Eckart conditions: Difference between revisions

From Citizendium
Jump to navigation Jump to search
imported>Paul Wormer
(This WP text is 100% mine.)
 
imported>Paul Wormer
 
(12 intermediate revisions by 4 users not shown)
Line 1: Line 1:
{{subpages}}
:<small>''This article assumes non-linear molecules everywhere''. </small>
In [[chemistry]] and [[physics]], the '''Eckart conditions''',<ref> C. Eckart, ''Some studies concerning rotating axes and polyatomic molecules'', Physical Review,
In [[chemistry]] and [[physics]], the '''Eckart conditions''',<ref> C. Eckart, ''Some studies concerning rotating axes and polyatomic molecules'', Physical Review,
vol. '''47''', pp. 552-558 (1935).</ref> named after Carl Eckart, sometimes  referred to as '''Sayvetz conditions''',<ref>Aaron Sayvetz, ''The Kinetic Energy of Polyatomic Molecules'', J. Chem. Phys. vol. '''7''', pp. 383-389 (1939).</ref> simplify the nuclear motion (rovibrational)  Schrödinger equation  that arises in the second step of the  [[Born-Oppenheimer approximation]]. The Eckart conditions allow to a large extent the separation of the external (rotation and translation) motions from the internal (vibration) motions.  Although the rotational and vibrational motions of the nuclei in a molecule cannot be fully separated, the Eckart conditions minimize the coupling between these two.   
vol. '''47''', pp. 552-558 (1935).</ref> named after [[Carl Eckart]], sometimes  referred to as '''Sayvetz conditions''',<ref>Aaron Sayvetz, ''The Kinetic Energy of Polyatomic Molecules'', J. Chem. Phys. vol. '''7''', pp. 383-389 (1939).</ref> simplify the nuclear motion (rovibrational)  Schrödinger equation of a [[molecule]]. The rovibrational Schrödinger equation arises in the second step of the  [[Born-Oppenheimer approximation]] (BO approximation).
 
The first step of the BO approximation yields a ''nuclear potential energy''. This potential is generated by solution of the ''electronic'' Schrödinger equation of the molecule for different nuclear constellations. The potential energy, which is a function of the nuclear coordinates,  enters the time-independent  Schrödinger equation for the "motion" of the nuclei&mdash;the second step of the BO approximation. The eigenvalues obtained from the solution of the nuclear Schrödinger equation  are the energies observed in the [[infrared spectrum|infrared]] and [[microwave spectrum|microwave spectra]]  of the molecule. Microwave and infrared spectra are associated with rotations and vibrations of the nuclei in the molecule.
 
In undergraduate texts it is usually assumed&mdash;but not proved&mdash;that rotations and vibrations can be separated, i.e., that nuclear wave functions can be factorized as,
:<math>
\Phi_\mathrm{nuc} =  \Phi_\mathrm{vib}\Phi_\mathrm{rot} \,. \qquad\qquad\qquad(1)
</math>
In these texts the coordinates are never made explicit. It is left to the reader to decide what the coordinates are that the wave functions depend on: space-fixed or body-fixed? If the latter, in what way are they fixed to the body, i.e., to the molecule?  Eckart was the first to derive the coordinates that give the separation of the ''external'' (rotational and translational) motions from the ''internal'' (vibrational) motions. He found that the factorization is never exact, as there are always non-vanishing vibration-rotation coupling terms, the so-called Coriolis terms.  Although the rotational and vibrational motions of the nuclei in a molecule cannot be fully separated, the Eckart conditions minimize the coupling between the two kinds of motion and accordingly make neglect of coupling a lesser  approximation.
 
== Rotational and translational Eckart conditions ==
The Eckart conditions can only be formulated for  [[semi-rigid molecule]]s, which are molecules with a potential energy surface ''V''('''R'''<sub>1</sub>, '''R'''<sub>2</sub>,..'''R'''<sub>''N''</sub>) that has a well-defined minimum for a set of nuclear positions '''R'''<sub>''A''</sub><sup>0</sup> (<math>\scriptstyle A=1,\ldots, N</math>). Tunneling of the nuclei through potential barriers is excluded.
 
The equilibrium coordinates of the nuclei&mdash;with respective masses ''M''<sub>''A''</sub>&mdash;are expressed with respect to an  orthonormal [[principal axes]] frame with origin in the center of mass in the moleculeBy definition, principal axes are eigenvectors of the [[inertia tensor]], which is a symmetric 3&times;3 tensor. Accordingly, the equilibrium coordinates satisfy the relations


