GF method: Difference between revisions

From Citizendium
Jump to navigation Jump to search
imported>Paul Wormer
(→‎The GF method: Clarified the method after some valid WP criticism)
imported>Mark Widmer
mNo edit summary
 
(3 intermediate revisions by one other user not shown)
Line 1: Line 1:
{{subpages}}
{{subpages}}


In [[chemistry]] and [[molecular physics]], Wilson's '''GF method''', sometimes referred to as the '''FG method''', is a classical mechanical  method to obtain certain ''internal coordinates'' for a vibrating [[semi-rigid molecule]], the so-called ''normal coordinates'' ''Q''<sub>k</sub> (also known as normal modes). Simultaneously, the vibrational energies  associated with these normal modes are obtained.
In [[chemistry]] and [[molecular physics]], Wilson's '''GF method''', sometimes referred to as the '''FG method''', is a method to obtain certain internal coordinates for a vibrating [[semi-rigid molecule]], the so-called ''normal coordinates'' (also known as normal modes). In Wilson's GF method it is assumed that the total molecular energy consists only of vibrational energies, i.e., the overall rotation and translation of the molecule is ignored, as are the motions of its electrons. Neglecting electrons, an ''N''-atomic molecule has 3''N'' degrees of freedom, and, as the description of the overall rotational and translational motions of a molecule requires 6 coordinates, the vibrational problem has 3''N''&minus;6 degrees of freedom.
Normal coordinates decouple the classical vibrational motions of the molecule making it possible to compute the vibrational amplitudes of atoms in a molecule.  In Wilson's GF method it is assumed that the molecular kinetic energy consists only of vibrations of the atoms, ''i.e.,''  the overall rotational and translational energies of the molecule are ignored. Normal coordinates are usually defined  in a classical mechanical context, but they also appear in the quantum mechanical descriptions of the vibrational motions of molecules and the Coriolis coupling between rotations and vibrations.


It follows from application of the [[Eckart conditions]] that the matrix '''G'''<sup>&minus;1</sup> gives the kinetic energy in terms of arbitrary linear internal coordinates, while '''F''' represents the ''harmonic'' potential energy in terms of these internal coordinates. Solution of the GF equations gives the matrix that transforms from general internal coordinates to the special set of normal coordinates. The associated eigen energies of the vibrations are obtained along the way.
The vibrations of the atoms in a molecule are all coupled; roughly speaking, an ''N''-atomic molecule can be seen as a system of 3''N''&minus;6 coupled one-dimensional [[harmonic oscillators]]. Normal coordinates decouple these vibrational motions, thus making it possible to compute the vibrational displacements of the atoms. Further, it is assumed that the vibrations are harmonic, which in general is only approximately the case.


The method is called after the two matrices '''F''' and '''G''' that enter the final equations. The name of the most important of its early investigators<ref name="Wilson41">E. B. Wilson, Jr. ''Some Mathematical Methods for the Study of Molecular Vibrations'', J. Chem. Phys. vol. '''9''', pp. 76-84 (1941)</ref> [[Edgar Bright Wilson|E. Bright Wilson]], Jr. (1908&ndash;1992) is often attached to the method.
Normal coordinates are often defined in a classical mechanical context, but they appear in the quantum mechanical descriptions of the vibrational motions of molecules as well, and also in the [[Coriolis coupling]] between rotations and vibrations.
 
It follows from application of the [[Eckart conditions]] that the inverse of a matrix '''G''' gives the vibrational kinetic energy in terms of arbitrary linearized internal coordinates, while '''F''' represents the ''harmonic'' potential energy in terms of the same internal coordinates. Solving the GF equations gives the matrix that transforms from the linearized internal coordinates to the normal coordinates. The associated energies of the vibrations are obtained after fixing initial amplitudes (classically) or vibrational quantum numbers (quantum mechanically).
 
