Вы находитесь на странице: 1из 11

PAPER

www.rsc.org/pccp | Physical Chemistry Chemical Physics

The molecular potential energy surface and vibrational energy levels of methyl uoride. Part IIw
Steven A. Manson, Mark M. Law,* Ian A. Atkinson and Grant A. Thomson
Received 28th February 2006, Accepted 19th April 2006 First published as an Advance Article on the web 31st May 2006 DOI: 10.1039/b603108k New analytical bending and stretching, ground electronic state, potential energy surfaces for CH3F are reported. The surfaces are expressed in bond-length, bond-angle internal coordinates. The four-dimensional stretching surface is an accurate, least squares t to over 2000 symmetrically unique ab initio points calculated at the CCSD(T) level. Similarly, the vedimensional bending surface is a t to over 1200 symmetrically unique ab initio points. This is an important rst stage towards a full nine-dimensional potential energy surface for the prototype CH3F molecule. Using these surfaces, highly excited stretching and (separately) bending vibrational energy levels of CH3F are calculated variationally using a nite basis representation method. The method uses the exact vibrational kinetic energy operator derived for XY3Z systems by Manson and Law (preceding paper, Part I, Phys. Chem. Chem. Phys., 2006, 8, DOI: 10.1039/ b603106d). We use the full C3v symmetry and the computer codes are designed to use an arbitrary potential energy function. Ultimately, these results will be used to design a compact basis for fully coupled stretchbend calculations of the vibrational energy levels of the CH3F system.

1. Introduction
Methyl uoride is a very important prototype for experimental and theoretical studies in spectroscopy,110 structure and bonding,1117 intramolecular dynamics3,4,6,1720 and reaction dynamics.17,2123 The spectroscopy and dynamics of a ve-atom molecule such as methyl uoride present a number of major challenges to both theory and experiment. There are nine internal degrees of freedom associated with the nuclear motion. Thus, the molecular potential energy function, which governs the vibrationrotation dynamics, is a nine-dimensional (9D) hypersurface. Its full exploration using ab initio electronic structure methods, including congurations probed near dissociation at spectroscopic accuracy, is still an impossible task. For the simplest ve-atom system, methane, the best ab initio potential energy surface is reliable up to only 13 000 cm1 above the minimum energy conguration, which is one third of the dissociation energy.24,25 Potential energy surfaces that do represent dissociation for methane exist, but they combine experimental and ab initio data.26 Similarly, accurate variational solutions of the 9D vibrational Schrodinger equation for methane have been attempted only very recently.24,25,2732 To our knowledge, only one attempt has been made to calculate variationally and in full dimension the vibrational energy levels of CH3F.2 However, there were a number of severe restrictions on the method described: the kinetic energy operator (KEO) was expressed in normal coordinates, which are only suitable for studying small amplitude vibrations; terms in the Hamiltonian operator were

Chemistry Department, University of Aberdeen, Meston Walk, Aberdeen, UK AB24 3UE. E-mail: m.m.law@abdn.ac.uk w For Part I see ref. 41.

neglected, the most serious of which was the restriction of the potential energy to terms involving at most three normal coordinates (this breaks the C3v symmetry and results in degenerate levels being split by as much as 10 cm1 for overtones of degenerate modes). Most earlier ab initio studies of the potential energy surface of methyl uoride have taken one of two approaches: the rst involves exploring, at a modest level of theory, a region close to the equilibrium conguration that can be described adequately with a low-order Taylor expansion in normal mode or other internal coordinates;2,11,13,15 the second approach explores, at a high level of theory, a few critical points on the potential energy surface.14,16,17,21,22 A third approach, taken by Luckhaus and Quack,4 was to focus on the three-dimensional potential energy surface of the CH chromophore in CHD2F. Here we will consider, using a relatively high level of theory, two reduced subspaces of the methyl uoride potential energy surface (ve-dimensional (5D) bending and fourdimensional (4D) stretching) each including large distortions from the equilibrium geometry. We consider determination of these surfaces to be an important rst step towards determining a full 9D ab initio potential energy surface for this prototype molecule. The choice of separating bending and stretching is motivated by our intention to tackle the vibrational calculations by solving initially separate bending and stretching vibrational Schrodinger equations (before combining the eigen functions of the low-dimension Hamiltonians to form a basis to tackle the full 9D problem). This approach has already been taken for methane.2729,31 We note that this may not be the optimum strategy since in the methyl halides (and in methane) the CH stretching modes are very strongly coupled to the bending modes by Fermi resonances.1,3,6,19
Phys. Chem. Chem. Phys., 2006, 8, 28552865 | 2855