== Definition of Eckart conditions ==
The Eckart conditions can only be formulated for a [[semi-rigid molecule]], which is a molecule with a [[potential energy surface]] ''V''('''R'''<sub>1</sub>, '''R'''<sub>2</sub>,..'''R'''<sub>''N''</sub>) that has a well-defined minimum for '''R'''<sub>''A''</sub><sup>0</sup> (<math>A=1,\ldots, N</math>). These equilibrium coordinates of the nuclei&mdash;with masses ''M''<sub>''A''</sub>&mdash;are expressed with respect to a fixed orthonormal principal axes frame and hence satisfy the relations
:<math>
:<math>
\sum_{A=1}^N M_A\,\big(\delta_{ij}|\mathbf{R}_A^0|^2 - R^0_{Ai} R^0_{Aj}\big) =  \lambda^0_i \delta_{ij} \quad\mathrm{and}\quad
\sum_{A=1}^N\, M_A\,\big[\delta_{ij}\,(R_A^0)^2 - R^0_{Ai} R^0_{Aj}\big] =  \lambda^0_i \delta_{ij} \qquad\mathrm{and}\qquad
\sum_{A=1}^N M_A \mathbf{R}_A^0 = \mathbf{0}.
\sum_{A=1}^N M_A \mathbf{R}_A^0 = \mathbf{0}.
</math>  
</math>  
Here &lambda;<sub>i</sub><sup>0</sup> is a principal [[Moment of inertia tensor#Principal moments of inertia|inertia moment]] of the equilibrium molecule.
 
The triplets '''R'''<sub>''A''</sub><sup>0</sup> = (''R''<sub>''A''1</sub><sup>0</sup>,  ''R''<sub>''A''2</sub><sup>0</sup>, ''R''<sub>''A''3</sub><sup>0</sup>) satisfying these conditions, enter the theory as a given set of real constants.
Here &lambda;<sub>i</sub><sup>0</sup> is a principal [[inertia moment]] of the equilibrium molecule.
The ''N'' triplets of real numbers '''R'''<sub>''A''</sub><sup>0</sup> = (''R''<sub>''A''1</sub><sup>0</sup>,  ''R''<sub>''A''2</sub><sup>0</sup>, ''R''<sub>''A''3</sub><sup>0</sup>) satisfying these conditions, enter the theory as a known set of constant parameters.  
 
Following Biedenharn and Louck<ref>L. C. Biedenharn and J. D. Louck, ''Angular Momentum in Quantum Physics'', Addison-Wesley, Reading (1981) p. 535.</ref> we introduce an orthonormal body-fixed frame, the ''Eckart frame'',   
Following Biedenharn and Louck<ref>L. C. Biedenharn and J. D. Louck, ''Angular Momentum in Quantum Physics'', Addison-Wesley, Reading (1981) p. 535.</ref> we introduce an orthonormal body-fixed frame, the ''Eckart frame'',   
:<math>\vec\mathbf{F} = \{ \vec{f}_1, \vec{f}_2, \vec{f}_3\}</math>.  
:<math>\vec\mathbf{F} \equiv \{ \vec{f}_1, \vec{f}_2, \vec{f}_3\} . </math>   
If we were tied to the Eckart frame, which&mdash;following the molecule&mdash;rotates and translates in space, we would observe the molecule in its equilibrium geometry when we would draw the nuclei at the points,  
If an observer were tied to the Eckart frame, which&mdash;following the molecule&mdash;rotates and translates in space, (s)he would constantly see the molecule in its equilibrium geometry, provided the nuclei are at the points,  
:<math>
:<math>
\vec{R}_A^0 \equiv \vec\mathbf{F} \cdot \mathbf{R}_A^0
\vec{R}_A^0 \equiv \vec\mathbf{F}\, \cdot\, \mathbf{R}_A^0
=\sum_{i=1}^3 \vec{f}_i\, R^0_{Ai},\quad A=1,\ldots,N  
=\sum_{i=1}^3 \vec{f}_i\, R^0_{Ai},\quad A=1,\ldots,N .
</math>.
</math>.
Let the elements of '''R'''<sub>''A''</sub> be the coordinates with respect to the Eckart frame of the position vector of nucleus ''A'' (<math>A=1,\ldots, N</math>).  Since we take the origin of the Eckart frame in the instantaneous center of mass, the following relation
The motion of the Eckart frame describes the external (translational and rotational) motion of the equilibrium molecule.  Possible internal motion of the nuclei away from their equilibrium points is the molecular  vibration. The Eckart frame is a non-inertial frame,<ref>An inertial frame is in uniform translational motion.</ref> hence one can expect pseudo (non-inertial) forces to act on the nuclei. Indeed, centrifugal and Coriolis forces will appear.
Let the elements of '''R'''<sub>''A''</sub> be the coordinates with respect to the Eckart frame of the position vector of nucleus ''A'' (<math>\scriptstyle A=1,\ldots, N</math>).  Since we take the origin of the Eckart frame in the instantaneous center of mass, the following relation holds,
:<math>
:<math>
\sum_A M_A \mathbf{R}_A = \mathbf{0}
\sum_A M_A \mathbf{R}_A = \mathbf{0}.
</math>  
</math>  
holds. We define ''displacement coordinates''  
We define ''displacement coordinates''  
:<math>\mathbf{d}_A\equiv\mathbf{R}_A-\mathbf{R}^0_A</math>.
:<math>\mathbf{d}_A\equiv\mathbf{R}_A-\mathbf{R}^0_A</math>.
Clearly the displacement coordinates satisfy the  '''translational Eckart conditions''',  
Clearly the displacement coordinates satisfy the  ''translational Eckart conditions'',  
:<math>
:<math>
\sum_{A=1}^N M_A \mathbf{d}_A = 0 .
\sum_{A=1}^N M_A \mathbf{d}_A = 0 .
</math>
</math>
The '''rotational Eckart conditions''' for the displacements are:
The ''rotational Eckart conditions'' for the displacements are:
:<math>
:<math>
\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,  
</math>
</math>
where <math>\times</math> indicates a [[vector product]].
where &times; indicates a [[cross product]].
These rotational conditions follow from the specific construction of the Eckart frame, see Biedenharn and Louck, ''loc. cit.'', page 538.  
These rotational conditions follow from the specific construction of the Eckart frame vectors <math>\scriptstyle \vec{f}_i</math>, see Biedenharn and Louck, ''loc. cit.'', page 538, for the details of their construction and how this construction leads to the rotational Eckart conditions.  