Evidently, the method is named after the two matrices '''F''' and '''G''' that enter the final equations. The name of the most important of its early investigators<ref name="Wilson41">E. B. Wilson, Jr. ''Some Mathematical Methods for the Study of Molecular Vibrations'', J. Chem. Phys. vol. '''9''', pp. 76-84 (1941)</ref> [[Edgar Bright Wilson|E. Bright Wilson]], Jr. (1908&ndash;1992) is often attached to the method.


== The GF method ==
== The GF method ==
A non-linear molecule consisting of ''N'' atoms has 3''N''&minus;6 internal degrees of freedom, because the positioning of a molecule in three-dimensional space requires three degrees of freedom and the description of its orientation requires another three degree of freedom. These six degrees of freedom must be subtracted from the 3''N'' degrees of freedom of a system of ''N'' particles.  
A non-linear molecule consisting of ''N'' atoms has 3''N''&minus;6 internal degrees of freedom, because the positioning of a molecule in three-dimensional space requires three degrees of freedom and the description of its orientation requires another three degrees of freedom. These six degrees of freedom must be subtracted from the 3''N'' degrees of freedom of a system of ''N'' particles.


The atoms in a molecule are bound by a [[potential energy surface]] (PES) (also known as [[force field (chemistry)|force field]]) which is a function of 3''N''&minus;6 coordinates.
The atoms in a molecule are bound by a [[potential energy surface]] (PES) (also known as [[force field (chemistry)|force field]]) which is a function of 3''N''&minus;6 coordinates.
The internal degrees of freedom ''q''<sub>1</sub>, ..., ''q''<sub>3''N''&minus;6</sub> describing the PES in an optimum way are often non-linear; they are for instance ''valence coordinates'', such as bending and torsion angles and bond stretches. It is possible to write the quantum mechanical kinetic energy operator for such [[curvilinear coordinates]], but it is hard to formulate a general theory applicable to any molecule. This is why Wilson<ref name="Wilson41"/> linearized the internal coordinates by assuming small displacements. The linearized version of the internal coordinate ''q''<sub>''t''</sub> is denoted by ''S''<sub>''t''</sub>.  
The internal degrees of freedom ''q''<sub>1</sub>, ..., ''q''<sub>3''N''&minus;6</sub> describing the PES in an optimum way are often non-linear; they are for instance ''valence coordinates'', such as bending and torsion angles and bond stretches. It is possible to write the quantum mechanical kinetic energy operator for such [[curvilinear coordinates]], but it is hard to formulate a general theory applicable to any molecule. This is why Wilson<ref name="Wilson41"/> linearized the internal coordinates by assuming small displacements. The linearized version of the internal coordinate ''q''<sub>''t''</sub> is denoted by ''S''<sub>''t''</sub>.  


The PES ''V'' can be [[Taylor series|Taylor-expanded]] around its minimum in terms of the ''S''<sub>''t''</sub>. The third term (the [[Hessian matrix|Hessian]] of ''V'') evaluated in the minimum is a force constant matrix '''F'''. In the ''harmonic approximation''&mdash;on which the method is based&mdash;the Taylor series is ended after this quadratic term. The second term, containing first derivatives, is zero because it is evaluated in the minimum of ''V''. The first term can be included in the zero of energy.
The PES ''V'' can be [[Taylor series|Taylor-expanded]] around its minimum in terms of the ''S''<sub>''t''</sub>. The third term (the [[Hessian matrix|Hessian]] of ''V'') evaluated in the minimum is a force constant matrix '''F'''. In the ''harmonic approximation''&mdash;on which the method is based&mdash;the Taylor series is ended after this quadratic term. The second term, containing first derivatives, is zero because it is evaluated in the minimum of ''V''. The first term can be included in the zero of energy.
Thus,
Thus,
:<math> 2V \approx \sum_{s,t=1}^{3N-6} F_{st} S_s\, S_t </math>.
:<math> 2V \approx \sum_{s,t=1}^{3N-6} F_{st} S_s\, S_t </math>.
Line 21: Line 23:
The classical vibrational kinetic energy has the form:
The classical vibrational kinetic energy has the form:
:<math> 2T = \sum_{s,t=1}^{3N-6} g_{st}(\mathbf{q})  \dot{S}_s\dot{S}_t ,</math>
:<math> 2T = \sum_{s,t=1}^{3N-6} g_{st}(\mathbf{q})  \dot{S}_s\dot{S}_t ,</math>
where ''g''<sub>''st''</sub> is an element of the [[metric tensor]] of the internal (curvilinear) coordinates. The dots indicate [[time derivative]]s. Evaluation of the metric tensor '''g'''('''q''') in the minimum '''q'''<sub>0</sub> of ''V'' gives  
where ''g''<sub>''st''</sub> is an element of the [[metric tensor]] of the internal (curvilinear) coordinates. The dots indicate [[time derivative]]s. Evaluation of the metric tensor '''g'''('''q''') in the minimum '''q'''<sub>0</sub> of ''V'' gives the [[positive define matrix|positive definite]] and [[symmetric matrix]] '''G''' &equiv; '''g'''('''q'''<sub>0</sub>)<sup>&minus;1</sup>.  
the [[positive define matrix|positive definite]] and [[symmetric matrix]] '''G''' &equiv; '''g'''('''q'''<sub>0</sub>)<sup>&minus;1</sup>.  
 
