be used For a given position of the nuclei the electronic Schrodinger equation is solved while treating the nuclei as stationary not coupled with the dynamics of the electrons This corresponding eigenvalue 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 Schrodinger 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 displaystyle 162 2 26 244 nbsp hypothetical calculation steps a series of smaller calculations requiring 126 2 N 15 876 N displaystyle 126 2 N 15 876 N nbsp with N being the number of grid points for the potential and a very small calculation requiring 36 2 1296 displaystyle 36 2 1296 nbsp steps can be performed In practice the scaling of the problem is larger than n 2 displaystyle n 2 nbsp and more approximations are applied in computational chemistry 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 Schrodinger equation Detailed description editThe 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 PS t o t a l displaystyle Psi mathrm total nbsp of a molecule as the product of an electronic wavefunction and a nuclear vibrational rotational wavefunction PS t o t a l ps e l e c t r o n i c ps n u c l e a r displaystyle Psi mathrm total psi mathrm electronic psi mathrm nuclear nbsp 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 note 1 that is the corresponding operator Tn is subtracted from the total molecular Hamiltonian In the remaining electronic Hamiltonian He 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 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 Schrodinger equation H e r R x r R E e x r R displaystyle H text e mathbf r mathbf R chi mathbf r mathbf R E text e chi mathbf r mathbf R nbsp where x r R displaystyle chi mathbf r mathbf R nbsp is the electronic wavefunction for given positions of nuclei fixed R is solved approximately note 2 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 Schrodinger equation one obtains Ee as a function of R This is the potential energy surface PES E e R displaystyle E e mathbf R nbsp 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 note 3 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 Schrodinger equation for the nuclear motion note 4 T n E e R ϕ R E ϕ R displaystyle T text n E text e mathbf R phi mathbf R E phi mathbf R nbsp 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 clarification needed 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 Derivation editIt 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 Schrodinger equation are well separated E 0 R E 1 R E 2 R for all R displaystyle E 0 mathbf R ll E 1 mathbf R ll E 2 mathbf R ll cdots text for all mathbf R nbsp We start from the exact non relativistic time independent molecular Hamiltonian H H e T n displaystyle H H text e T text n nbsp with H e i 1 2 i 2 i A Z A r i A i gt j 1 r i j B gt A Z A Z B R A B and T n A 1 2 M A A 2 displaystyle H text e sum i frac 1 2 nabla i 2 sum i A frac Z A r iA sum i gt j frac 1 r ij sum B gt A frac Z A Z B R AB quad text and quad T text n sum A frac 1 2M A nabla A 2 nbsp The position vectors r r i displaystyle mathbf r equiv mathbf r i nbsp of the electrons and the position vectors R R A R A x y R A y z R A z x displaystyle mathbf R equiv mathbf R A R Axy R Ayz R Azx nbsp of the nuclei are with respect to a Cartesian inertial frame Distances between particles are written as r i A r i R A displaystyle r iA equiv mathbf r i mathbf R A nbsp distance between electron i and nucleus A and similar definitions hold for r i j displaystyle r ij nbsp and R A B displaystyle R AB nbsp 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 T n A a x y z P A a P A a 2 M A with P A a i R A a displaystyle T text n sum A sum alpha x y z frac P A alpha P A alpha 2M A quad text with quad P A alpha i frac partial partial R A alpha nbsp Suppose we have K electronic eigenfunctions x k r R displaystyle chi k mathbf r mathbf R nbsp of H e displaystyle H text e nbsp that is we have solved H e x k r R E k R x k r R for k 1 K displaystyle H text e chi k mathbf r mathbf R E k mathbf R chi k mathbf r mathbf R quad text for quad k 1 ldots K nbsp The electronic wave functions x k displaystyle chi k nbsp will be taken to be real which is possible when there are no