Finally, for a better understanding of the Eckart frame it may be useful to remark that it becomes a principal axes frame in the case that the molecule is a [[rigid rotor]], that is, when all ''N'' displacement vectors are zero.
Finally, notice that an Eckart frame becomes a principal axes frame in the case that the molecule is a [[rigid rotor]], that is, when all ''N'' displacement vectors are zero, which occurs when the molecule is not vibrating.


==Separation of external and internal coordinates==
==Separation of external and internal coordinates==
The ''N'' position vectors  <math>\vec{R}_A</math> of the nuclei constitute a 3''N'' dimensional linear space '''R'''<sup>3N</sup>: the ''configuration space''. The Eckart conditions give an orthogonal  direct sum decomposition of this space
The ''N'' position vectors  <math>\scriptstyle \vec{R}_A</math> of the nuclei in a molecule constitute a 3''N'' dimensional linear space '''R'''<sup>3N</sup>: the ''nuclear configuration space''. From the Eckart conditions follows that the nuclear configuration space is an orthogonal  direct sum. This direct sum decomposition of the nuclear configuration space gives a separation of the nuclear coordinates in vibration coordinates on the one hand and rotation/translation coordinates on the other.
:<math>
:<math>
\mathbf{R}^{3N} = \mathbf{R}_\textrm{ext}\oplus\mathbf{R}_\textrm{int}.
\mathbf{R}^{3N} = \mathbf{R}_\textrm{ext}\oplus\mathbf{R}_\textrm{int}.
</math>
</math>
The elements of the 3''N''-6 dimensional subspace '''R'''<sub>int</sub> are referred to as ''internal coordinates'', because they are invariant under overall translation and rotation of the molecule and, thus, depend only on the internal (vibrational) motions.  The elements of the 6-dimensional subspace '''R'''<sub>ext</sub> are referred to as ''external coordinates'', because they are associated with the overall translation and rotation of the molecule.


To clarify this nomenclature we define first  a basis for '''R'''<sub>ext</sub>. To that end we introduce the following 6 vectors (i=1,2,3):
The elements of the 3''N''&minus;6 dimensional subspace '''R'''<sub>int</sub> are referred to as ''internal coordinates'', because they are invariant under overall translation and rotation of the molecule and, thus, depend only on the internal (vibrational) motions.  The elements of the 6-dimensional subspace '''R'''<sub>ext</sub> are referred to as ''external coordinates'', because they are associated with the overall translation and rotation of the molecule.
 
To show the direct sum decomposition, we define first  a basis for '''R'''<sub>ext</sub> and introduce the following 6 vectors:
:<math>
:<math>
\begin{align}
\begin{align}
\vec{s}^A_{i}  &\equiv  \vec{f}_i \\
\vec{s}^A_{i}  \equiv  \vec{f}_i \quad\hbox{and}\quad
\vec{s}^A_{i+3} &\equiv  \vec{f}_i \times\vec{R}_A^0 .\\
\vec{s}^A_{i+3} \equiv  \vec{f}_i \times\vec{R}_A^0,\qquad i=1,2,3,\quad A=1,\ldots,N\\
\end{align}
\end{align}
</math>
</math>
An orthogonal, unnormalized, basis for '''R'''<sub>ext</sub> is,
An orthogonal, unnormalized, basis for '''R'''<sub>ext</sub> is then,
:<math>
:<math>
\vec{S}_t \equiv \operatorname{row}(\sqrt{M_1}\;\vec{s}^{\,1}_{t}, \ldots, \sqrt{M_N} \;\vec{s}^{\,N}_{t})
\vec{S}_t \equiv \operatorname{row}(\sqrt{M_1}\;\vec{s}^{\,1}_{t}, \ldots, \sqrt{M_N} \;\vec{s}^{\,N}_{t})
\quad\mathrm{for}\quad t=1,\ldots, 6.
\quad\mathrm{for}\quad t=1,\ldots, 6.
</math>
</math>
A mass-weighted displacement vector can be written as
 