It is assumed that both '''F''' and '''G''' can be determined. The (inverse of the) kinetic energy matrix '''G''' is usually obtained by analytic methods and the force constant matrix '''F''' by fitting to experimental infrared data and/or quantum chemical calculations. Knowing the matrices, one can solve the following two problems simultaneously
It is assumed that both '''F''' and '''G''' can be determined. The (inverse of the) kinetic energy matrix '''G''' is usually obtained by analytic methods and the force constant matrix '''F''' by fitting to experimental infrared data and/or quantum chemical calculations. Knowing the matrices, one can solve the following two problems simultaneously
:<math> \mathbf{L}^\mathrm{T} \mathbf{F} \mathbf{L} =\boldsymbol{\Phi}
:<math> \mathbf{L}^\mathrm{T} \mathbf{F} \mathbf{L} =\boldsymbol{\Phi}
\quad \mathrm{and}\quad \mathbf{L}^\mathrm{T} \mathbf{G}^{-1} \mathbf{L} = \mathbf{E},
\quad \mathrm{and}\quad \mathbf{L}^\mathrm{T} \mathbf{G}^{-1} \mathbf{L} = \mathbf{E},
Line 34: Line 35:
</math>
</math>
where <math>\boldsymbol{\Phi}=\operatorname{diag}(f_1,\ldots, f_{3N-6}) </math>.  
where <math>\boldsymbol{\Phi}=\operatorname{diag}(f_1,\ldots, f_{3N-6}) </math>.  
In [[numerical mathematics]] several algorithms are known to solve generalized eigenvalue problems. The solution yields the matrix '''L'''   and the diagonal matrix '''&Phi;''' that has the vibrational [[force constant]]s on the diagonal.  
In [[numerical mathematics]] several algorithms are known to solve generalized eigenvalue problems. The solution yields the matrix '''L''' and the diagonal matrix '''&Phi;''' that has the vibrational [[force constant]]s on the diagonal.  
The matrix '''L'''<sup>&minus;1</sup> contains the ''normal coordinates'' ''Q''<sub>k</sub> in its rows:
The matrix '''L'''<sup>&minus;1</sup> contains the ''normal coordinates'' ''Q''<sub>k</sub> in its rows:
:<math> Q_k = \sum_{t=1}^{3N-6} (\mathbf{L}^{-1})_{kt} S_t , \quad k=1,\ldots, 3N-6. \,</math>
:<math> Q_k = \sum_{t=1}^{3N-6} (\mathbf{L}^{-1})_{kt} S_t , \quad k=1,\ldots, 3N-6. \,</math>
Because of the form of the generalized eigenvalue problem, the method is called the GF method,
Because of the form of the generalized eigenvalue problem, the method is called the GF method,
often with the name of its originator attached to it: '''Wilson's GF method'''. It is of interest to point out that by matrix transposition of both sides of the equation and by use of the fact
often with the name of its originator attached to it: ''Wilson's GF method''. It is of interest to point out that by matrix transposition of both sides of the equation and by use of the fact
that both '''G''' and '''F''' are symmetric, (as is a diagonal matrix), one can recast this equation into a very similar one for '''FG''' . This is why the method is also referred to as '''Wilson's FG method'''.  
that both '''G''' and '''F''' are symmetric, (as is a diagonal matrix), one can recast this equation into a very similar one for '''FG'''. This is why the method is also referred to as ''Wilson's FG method''.