magnetic or spin interactions The parametric dependence of the functions x k displaystyle chi k nbsp on the nuclear coordinates is indicated by the symbol after the semicolon This indicates that although x k displaystyle chi k nbsp is a real valued function of r displaystyle mathbf r nbsp its functional form depends on R displaystyle mathbf R nbsp For example in the molecular orbital linear combination of atomic orbitals LCAO MO approximation x k displaystyle chi k nbsp 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 R displaystyle mathbf R nbsp the LCAO coefficients obtain different values and we see corresponding changes in the functional form of the MO x k displaystyle chi k nbsp We will assume that the parametric dependence is continuous and differentiable so that it is meaningful to consider P A a x k r R i x k r R R A a for a x y z displaystyle P A alpha chi k mathbf r mathbf R i frac partial chi k mathbf r mathbf R partial R A alpha quad text for quad alpha x y z nbsp which in general will not be zero The total wave function PS R r displaystyle Psi mathbf R mathbf r nbsp is expanded in terms of x k r R displaystyle chi k mathbf r mathbf R nbsp PS R r k 1 K x k r R ϕ k R displaystyle Psi mathbf R mathbf r sum k 1 K chi k mathbf r mathbf R phi k mathbf R nbsp with x k r R x k r R r d k k displaystyle langle chi k mathbf r mathbf R chi k mathbf r mathbf R rangle mathbf r delta k k nbsp and where the subscript r displaystyle mathbf r nbsp indicates that the integration implied by the bra ket notation is over electronic coordinates only By definition the matrix with general element H e R k k x k r R H e x k r R r d k k E k R displaystyle big mathbb H text e mathbf R big k k equiv langle chi k mathbf r mathbf R H text e chi k mathbf r mathbf R rangle mathbf r delta k k E k mathbf R nbsp is diagonal After multiplication by the real function x k r R displaystyle chi k mathbf r mathbf R nbsp from the left and integration over the electronic coordinates r displaystyle mathbf r nbsp the total Schrodinger equation H PS R r E PS R r displaystyle H Psi mathbf R mathbf r E Psi mathbf R mathbf r nbsp is turned into a set of K coupled eigenvalue equations depending on nuclear coordinates only H n R H e R ϕ R E ϕ R displaystyle mathbb H text n mathbf R mathbb H text e mathbf R boldsymbol phi mathbf R E boldsymbol phi mathbf R nbsp The column vector ϕ R displaystyle boldsymbol phi mathbf R nbsp has elements ϕ k R k 1 K displaystyle phi k mathbf R k 1 ldots K nbsp The matrix H e R displaystyle mathbb H text e mathbf R nbsp is diagonal and the nuclear Hamilton matrix is non diagonal its off diagonal vibronic coupling terms H n R k k displaystyle big mathbb H text n mathbf R big k k nbsp 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 T n displaystyle T text n nbsp as T n R k k H n R k k d k k T n A a 1 M A x k P A a x k r P A a x k T n x k r displaystyle T text n mathbf R k k equiv big mathbb H text n mathbf R big k k delta k k T text n sum A alpha frac 1 M A langle chi k P A alpha chi k rangle mathbf r P A alpha langle chi k T text n chi k rangle mathbf r nbsp The diagonal k k displaystyle k k nbsp matrix elements x k P A a x k r displaystyle langle chi k P A alpha chi k rangle mathbf r nbsp of the operator P A a displaystyle P A alpha nbsp vanish because we assume time reversal invariant so x k displaystyle chi k nbsp can be chosen to be always real The off diagonal matrix elements satisfy x k P A a x k r x k P A a H e x k r E k R E k R displaystyle langle chi k P A alpha chi k rangle mathbf r frac langle chi k P A alpha H text e chi k rangle mathbf r E k mathbf R E k mathbf R nbsp The matrix element in the numerator is x k P A a H e x k r i Z A i x k r i A a r i A 3 x k r with r i A r i R A displaystyle langle chi k P A alpha H mathrm e chi k rangle mathbf r iZ A sum i left langle chi k left frac mathbf r iA alpha r iA 3 right chi k right rangle mathbf r quad text with quad mathbf r iA equiv mathbf r i mathbf R A nbsp The matrix element of the one electron operator appearing on the right side is finite When the two surfaces come close E k R E k R displaystyle E k mathbf R approx E k mathbf R nbsp 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 