A ''mass-weighted displacement vector'' is
:<math>
:<math>
\vec{D} \equiv \operatorname{col}(\sqrt{M_1}\;\vec{d}^{\,1}, \ldots, \sqrt{M_N}\;\vec{d}^{\,N})
\vec{D} \equiv \operatorname{col}(\sqrt{M_1}\;\vec{d}^{\,1}, \ldots, \sqrt{M_N}\;\vec{d}^{\,N})
Line 70: Line 91:
:<math>\, \vec{S}_i \cdot \vec{D} = \sum_{A=1}^N \; M_A  \big(\vec{f}_i \times\vec{R}_A^0\big) \cdot \vec{d}^{\,A}=\vec{f}_i \cdot \sum_{A=1}^N M_A \vec{R}_A^0 \times\vec{d}^A =  \sum_{A=1}^N M_A \big( \mathbf{R}_A^0 \times \mathbf{d}_A\big)_i = 0,
:<math>\, \vec{S}_i \cdot \vec{D} = \sum_{A=1}^N \; M_A  \big(\vec{f}_i \times\vec{R}_A^0\big) \cdot \vec{d}^{\,A}=\vec{f}_i \cdot \sum_{A=1}^N M_A \vec{R}_A^0 \times\vec{d}^A =  \sum_{A=1}^N M_A \big( \mathbf{R}_A^0 \times \mathbf{d}_A\big)_i = 0,
</math>
</math>
where the zero follows because of the rotational Eckart conditions. We conclude that the displacement vector <math>\vec{D}</math> belongs to the orthogonal complement of '''R'''<sub>ext</sub>, so that it is an internal vector.
where the zero follows because of the rotational Eckart conditions. We conclude that the displacement vector <math>\scriptstyle \vec{D}</math> belongs to the orthogonal complement of '''R'''<sub>ext</sub>, and hence that a mass-weighted displacement vector is an internal vector.


We obtain a basis for the internal space by defining 3''N''-6 linearly independent vectors
A basis for the internal space is obtained by introduction of 3''N''&minus;6 linearly independent vectors
:<math>
:<math>
\vec{Q}_r \equiv \operatorname{row}(\frac{1}{\sqrt{M_1}}\;\vec{q}_r^{\,1}, \ldots, \frac{1}{\sqrt{M_N}}\;\vec{q}_r^{\,N}), \quad\mathrm{for}\quad  r=1,\ldots, 3N-6.
\vec{Q}_r \equiv \operatorname{row}(\frac{1}{\sqrt{M_1}}\;\vec{q}_r^{\,1}, \ldots, \frac{1}{\sqrt{M_N}}\;\vec{q}_r^{\,N}), \quad\mathrm{for}\quad  r=1,\ldots, 3N-6.
</math>
</math>
The vectors <math>\vec{q}^A_r</math> could be  [[GF method#Normal coordinates in terms of Cartesian displacement coordinates|Wilson's s-vectors]] or could be obtained in the harmonic approximation by diagonalizing the Hessian of ''V''.  
The vectors <math>\scriptstyle \vec{q}^A_r</math> could be  [[GF method#Normal coordinates in terms of Cartesian displacement coordinates|Wilson's s-vectors]] or could be obtained in the harmonic approximation by diagonalizing the Hessian of ''V''.  
We next introduce  internal (vibrational) modes,
We next introduce  internal (vibrational) modes,
:<math>
:<math>
Line 82: Line 103:
\quad\mathrm{for}\quad r=1,\ldots, 3N-6.
\quad\mathrm{for}\quad r=1,\ldots, 3N-6.
</math>
</math>
The physical meaning of ''q''<sub>r</sub> depends on the vectors <math>\vec{q}^A_r</math>. For instance, ''q''<sub>r</sub> could be a [[Infrared spectroscopy#Theory|symmetric stretching mode]], in which two C&mdash;H bonds are simultaneously stretched and contracted.
The physical meaning of ''q''<sub>r</sub> depends on the vectors <math>\scriptstyle \vec{q}^A_r</math>. For instance, ''q''<sub>r</sub> could be a [[Infrared spectroscopy#Theory|symmetric stretching mode]], in which two C&mdash;H bonds are simultaneously stretched and contracted.


We already saw that the corresponding external modes are zero because of the Eckart conditions,
We already saw that the corresponding external modes are zero because of the Eckart conditions,
Line 91: Line 112:


==Overall translation and rotation==
==Overall translation and rotation==
The vibrational (internal) modes are invariant under translation and infinitesimal rotation of the equilibrium (reference) molecule if and only if the Eckart conditions apply. This will be shown in this subsection.
As we stated above, the internal (vibrational) coordinates are invariant under translation and infinitesimal rotation of the equilibrium molecule, hence their name. We will now show that these invariances hold if and only if the Eckart conditions apply.  