We introduce the vectors
We introduce the vectors
:<math>\mathbf{s} = \operatorname{col}(S_1,\ldots, S_{3N-6})
:<math>\mathbf{s} = \operatorname{col}(S_1,\ldots, S_{3N-6})
\quad\mathrm{and}\quad
\quad\mathrm{and}\quad
Line 50: Line 51:
\mathbf{s} = \mathbf{L} \mathbf{Q}.
\mathbf{s} = \mathbf{L} \mathbf{Q}.
</math>
</math>
Upon use of the results of the generalized eigenvalue equation the classical (i.e., non-quantum mechanical) energy ''E'' = ''T '' + ''V'' of the molecule becomes in the harmonic approximation,  
Upon use of the results of the generalized eigenvalue equation the classical (i.e., non-quantum mechanical) energy ''E'' = ''T '' + ''V'' of the molecule becomes in the harmonic approximation,  
:<math>   
:<math>   
2E = \dot{\mathbf{s}}^\mathrm{T} \mathbf{G}^{-1}\dot{\mathbf{s}}+
2E = \dot{\mathbf{s}}^\mathrm{T} \mathbf{G}^{-1}\dot{\mathbf{s}}+
Line 57: Line 58:


::<math>
::<math>
= \dot{\mathbf{Q}}^\mathrm{T} \; \left( \mathbf{L}^\mathrm{T} \mathbf{G}^{-1} \mathbf{L}\right) \;  \dot{\mathbf{Q}}+
= \dot{\mathbf{Q}}^\mathrm{T} \; \left( \mathbf{L}^\mathrm{T} \mathbf{G}^{-1} \mathbf{L}\right) \;  \dot{\mathbf{Q}}+
\mathbf{Q}^\mathrm{T} \left( \mathbf{L}^\mathrm{T}\mathbf{F}\mathbf{L}\right)\;  \mathbf{Q}  
\mathbf{Q}^\mathrm{T} \left( \mathbf{L}^\mathrm{T}\mathbf{F}\mathbf{L}\right)\;  \mathbf{Q}  
</math>
</math>


Line 66: Line 67:


::<math>
::<math>
  = \sum_{t=1}^{3N-6} \big( \dot{Q}_t^2 + f_t Q_t^2 \big).
  = \sum_{t=1}^{3N-6} \big( \dot{Q}_t^2 + f_t Q_t^2 \big).
</math>
</math>
Solution of the GF equations yields ''Q''<sub>''t''</sub> and ''f''<sub>''t''</sub>. Classically, the time dependence of the [[harmonic oscillator (classical)|harmonic oscillator]]s is given by sine-type functions, so that the time derivatives of the ''Q''<sub>''t''</sub>'s are simple cosine-type functions. Hence ''E'' is given up to a choice of 3''N''&minus;6 initial vibrational amplitudes ''A''<sub>''t''</sub>.
Solution of the GF equations yields ''Q''<sub>''t''</sub> and ''f''<sub>''t''</sub>. Classically, the time dependence of the [[harmonic oscillator (classical)|harmonic oscillator]]s is given by sine-type functions, so that the time derivatives of the ''Q''<sub>''t''</sub>'s are simple cosine-type functions. Hence ''E'' is given up to a choice of 3''N''&minus;6 initial vibrational amplitudes ''A''<sub>''t''</sub>.