This journal is

 c

the Owner Societies 2006

Knowledge of the molecular potential energy surface is a prerequisite for determining vibrationrotation (and other) properties. On the other hand, the ability to compute vibrationrotation energy levels (and transition intensities) opens up the possibility of extracting detailed information on the potential energy surface from experimental high resolution spectroscopic data.33 Previous eorts to derive potential energy surface information for methyl uoride from spectroscopic data12,15 have relied on perturbation theory or other approximate methods to compute ro-vibrational properties. The application of perturbation theory to methyl uoride (and its isotopomers) is severely complicated by the presence of many Fermi and Coriolis resonances.15 In this context, it is important to note that methyl uoride is an important prototype for developing and testing new ab initio electronic structure methods.13,34 High resolution spectroscopic data aord the most stringent tests for comparison with ro-vibrational properties computed from an ab initio molecular potential energy surface. The ability to calculate rotationally and vibrationally excited bound-state energies and wavefunctions precisely and eciently for molecules such as methyl uoride would undoubtedly overcome the limitations currently imposed by the enforced reliance on perturbation theory. Ultimately, our goal is to develop a computational method to solve the ro-vibrational Schrodinger equation for species such as CH3F accurately and over a wide energy range. The accomplishment of such methods for three-atom33,35,36 and four-atom systems33,37,3840 have resolved dicult dynamical problems, including the interpretation of high resolution molecular spectra, and have allowed the determination of accurate potential energy surfaces (PESs) by tting to such data. The structure of the present paper is as follows. Details of the determination of the ab initio potential energy at points on extensive grids in stretching and bending subspaces are given in Sections 2.1 and 2.2. The subsequent tting of suitable analytical functions for the stretching and bending surfaces are reported in Sections 2.3 and 2.4 along with discussion of the accuracy and range of the two surfaces. Our stretching and bending vibrational energy level calculations are described in Section 3. These calculations make use of the exact vibrational kinetic energy operator derived in Part I.41

investigation of ve stationary points on the potential energy surface for the CH3Cl + F - CH3F + Cl, SN2 reaction.21 For the present purpose, calculating the potential energy at a wide range of congurations of the CH3F molecule, this basis set (and CCSD(T) level of theory) provides a good compromise between accuracy and speed of calculation. A single point energy calculation (at a C3v symmetry conguration) typically took about 13 CPU minutes on a Compaq Alpha XP1000/667 workstation. Calculations at geometries distorted away from C3v symmetry took up to four times longer. Test calculations using the full aug-cc-pVQZ basis took over ve times longer than the corresponding calculations using the (A)VT/QZ basis with only marginal gains in reproduction of the experimental geometry. The aug-cc-pVQZ basis was not considered further. The optimised equilibrium geometry determined by the CCSD(T)/(A)VT/QZ method is: RCH = 1.0899 A, RCF = e e and bHCF = 108.721. This agrees reasonably well 1.3872 A e with the experimental geometry: RCH = 1.0870 A, RCF = e e 1.3827 A and bHCF = 108.671.14 e The harmonic wavenumbers for 12CH3F (o1. . .o6) computed using the CCSD(T)/(A)VT/QZ method are 3049.10, 1494.21, 1065.70, 3140.49, 1512.35, 1205.35 cm1. All of these except o1 agree with the empirical values within the estimated uncertainties associated with the latter.12 We note that the wavenumber o1 is the most strongly aected by Fermi resonance and other anharmonic corrections.12 2.2. Calculation of 4D stretching and 5D bending ab initio grids The computational demands of calculating multidimensional ab initio PESs make it impractical to carry out the full calculation on a single workstation. We have used the massively parallel CSAR supercomputer facilities at Manchester University in order to determine the required ab initio energies. The surfaces were initially determined by performing electronic structure calculations at a large number of molecular geometries to generate 4D stretching and 5D bending grids of data points. CH3F possesses three equivalent hydrogen atoms and so there are a number of geometries of the molecule which, by symmetry, will all have the same energy. Thus we calculated only symmetrically unique points. The geometry of the molecule was specied using four internal bond vectors (three CH and one CF) whose relative orientation was dened using the polyspherical coordinates b and f as described in Part I.41 Briey, the radial coordinates t1, t2, t3 and t4 correspond to the three CH bond lengths and the CF bond length, respectively. The angle bi is the angle between the CF vector and the CH vector for Hi (for i = 1,2 and 3). The angle f3 is the dihedral angle between the H1CF and H2CF planes whilst f2 is the dihedral angle between the H3CF and H1CF planes. We have calculated the energy at over 2000 (symmetrically unique) stretching points and 1200 (symmetrically unique) bending points. In choosing the grids of points, the following constraints have been employed: (i) All points are less than 30 000 cm1 above the equilibrium energy.
This journal is
 c