An overall  translation of the reference molecule is given by
An overall  translation of the reference ( = equilibrium) molecule is given by
:<math> \vec{R}_{A}^0  \mapsto \vec{R}_{A}^0 + \vec{t} </math>'
:<math> \vec{R}_{A}^0  \mapsto \vec{R}_{A}^0 + \vec{t} </math>'
for any arbitrary 3-vector <math>\vec{t}</math>.
for any arbitrary 3-vector <math>\scriptstyle \vec{t}</math>.
An infinitesimal rotation of the molecule is given by
An infinitesimal rotation of the molecule is given by
:<math>  
:<math>  
\vec{R}_A^0  \mapsto \vec{R}_A^0 + \Delta\phi \; ( \vec{n}\times \vec{R}_A^0)
\vec{R}_A^0  \mapsto \vec{R}_A^0 + \Delta\phi \; ( \vec{n}\times \vec{R}_A^0)
</math>
</math>
where Δφ is an infinitesimal angle, Δφ  >> (Δφ)², and <math>\vec{n}</math> is an arbitrary unit vector. From the orthogonality of <math>\vec{Q}_r</math> to the external space follows that the <math>\vec{q}^A_r</math> satisfy
where Δφ is an infinitesimal angle, Δφ  >> (Δφ)², and <math>\scriptstyle \vec{n}</math> is an arbitrary unit vector. From the orthogonality of <math>\scriptstyle \vec{Q}_r</math> to the external space follows that the <math>\scriptstyle \vec{q}^A_r</math> satisfy
:<math>
:<math>
\sum_{A=1}^N \vec{q}^{\,A}_r = \vec{0} \quad\mathrm{and}\quad \sum_{A=1}^N \vec{R}^0_A\times  
\sum_{A=1}^N \vec{q}^{\,A}_r = \vec{0} \quad\mathrm{and}\quad \sum_{A=1}^N \vec{R}^0_A\times  
Line 110: Line 131:
q_r - \vec{t}\cdot\sum_A \vec{q}^{\,A}_r = q_r.
q_r - \vec{t}\cdot\sum_A \vec{q}^{\,A}_r = q_r.
</math>
</math>
Clearly, <math>\vec{q}^A_r</math> is invariant under translation if and only if  
Clearly, <math>\scriptstyle \vec{q}^A_r</math> is invariant under translation if and only if  
:<math>
:<math>
\sum_A \vec{q}^{\,A}_r = 0,
\sum_A \vec{q}^{\,A}_r = 0,
</math>
</math>
because the vector <math>\vec{t}</math> is arbitrary. So, the translational Eckart conditions imply the translational invariance of the vectors belonging to internal space and conversely. Under rotation we have,
because the vector <math>\scriptstyle \vec{t}</math> is arbitrary. So, the translational Eckart conditions imply the translational invariance of the vectors belonging to internal space and conversely. Under rotation we have,
:<math>
:<math>
q_r \mapsto  \sum_A\vec{q}^{\,A}_r \cdot \big(\vec{d}^A - \Delta\phi \; ( \vec{n}\times \vec{R}_A^0) \big) =
q_r \mapsto  \sum_A\vec{q}^{\,A}_r \cdot \big(\vec{d}^A - \Delta\phi \; ( \vec{n}\times \vec{R}_A^0) \big) =
Line 148: Line 169:
= \sum_{A=1}^N M_A \dot{\mathbf{d}}_A\cdot \dot{\mathbf{d}}_A.
= \sum_{A=1}^N M_A \dot{\mathbf{d}}_A\cdot \dot{\mathbf{d}}_A.
</math>
</math>
Because the Eckart frame is non-inertial, the total kinetic energy comprises also centrifugal and Coriolis energies. These  stay out of the present discussion. The vibrational energy is written in terms of the displacement coordinates, which are linearly dependent because they are contaminated by the 6 external modes, which are zero, i.e., the '''d'''<sub>''A''</sub>'s satisfy 6 linear relations. It is possible to write the vibrational energy solely in terms of the internal modes ''q''<sub>r</sub> (''r'' =1, ..., 3''N''-6) as we will now show. We write the different modes in terms of the displacements
Because the Eckart frame is non-inertial (moves non-uniformly in space), the total kinetic energy comprises also centrifugal and Coriolis energies. These kinetic energy terms couple the vibrational with the rotational motions. See [[Molecular Hamiltonian#Watson's nuclear motion Hamiltonian|Watson's Hamiltonian]] for the explicit form of the coupling terms.
 
In this section we will ignore the coupling. Quantum mechanically the neglect of coupling implies that the rotation/translation Schrödinger equation (depending on the external coordinates) can be solved separately from the vibrational Schrödinger equation, which depends solely on the internal coordinates. The nuclear wave function then gets the product form of equation (1).
This is an important computational and conceptual simplification, that however, amounts to neglect of some observable effects on the vibration spectrum.  
 