The Lagrangian ''L'' = ''T'' &minus; ''V'' is
The Lagrangian ''L'' = ''T'' &minus; ''V'' is
Line 80: Line 81:




In summary, by solving the GF equations the coupled equations have been uncoupled into a set of 3''N''&minus;6 one-dimensional harmonic oscillator equations. These ordinary second-order differential equations are easily solved yielding ''Q''<sub>''t''</sub> as sine-type functions of time, see the article on [[harmonic oscillator (classical)|harmonic oscillator]]s. Classically, one must further specify amplitudes ''A''<sub>''t''</sub> as initial conditions in order to fix the 3''N''&minus;6 vibrational energies completely.
In summary, by solving the GF equations the coupled equations have been uncoupled into a set of 3''N''&minus;6 one-dimensional harmonic oscillator equations. These ordinary second-order differential equations are easily solved yielding the ''Q''<sub>''t''</sub>'s as sine-type functions of time, see the article on [[harmonic oscillator (classical)|harmonic oscillator]]s. Classically, one must further specify amplitudes ''A''<sub>''t''</sub> as initial conditions in order to fix the 3''N''&minus;6 vibrational energies completely. Since the classical vibrational energy of a single mode is given by &frac12; ''f''<sub>''t''</sub> (''A''<sub>''t''</sub>)&sup2; this means that one must fix the initial amount of energy that is contained in each mode.


The uncoupled harmonic oscillator equations can be solved quantum mechanically as well. The energies are (<math>\hbar</math> is [[Planck's constant|Planck's reduced constant]])
The uncoupled harmonic oscillator equations can be solved quantum mechanically as well. The energies are (<math>\hbar</math> is [[Planck's constant|Planck's reduced constant]])
Line 86: Line 87:
E^{(t)}_{v_t} = \hbar \sqrt{f_t}\;\big(v_t+\tfrac{1}{2}\big), \qquad v_t=0,1,2, \ldots
E^{(t)}_{v_t} = \hbar \sqrt{f_t}\;\big(v_t+\tfrac{1}{2}\big), \qquad v_t=0,1,2, \ldots
</math>
</math>
which implies that quantum mechanically 3''N''&minus;6 vibrational quantum numbers ''v''<sub>''t''</sub> must specified in order to fix the vibrational energy of the molecule completely.
which implies that quantum mechanically 3''N''&minus;6 vibrational quantum numbers ''v''<sub>''t''</sub> must specified in order to fix the vibrational energy of the molecule completely. Hence, also quantum mechanically one must establish how much initial energy is present in each mode. Of course, quantum mechanics can account for excitations of vibrational modes by absorption of (infrared) photons. Classical mechanics cannot treat these quantized excitations.