A displaystyle P alpha A nbsp 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 P a A displaystyle P alpha A nbsp 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 T n E k R ϕ k R E ϕ k R for k 1 K displaystyle T text n E k mathbf R phi k mathbf R E phi k mathbf R quad text for quad k 1 ldots K nbsp 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 correct symmetry editTo include the correct symmetry within the Born Oppenheimer BO approximation 2 5 a molecular system presented in terms of mass dependent nuclear coordinates q displaystyle mathbf q nbsp and formed by the two lowest BO adiabatic potential energy surfaces PES u 1 q displaystyle u 1 mathbf q nbsp and u 2 q displaystyle u 2 mathbf q nbsp 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 q displaystyle u 2 mathbf q nbsp becomes a closed PES in the region of interest with the exception of sporadic infinitesimal sites surrounding degeneracy points formed by u 1 q displaystyle u 1 mathbf q nbsp and u 2 q displaystyle u 2 mathbf q nbsp designated as 1 2 degeneracy points The starting point is the nuclear adiabatic BO matrix equation written in the form 6 ℏ 2 2 m t 2 PS u E PS 0 displaystyle frac hbar 2 2m nabla tau 2 Psi mathbf u E Psi 0 nbsp where PS q displaystyle Psi mathbf q nbsp is a column vector containing the unknown nuclear wave functions ps k q displaystyle psi k mathbf q nbsp u q displaystyle mathbf u mathbf q nbsp is a diagonal matrix containing the corresponding adiabatic potential energy surfaces u k q displaystyle u k mathbf q nbsp m is the reduced mass of the nuclei E is the total energy of the system displaystyle nabla nbsp is the gradient operator with respect to the nuclear coordinates q displaystyle mathbf q nbsp and t q displaystyle mathbf tau mathbf q nbsp is a matrix containing the vectorial non adiabatic coupling terms NACT t j k z j z k displaystyle mathbf tau jk langle zeta j nabla zeta k rangle nbsp Here z n displaystyle zeta n rangle nbsp 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 ℏ 2 2 m 2 ps 1 u 1 E ps 1 ℏ 2 2 m 2 t 12 t 12 ps 2 0 displaystyle frac hbar 2 2m nabla 2 psi 1 tilde u 1 E psi 1 frac hbar 2 2m 2 mathbf tau 12 nabla nabla mathbf tau 12 psi 2 0 nbsp ℏ 2 2 m 2 ps 2 u 2 E ps 2 ℏ 2 2 m 2 t 12 t 12 ps 1 0 displaystyle frac hbar 2 2m nabla 2 psi 2 tilde u 2 E psi 2 frac hbar 2 2m 2 mathbf tau 12 nabla nabla mathbf tau 12 psi 1 0 nbsp where u k q u k q ℏ 2 2 m t 12 2 displaystyle tilde u k mathbf q u k mathbf q hbar 2 2m tau 12 2 nbsp k 1 2 and t 12 t 12 q displaystyle mathbf tau 12 mathbf tau 12 mathbf q nbsp is the vectorial NACT responsible for the coupling between u 1 q displaystyle u 1 mathbf q nbsp and u 2 q displaystyle u 2 mathbf q nbsp Next a new function is introduced 7 x ps 1 i ps 2 displaystyle chi psi 1 i psi 2 nbsp and the corresponding rearrangements are made Multiplying the second equation by i and combining it with the first equation yields the complex equation ℏ 2 2 m 2 x u 1 E x i ℏ 2 2 m 2 t 12 t 12 x i u 1 u 2 ps 2 0 displaystyle frac hbar 2 2m nabla 2 chi tilde u 1 E chi i frac hbar 2 2m 2 mathbf tau 12 nabla nabla mathbf tau 12 chi i u 1 u 2 psi 2 0 nbsp The last term in this equation can be deleted for the following reasons At those points where u 2 q displaystyle u 2 mathbf q nbsp is classically closed ps 2 q 0 displaystyle psi 2 mathbf q sim 0 nbsp by definition and at those points where u 2 q displaystyle u 2 mathbf q nbsp becomes classically allowed which happens at the vicinity of the 1 2 degeneracy points this implies that u 1 q u 2 q displaystyle u 1 mathbf q sim u 2 mathbf q nbsp or u 1 q u 2 q 0 displaystyle u 1 mathbf q u 2 mathbf q sim 0 nbsp Consequently the last term is indeed negligibly small at every point in the region of interest and the equation simplifies to become ℏ 2 2 m 2 x u 1 E x i ℏ 2 2 m 2 t 12 t 12 x 0 displaystyle frac hbar 2 2m nabla 2 chi tilde u 1 E chi i frac hbar 2 2m 2 mathbf tau 12 nabla nabla mathbf tau 12 chi 0 nbsp 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 q displaystyle u 0 mathbf q nbsp which coincides with u 1 q displaystyle u 1 mathbf q nbsp at the asymptotic region The equation with an elastic potential can be solved in a straightforward manner