The vibrational energy is written in terms of the displacement coordinates, which are linearly dependent because they are "contaminated" by the 6 external coordinates, which are zero, i.e., the '''d'''<sub>''A''</sub>'s satisfy 6 linear relations.  
An essential result is the fact that it is possible to write the vibrational energy solely in terms of the internal coordinates ''q''<sub>r</sub> (''r'' =1, ..., 3''N''&minus;6). In order to exhibit this result, we write the different coordinates in terms of the displacements
:<math>
:<math>
\begin{align}
\begin{align}
Line 156: Line 183:
\end{align}
\end{align}
</math>
</math>
The parenthesized expressions define  a matrix '''B''' relating the internal and external modes to the displacements. The matrix '''B''' may be partitioned in an internal (3''N''-6 x 3''N'')  and an external (6 x 3''N'') part,
The parenthesized expressions define  a matrix '''B''' relating the internal and external modes to the displacements. The matrix '''B''' may be partitioned in an internal (3''N''&minus;6 x 3''N'')  and an external (6 x 3''N'') part,
:<math>
:<math>
\mathbf{v}\equiv  
\mathbf{v}\equiv  
Line 204: Line 231:
\end{pmatrix},
\end{pmatrix},
</math>
</math>
where '''G'''<sup>-1</sup> is of dimension (3''N''-6 x 3''N''-6) and '''N'''<sup>-1</sup> is (6 x 6).
where '''G'''<sup>&minus;1</sup> is of dimension (3''N''&minus;6 x 3''N''&minus;6) and '''N'''<sup>&minus;1</sup> is (6 x 6).
The kinetic energy becomes
The kinetic energy becomes
:<math>
:<math>
Line 223: Line 250:
:<math>  \mathbf{H}\mathbf{C} = \mathbf{M} \mathbf{C} \boldsymbol{\Phi},
:<math>  \mathbf{H}\mathbf{C} = \mathbf{M} \mathbf{C} \boldsymbol{\Phi},
</math>
</math>
where  '''H''' is a 3''N'' x 3''N'' symmetric matrix of second derivatives of the potential <math>V(\mathbf{R}_1, \mathbf{R}_2,\ldots, \mathbf{R}_N)</math>. '''H''' is the [[Hessian matrix]] of ''V'' in the equilibrium <math>\mathbf{R}_1^0,\ldots,  \mathbf{R}_N^0</math>. The diagonal matrix '''M''' contains the masses on the diagonal.  
where  '''H''' is a 3''N'' x 3''N'' [[symmetric matrix]] of second derivatives of the potential <math>\scriptstyle V(\mathbf{R}_1, \mathbf{R}_2,\ldots, \mathbf{R}_N)</math>. '''H''' is the [[Hessian matrix]] of ''V'' in the equilibrium <math>\scriptstyle \mathbf{R}_1^0,\ldots,  \mathbf{R}_N^0</math>. The [[diagonal matrix]] '''M''' contains the masses on the diagonal.  
The diagonal matrix <math>\boldsymbol{\Phi}</math> contains the eigenvalues, while
The diagonal matrix <math>\scriptstyle \boldsymbol{\Phi}</math> contains the [[eigenvalue]]s, while
the columns of '''C''' contain the eigenvectors.
the columns of '''C''' contain the [[eigenvector]]s.


It can be shown that the  invariance  of ''V'' under simultaneous translation over '''t'''  of all nuclei implies that vectors '''T''' = ('''t''', ... , '''t''') are in the kernel of '''H'''.
It can be shown that the  invariance  of ''V'' under simultaneous translation over '''t'''  of all nuclei implies that vectors '''T''' = ('''t''', ... , '''t''') are in the kernel of '''H'''.
Line 242: Line 269:


By the definition in this article the non-zero frequency modes are internal modes, since they  are within the orthogonal complement of  '''R'''<sub>ext</sub>. The generalized orthogonalities:
By the definition in this article the non-zero frequency modes are internal modes, since they  are within the orthogonal complement of  '''R'''<sub>ext</sub>. The generalized orthogonalities:
<math> \mathbf{C}^\mathrm{T} \mathbf{M} \mathbf{C} = \mathbf{I} </math>  
<math>\scriptstyle  \mathbf{C}^\mathrm{T} \mathbf{M} \mathbf{C} = \mathbf{I} </math>  
applied to the "internal" (non-zero eigenvalue) and "external" (zero-eigenvalue) columns of '''C'''
applied to the "internal" (non-zero eigenvalue) and "external" (zero-eigenvalue) columns of '''C'''
are equivalent to the Eckart conditions.
are equivalent to the Eckart conditions.
Line 255: Line 282:
* D. Papoušek and M. R. Aliev, ''Molecular Vibrational-Rotational Spectra'', Elsevier (1982).  
* D. Papoušek and M. R. Aliev, ''Molecular Vibrational-Rotational Spectra'', Elsevier (1982).  
* S. Califano, ''Vibrational States'', Wiley, New York-London (1976). ISBN 0 471 12996 8
* S. Califano, ''Vibrational States'', Wiley, New York-London (1976). ISBN 0 471 12996 8
[[Category: CZ Live]]
[[Category: Physics Workgroup]]
[[Category: Chemistry Workgroup]]

Latest revision as of 02:14, 23 February 2010

This article is developed but not approved.
Main Article
Discussion
Related Articles  [?]
Bibliography  [?]
External Links  [?]
Citable Version  [?]
 
This editable, developed Main Article is subject to a disclaimer.
This article assumes non-linear molecules everywhere.

In chemistry and physics, the Eckart conditions,[1] named after Carl Eckart, sometimes referred to as Sayvetz conditions,[2] simplify the nuclear motion (rovibrational) Schrödinger equation of a molecule. The rovibrational Schrödinger equation arises in the second step of the Born-Oppenheimer approximation (BO approximation).