==Normal coordinates in terms of Cartesian displacement coordinates ==
==Normal coordinates in terms of Cartesian displacement coordinates ==
Line 93: Line 94:
the corresponding equilibrium position. Then
the corresponding equilibrium position. Then
<math> \mathbf{d}_A \equiv \mathbf{R}_A -\mathbf{R}_A^0 </math> is by definition the ''Cartesian displacement coordinate'' of nucleus A.
<math> \mathbf{d}_A \equiv \mathbf{R}_A -\mathbf{R}_A^0 </math> is by definition the ''Cartesian displacement coordinate'' of nucleus A.
Wilson's linearizing of the internal curvilinear coordinates ''q''<sub>''t''</sub> expresses the coordinate ''S''<sub>''t''</sub> in terms of the displacement coordinates  
Wilson's linearizing of the internal curvilinear coordinates ''q''<sub>''t''</sub> expresses the coordinate ''S''<sub>''t''</sub> in terms of the displacement coordinates  
:<math>
:<math>
S_t =\sum_{A=1}^N \sum_{i=1}^3 s^t_{Ai} \, d_{Ai}= \sum_{A=1}^N \mathbf{s}^t_{A} \cdot \mathbf{d}_{A}, \quad \mathrm{for}\quad t = 1,\ldots,3N-6,  
S_t =\sum_{A=1}^N \sum_{i=1}^3 s^t_{Ai} \, d_{Ai}= \sum_{A=1}^N \mathbf{s}^t_{A} \cdot \mathbf{d}_{A}, \quad \mathrm{for}\quad t = 1,\ldots,3N-6,  
</math>
</math>
where '''s'''<sub>A</sub><sup>t</sup> is known as a ''Wilson s-vector''.
where '''s'''<sub>A</sub><sup>t</sup> is known as a ''Wilson s-vector''.
If we put the <math>s^t_{Ai}</math> into a 3''N''-6 x 3''N'' matrix '''B''', this equation becomes in matrix language
If we put the <math>s^t_{Ai}</math> into a 3''N''-6 x 3''N'' matrix '''B''', this equation becomes in matrix language
:<math> \mathbf{s} = \mathbf{B} \mathbf{d}. </math>
:<math> \mathbf{s} = \mathbf{B} \mathbf{d}. </math>
The actual form of the matrix elements of '''B''' can be fairly complicated.
The actual form of the matrix elements of '''B''' can be fairly complicated.
Especially for a torsion angle, which involves 4 atoms, it requires tedious vector algebra to derive the corresponding values of the <math>s^t_{Ai}</math>. See for more details on this method, known as
Especially for a torsion angle, which involves 4 atoms, it requires tedious vector algebra to derive the corresponding values of the <math>s^t_{Ai}</math>. See for more details on this method, known as
the ''Wilson s-vector method'', the book by Wilson ''et al.'', or [[molecular vibration]]. Now,  
the ''Wilson s-vector method'', the book by Wilson ''et al.'', or [[molecular vibration]]. Now,  
:<math>
:<math>
\mathbf{Q} = \mathbf{L}^{-1} \mathbf{s} = \mathbf{L}^{-1} \mathbf{B} \mathbf{d} \equiv
\mathbf{Q} = \mathbf{L}^{-1} \mathbf{s} = \mathbf{L}^{-1} \mathbf{B} \mathbf{d} \equiv
Line 111: Line 112:
Q_k = \sum_{A=1}^N \sum_{i=1}^3 D^k_{Ai}\, d_{Ai} \quad \mathrm{for}\quad k=1,\ldots, 3N-6.
Q_k = \sum_{A=1}^N \sum_{i=1}^3 D^k_{Ai}\, d_{Ai} \quad \mathrm{for}\quad k=1,\ldots, 3N-6.
</math>
</math>
Here '''D''' is a 3''N''-6 x 3''N'' matrix which is given by (i) the linearization of the internal coordinates '''q''' (an algebraic process) and (ii) solution of Wilson's GF equations (a numeric process).
Here '''D''' is a 3''N''-6 x 3''N'' matrix which is given by (i) the linearization of the internal coordinates '''q''' (an algebraic process) and (ii) solution of Wilson's GF equations (a numeric process).


==Relation with Eckart conditions ==
==Relation with Eckart conditions ==
Line 122: Line 123:
\sum_{A=1}^N \mathbf{R}^0_A\times \mathbf{s}^t_A= 0,  \quad t=1,\ldots,3N-6.
\sum_{A=1}^N \mathbf{R}^0_A\times \mathbf{s}^t_A= 0,  \quad t=1,\ldots,3N-6.
</math>
</math>
These conditions follow from the [[Eckart conditions]] that hold for the displacement vectors,  
These conditions follow from the [[Eckart conditions]] that hold for the displacement vectors,  
:<math> \sum_{A=1}^N M_A\; \mathbf{d}_{A} = 0 \quad\mathrm{and}\quad
:<math> \sum_{A=1}^N M_A\; \mathbf{d}_{A} = 0 \quad\mathrm{and}\quad
\sum_{A=1}^N M_A\;  \mathbf{R}^0_{A} \times \mathbf{d}_{A} = 0.
\sum_{A=1}^N M_A\;  \mathbf{R}^0_{A} \times \mathbf{d}_{A} = 0.