by substitution Thus if x 0 displaystyle chi 0 nbsp is the solution of this equation it is presented as x 0 q G 3 0 q exp i G d q t q G displaystyle chi 0 mathbf q Gamma xi 0 mathbf q exp left i int Gamma d mathbf q cdot mathbf tau mathbf q Gamma right nbsp where G displaystyle Gamma nbsp is an arbitrary contour and the exponential function contains the relevant symmetry as created while moving along G displaystyle Gamma nbsp The function 3 0 q displaystyle xi 0 mathbf q nbsp can be shown to be a solution of the unperturbed elastic equation ℏ 2 2 m 2 3 0 u 0 E 3 0 0 displaystyle frac hbar 2 2m nabla 2 xi 0 u 0 E xi 0 0 nbsp Having x 0 q G displaystyle chi 0 mathbf q Gamma nbsp the full solution of the above decoupled equation takes the form x q G x 0 q G h q G displaystyle chi mathbf q Gamma chi 0 mathbf q Gamma eta mathbf q Gamma nbsp where h q G displaystyle eta mathbf q Gamma nbsp satisfies the resulting inhomogeneous equation ℏ 2 2 m 2 h u 1 E h i ℏ 2 2 m 2 t 12 t 12 h u 1 u 0 x 0 displaystyle frac hbar 2 2m nabla 2 eta tilde u 1 E eta i frac hbar 2 2m 2 mathbf tau 12 nabla nabla mathbf tau 12 eta u 1 u 0 chi 0 nbsp 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 10 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 editAdiabatic ionization Adiabatic process quantum mechanics Avoided crossing Born Huang approximation Franck Condon principle Kohn anomalyNotes edit 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 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 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 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 citation needed Typically the Schrodinger equation for molecules cannot be solved exactly Approximation methods include the Hartree Fock method 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 citation needed 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 citation needed References edit Hanson David The Born Oppenheimer Approximation Chemistry Libretexts Chemical Education Digital Library Retrieved 2 August 2022 a b 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 Bird Kai Sherwin Martin K 2006 American Prometheus The Triumph and Tragedy of J Robert Oppenheimer 1st ed Vintage Books pp 65 66 ISBN 978 0375726262 T H Cormen C E Leiserson R L Rivest C Stein Introduction to Algorithms 3rd ed MIT Press Cambridge MA 2009 28 2 Born M Huang K 1954 IV Dynamical Theory of Crystal Lattices New York Oxford University Press Born Oppenheimer Approach Diabatization and Topological Matrix Beyond Born Oppenheimer Electronic Nonadiabatic Coupling Terms and Conical Intersections Hoboken NJ USA John Wiley amp Sons Inc 28 March 2006 pp 26 57 doi 10 1002 0471780081 ch2 ISBN 978 0 471 78008 3 Baer Michael Englman Robert 1997 A modified Born Oppenheimer equation application to conical intersections and other types of singularities Chemical Physics Letters Elsevier BV 265 1 2 105 108 Bibcode 1997CPL 265 105B doi 10 1016 s0009 2614 96 01411 x ISSN 0009 2614 Baer Roi Charutz David M Kosloff Ronnie Baer Michael 22 November 1996 A study of conical intersection effects on scattering processes The validity of adiabatic single surface approximations within a quasi Jahn Teller model The Journal of Chemical Physics AIP Publishing 105 20 9141 9152 Bibcode 1996JChPh 105 9141B doi 10 1063 1 472748 ISSN 0021 9606 Adhikari Satrajit Billing Gert D 1999 The conical intersection effects and adiabatic single surface approximations on scattering processes A time dependent wave packet approach The Journal of Chemical Physics AIP Publishing 111 1 40 47 Bibcode 1999JChPh 111 40A doi 10 1063 1 479360 ISSN 0021 9606 Charutz David M Baer Roi Baer Michael 1997 A study of degenerate vibronic coupling effects on scattering processes are resonances affected by degenerate vibronic coupling Chemical Physics Letters Elsevier BV 265 6 629 637 Bibcode 1997CPL 265 629C doi 10 1016 s0009 2614 96 01494 7 ISSN 0009 2614 External links editResources related to the Born Oppenheimer approximation The original article in German Translation by S M Blinder Another version of the same 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 amp oldid 1175347603, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.