The first step of the BO approximation yields a nuclear potential energy. This potential is generated by solution of the electronic Schrödinger equation of the molecule for different nuclear constellations. The potential energy, which is a function of the nuclear coordinates, enters the time-independent Schrödinger equation for the "motion" of the nuclei—the second step of the BO approximation. The eigenvalues obtained from the solution of the nuclear Schrödinger equation are the energies observed in the infrared and microwave spectra of the molecule. Microwave and infrared spectra are associated with rotations and vibrations of the nuclei in the molecule.

In undergraduate texts it is usually assumed—but not proved—that rotations and vibrations can be separated, i.e., that nuclear wave functions can be factorized as,

In these texts the coordinates are never made explicit. It is left to the reader to decide what the coordinates are that the wave functions depend on: space-fixed or body-fixed? If the latter, in what way are they fixed to the body, i.e., to the molecule? Eckart was the first to derive the coordinates that give the separation of the external (rotational and translational) motions from the internal (vibrational) motions. He found that the factorization is never exact, as there are always non-vanishing vibration-rotation coupling terms, the so-called Coriolis terms. Although the rotational and vibrational motions of the nuclei in a molecule cannot be fully separated, the Eckart conditions minimize the coupling between the two kinds of motion and accordingly make neglect of coupling a lesser approximation.

Rotational and translational Eckart conditions

The Eckart conditions can only be formulated for semi-rigid molecules, which are molecules with a potential energy surface V(R1, R2,..RN) that has a well-defined minimum for a set of nuclear positions RA0 (). Tunneling of the nuclei through potential barriers is excluded.

The equilibrium coordinates of the nuclei—with respective masses MA—are expressed with respect to an orthonormal principal axes frame with origin in the center of mass in the molecule. By definition, principal axes are eigenvectors of the inertia tensor, which is a symmetric 3×3 tensor. Accordingly, the equilibrium coordinates satisfy the relations

Here λi0 is a principal inertia moment of the equilibrium molecule. The N triplets of real numbers RA0 = (RA10, RA20, RA30) satisfying these conditions, enter the theory as a known set of constant parameters.

Following Biedenharn and Louck[3] we introduce an orthonormal body-fixed frame, the Eckart frame,

If an observer were tied to the Eckart frame, which—following the molecule—rotates and translates in space, (s)he would constantly see the molecule in its equilibrium geometry, provided the nuclei are at the points,

.

The motion of the Eckart frame describes the external (translational and rotational) motion of the equilibrium molecule. Possible internal motion of the nuclei away from their equilibrium points is the molecular vibration. The Eckart frame is a non-inertial frame,[4] hence one can expect pseudo (non-inertial) forces to act on the nuclei. Indeed, centrifugal and Coriolis forces will appear.

Let the elements of RA be the coordinates with respect to the Eckart frame of the position vector of nucleus A (). Since we take the origin of the Eckart frame in the instantaneous center of mass, the following relation holds,

We define displacement coordinates

.

Clearly the displacement coordinates satisfy the translational Eckart conditions,

The rotational Eckart conditions for the displacements are:

where × indicates a cross product. These rotational conditions follow from the specific construction of the Eckart frame vectors , see Biedenharn and Louck, loc. cit., page 538, for the details of their construction and how this construction leads to the rotational Eckart conditions.

Finally, notice that an Eckart frame becomes a principal axes frame in the case that the molecule is a rigid rotor, that is, when all N displacement vectors are zero, which occurs when the molecule is not vibrating.

Separation of external and internal coordinates

The N position vectors of the nuclei in a molecule constitute a 3N dimensional linear space R3N: the nuclear configuration space. From the Eckart conditions follows that the nuclear configuration space is an orthogonal direct sum. This direct sum decomposition of the nuclear configuration space gives a separation of the nuclear coordinates in vibration coordinates on the one hand and rotation/translation coordinates on the other.

The elements of the 3N−6 dimensional subspace Rint are referred to as internal coordinates, because they are invariant under overall translation and rotation of the molecule and, thus, depend only on the internal (vibrational) motions. The elements of the 6-dimensional subspace Rext are referred to as external coordinates, because they are associated with the overall translation and rotation of the molecule.

To show the direct sum decomposition, we define first a basis for Rext and introduce the following 6 vectors:

An orthogonal, unnormalized, basis for Rext is then,

A mass-weighted displacement vector is

For i=1,2,3,

where the zero follows because of the translational Eckart conditions. For i=4,5,6

where the zero follows because of the rotational Eckart conditions. We conclude that the displacement vector belongs to the orthogonal complement of Rext, and hence that a mass-weighted displacement vector is an internal vector.

A basis for the internal space is obtained by introduction of 3N−6 linearly independent vectors

The vectors could be Wilson's s-vectors or could be obtained in the harmonic approximation by diagonalizing the Hessian of V. We next introduce internal (vibrational) modes,

The physical meaning of qr depends on the vectors . For instance, qr could be a symmetric stretching mode, in which two C—H bonds are simultaneously stretched and contracted.

We already saw that the corresponding external modes are zero because of the Eckart conditions,

Overall translation and rotation

As we stated above, the internal (vibrational) coordinates are invariant under translation and infinitesimal rotation of the equilibrium molecule, hence their name. We will now show that these invariances hold if and only if the Eckart conditions apply.