Latest revision as of 10:52, 3 November 2021

This article is basically copied from an external source and has not been approved.
Main Article
Discussion
Related Articles  [?]
Bibliography  [?]
External Links  [?]
Citable Version  [?]
 
This editable Main Article is under development and subject to a disclaimer.
The content on this page originated on Wikipedia and is yet to be significantly improved. Contributors are invited to replace and add material to make this an original article.

In chemistry and molecular physics, Wilson's GF method, sometimes referred to as the FG method, is a method to obtain certain internal coordinates for a vibrating semi-rigid molecule, the so-called normal coordinates (also known as normal modes). In Wilson's GF method it is assumed that the total molecular energy consists only of vibrational energies, i.e., the overall rotation and translation of the molecule is ignored, as are the motions of its electrons. Neglecting electrons, an N-atomic molecule has 3N degrees of freedom, and, as the description of the overall rotational and translational motions of a molecule requires 6 coordinates, the vibrational problem has 3N−6 degrees of freedom.

The vibrations of the atoms in a molecule are all coupled; roughly speaking, an N-atomic molecule can be seen as a system of 3N−6 coupled one-dimensional harmonic oscillators. Normal coordinates decouple these vibrational motions, thus making it possible to compute the vibrational displacements of the atoms. Further, it is assumed that the vibrations are harmonic, which in general is only approximately the case.

Normal coordinates are often defined in a classical mechanical context, but they appear in the quantum mechanical descriptions of the vibrational motions of molecules as well, and also in the Coriolis coupling between rotations and vibrations.

It follows from application of the Eckart conditions that the inverse of a matrix G gives the vibrational kinetic energy in terms of arbitrary linearized internal coordinates, while F represents the harmonic potential energy in terms of the same internal coordinates. Solving the GF equations gives the matrix that transforms from the linearized internal coordinates to the normal coordinates. The associated energies of the vibrations are obtained after fixing initial amplitudes (classically) or vibrational quantum numbers (quantum mechanically).

Evidently, the method is named after the two matrices F and G that enter the final equations. The name of the most important of its early investigators[1] E. Bright Wilson, Jr. (1908–1992) is often attached to the method.

The GF method

A non-linear molecule consisting of N atoms has 3N−6 internal degrees of freedom, because the positioning of a molecule in three-dimensional space requires three degrees of freedom and the description of its orientation requires another three degrees of freedom. These six degrees of freedom must be subtracted from the 3N degrees of freedom of a system of N particles.

The atoms in a molecule are bound by a potential energy surface (PES) (also known as force field) which is a function of 3N−6 coordinates. The internal degrees of freedom q1, ..., q3N−6 describing the PES in an optimum way are often non-linear; they are for instance valence coordinates, such as bending and torsion angles and bond stretches. It is possible to write the quantum mechanical kinetic energy operator for such curvilinear coordinates, but it is hard to formulate a general theory applicable to any molecule. This is why Wilson[1] linearized the internal coordinates by assuming small displacements. The linearized version of the internal coordinate qt is denoted by St.

The PES V can be Taylor-expanded around its minimum in terms of the St. The third term (the Hessian of V) evaluated in the minimum is a force constant matrix F. In the harmonic approximation—on which the method is based—the Taylor series is ended after this quadratic term. The second term, containing first derivatives, is zero because it is evaluated in the minimum of V. The first term can be included in the zero of energy. Thus,

.

The classical vibrational kinetic energy has the form:

where gst is an element of the metric tensor of the internal (curvilinear) coordinates. The dots indicate time derivatives. Evaluation of the metric tensor g(q) in the minimum q0 of V gives the positive definite and symmetric matrix Gg(q0)−1.