An overall translation of the reference ( = equilibrium) molecule is given by

'

for any arbitrary 3-vector . An infinitesimal rotation of the molecule is given by

where Δφ is an infinitesimal angle, Δφ >> (Δφ)², and is an arbitrary unit vector. From the orthogonality of to the external space follows that the satisfy

Now, under translation

Clearly, is invariant under translation if and only if

because the vector is arbitrary. So, the translational Eckart conditions imply the translational invariance of the vectors belonging to internal space and conversely. Under rotation we have,

Rotational invariance follows if and only if

The external modes, on the other hand, are not invariant and it is not difficult to show that they change under translation as follows:

where M is the total mass of the molecule. They change under infinitesimal rotation as follows

where I0 is the inertia tensor of the equilibrium molecule. This behavior shows that the first three external modes describe the overall translation of the molecule, while the modes 4, 5, and, 6 describe the overall rotation.

Vibrational energy

The vibrational energy of the molecule can be written in terms of coordinates with respect to the Eckart frame as

Because the Eckart frame is non-inertial (moves non-uniformly in space), the total kinetic energy comprises also centrifugal and Coriolis energies. These kinetic energy terms couple the vibrational with the rotational motions. See Watson's Hamiltonian for the explicit form of the coupling terms.

In this section we will ignore the coupling. Quantum mechanically the neglect of coupling implies that the rotation/translation Schrödinger equation (depending on the external coordinates) can be solved separately from the vibrational Schrödinger equation, which depends solely on the internal coordinates. The nuclear wave function then gets the product form of equation (1). This is an important computational and conceptual simplification, that however, amounts to neglect of some observable effects on the vibration spectrum.

The vibrational energy is written in terms of the displacement coordinates, which are linearly dependent because they are "contaminated" by the 6 external coordinates, which are zero, i.e., the dA's satisfy 6 linear relations. An essential result is the fact that it is possible to write the vibrational energy solely in terms of the internal coordinates qr (r =1, ..., 3N−6). In order to exhibit this result, we write the different coordinates in terms of the displacements

The parenthesized expressions define a matrix B relating the internal and external modes to the displacements. The matrix B may be partitioned in an internal (3N−6 x 3N) and an external (6 x 3N) part,

We define the matrix M by

and from the relations given in the previous sections follow the matrix relations

and

We define

By using the rules for block matrix multiplication we can show that

where G−1 is of dimension (3N−6 x 3N−6) and N−1 is (6 x 6). The kinetic energy becomes

where we used that the last 6 components of v are zero. This form of the kinetic energy of vibration enters Wilson's GF method. It is of some interest to point out that the potential energy in the harmonic approximation can be written as follows

where H is the Hessian of the potential in the minimum and F, defined by this equation, is the F matrix of the GF method.

Relation to the harmonic approximation

In the harmonic approximation to the nuclear vibrational problem, expressed in displacement coordinates, one must solve the generalized eigenvalue problem

where H is a 3N x 3N symmetric matrix of second derivatives of the potential . H is the Hessian matrix of V in the equilibrium . The diagonal matrix M contains the masses on the diagonal. The diagonal matrix contains the eigenvalues, while the columns of C contain the eigenvectors.

It can be shown that the invariance of V under simultaneous translation over t of all nuclei implies that vectors T = (t, ... , t) are in the kernel of H. From the invariance of V under an infinitesimal rotation of all nuclei around s it can been shown that also the vectors S = (s x R10, ..., s x RN0) are in the kernel of H :

Thus, 6 columns of C corresponding to eigenvalue zero, are determined algebraically. (If the generalized eigenvalue problem is solved numerically, one will find in general 6 linearly independent linear combinations of S and T). The eigenspace corresponding to eigenvalue zero is at least of dimension 6 (often it is exactly of dimension 6, since the other eigenvalues, which are force constants, are never zero for molecules in their ground state). Thus, T and S correspond to the overall (external) motions: translation and rotation, respectively. They are zero-energy modes because space is homogeneous (force-free) and isotropic (torque-free).

By the definition in this article the non-zero frequency modes are internal modes, since they are within the orthogonal complement of Rext. The generalized orthogonalities: applied to the "internal" (non-zero eigenvalue) and "external" (zero-eigenvalue) columns of C are equivalent to the Eckart conditions.

References

  1. C. Eckart, Some studies concerning rotating axes and polyatomic molecules, Physical Review, vol. 47, pp. 552-558 (1935).
  2. Aaron Sayvetz, The Kinetic Energy of Polyatomic Molecules, J. Chem. Phys. vol. 7, pp. 383-389 (1939).
  3. L. C. Biedenharn and J. D. Louck, Angular Momentum in Quantum Physics, Addison-Wesley, Reading (1981) p. 535.
  4. An inertial frame is in uniform translational motion.

The classic work is:

  • E. Bright Wilson, Jr., J. C. Decius and Paul C. Cross, Molecular Vibrations, Mc-Graw-Hill (1955). Reprinted by Dover (1980).

More advanced book are:

  • D. Papoušek and M. R. Aliev, Molecular Vibrational-Rotational Spectra, Elsevier (1982).
  • S. Califano, Vibrational States, Wiley, New York-London (1976). ISBN 0 471 12996 8