It is assumed that both F and G can be determined. The (inverse of the) kinetic energy matrix G is usually obtained by analytic methods and the force constant matrix F by fitting to experimental infrared data and/or quantum chemical calculations. Knowing the matrices, one can solve the following two problems simultaneously

where is the identity matrix. This can be done because the two equations are equivalent to the generalized eigenvalue problem

where . In numerical mathematics several algorithms are known to solve generalized eigenvalue problems. The solution yields the matrix L and the diagonal matrix Φ that has the vibrational force constants on the diagonal. The matrix L−1 contains the normal coordinates Qk in its rows:

Because of the form of the generalized eigenvalue problem, the method is called the GF method, often with the name of its originator attached to it: Wilson's GF method. It is of interest to point out that by matrix transposition of both sides of the equation and by use of the fact that both G and F are symmetric, (as is a diagonal matrix), one can recast this equation into a very similar one for FG. This is why the method is also referred to as Wilson's FG method.

We introduce the vectors

which satisfy the relation

Upon use of the results of the generalized eigenvalue equation the classical (i.e., non-quantum mechanical) energy E = T + V of the molecule becomes in the harmonic approximation,

Solution of the GF equations yields Qt and ft. Classically, the time dependence of the harmonic oscillators is given by sine-type functions, so that the time derivatives of the Qt's are simple cosine-type functions. Hence E is given up to a choice of 3N−6 initial vibrational amplitudes At.

The Lagrangian L = TV is

The corresponding Lagrange equations are identical to a set of harmonic oscillator (Newton) equations


In summary, by solving the GF equations the coupled equations have been uncoupled into a set of 3N−6 one-dimensional harmonic oscillator equations. These ordinary second-order differential equations are easily solved yielding the Qt's as sine-type functions of time, see the article on harmonic oscillators. Classically, one must further specify amplitudes At as initial conditions in order to fix the 3N−6 vibrational energies completely. Since the classical vibrational energy of a single mode is given by ½ ft (At)² this means that one must fix the initial amount of energy that is contained in each mode.

The uncoupled harmonic oscillator equations can be solved quantum mechanically as well. The energies are ( is Planck's reduced constant)

which implies that quantum mechanically 3N−6 vibrational quantum numbers vt must specified in order to fix the vibrational energy of the molecule completely. Hence, also quantum mechanically one must establish how much initial energy is present in each mode. Of course, quantum mechanics can account for excitations of vibrational modes by absorption of (infrared) photons. Classical mechanics cannot treat these quantized excitations.

Normal coordinates in terms of Cartesian displacement coordinates

Often the normal coordinates are expressed as linear combinations of Cartesian displacement coordinates. Let RA be the position vector of nucleus A and RA0 the corresponding equilibrium position. Then is by definition the Cartesian displacement coordinate of nucleus A. Wilson's linearizing of the internal curvilinear coordinates qt expresses the coordinate St in terms of the displacement coordinates

where sAt is known as a Wilson s-vector. If we put the into a 3N-6 x 3N matrix B, this equation becomes in matrix language

The actual form of the matrix elements of B can be fairly complicated. Especially for a torsion angle, which involves 4 atoms, it requires tedious vector algebra to derive the corresponding values of the . See for more details on this method, known as the Wilson s-vector method, the book by Wilson et al., or molecular vibration. Now,

In summation language:

Here D is a 3N-6 x 3N matrix which is given by (i) the linearization of the internal coordinates q (an algebraic process) and (ii) solution of Wilson's GF equations (a numeric process).

Relation with Eckart conditions

From the invariance of the internal coordinates St under overall rotation and translation of the molecule, follows the same for the linearized coordinates stA. It can be shown that this implies that the following 6 conditions are satisfied by the internal coordinates,

These conditions follow from the Eckart conditions that hold for the displacement vectors,

See this article for more details.

References

Cited reference

  1. 1.0 1.1 E. B. Wilson, Jr. Some Mathematical Methods for the Study of Molecular Vibrations, J. Chem. Phys. vol. 9, pp. 76-84 (1941)

Further references

  • E. B. Wilson, J. C. Decius, and P. C. Cross, Molecular Vibrations, McGraw-Hill, New York, 1955 (Reprinted by Dover 1980).
  • D. Papoušek and M. R. Aliev, Molecular Vibrational-Rotational Spectra Elsevier, Amsterdam, 1982.
  • S. Califano, Vibrational States, Wiley, London, 1976.