2. Potential energy surfaces


2.1. Computational method and basis set All our ab initio calculations were undertaken at the CCSD(T) level of theory using the MOLPRO package.42 We have used the BornOppenheimer approximation and consider only the ground electronic state. We have not attempted to make relativistic nor other, small, corrections.43,44 We denote the basis set used (A)VT/QZ. It comprises the (s,p,d,f) functions of Dunnings aug-cc-pVQZ basis for carbon and uorine, with the (s,p) functions of the aug-cc-pVTZ basis plus the (d) function of the cc-pVTZ basis for hydrogen. This gives a basis set consisting of 178 contracted Gaussian-type orbitals in total. This basis set has been used successfully in an
2856 | Phys. Chem. Chem. Phys., 2006, 8, 28552865

the Owner Societies 2006

For the stretching surface: (ii) The b and f angles are held at their calculated equilibrium values; (iii) The four bond lengths are varied in the range 0.40 to +1.0 A from their equilibrium values; (iv) Points are calculated at regular 0.1 A intervals in each dimension, with the interval being reduced to 0.05 A close to equilibrium. For the bending surface: (v) All four bond length vectors are held at their calculated equilibrium values; (vi) The HCF angles (b) are varied in the range 551 to 1701, for each of the 33 unique combinations of interplanar angles (f)the latter ranging from 501 to +501 from equilibrium; (vii) For the f angles, points are calculated at 101 intervals; (viii) For the b angles, ve points spaced at equal energy on either side of the calculated equilibrium b value are used, [55.0, 60.0, 66.0, 74.0, 84.0, 136.0, 147.0, 156.0, 164.0, 170.0]1, giving 11 points in each b dimension. In order to produce an analytical representation of the potential energy surface for use in dynamical calculations, it was necessary to t the ab initio energies computed at the points discussed above to suitable functional forms. Fitting molecular potential energy hypersurfaces is a challenging problem,19,45,46 but the powerful interactive non-linear least squares tting program I-NoLLS45 is available and has been used to t suitable functional forms to the ab initio grid points. We have exploited the full C3v symmetry in the ts. The tted potential energy surfaces are available as FORTRAN subroutines from the corresponding author on request. 2.3. Stretching potential energy surface

Thus, 53 symmetrically unique parameters were optimised to t the 2028 equally weighted data points, with an average deviation of 1.8 cm1. The ab initio energies are tted to within a few cm1 except for above 20 000 cm1, where there are 12 points deviating by between 10 and 40 cm1. The resulting surface has been examined carefully and reects the correct physical behaviour at short CH and CF bond lengths. The surface also shows realistic behaviour at large stretching geometries, although we do not expect it to represent dissociation accurately. 2.4. Bending potential energy surface

A power series expansion was also used in tting a functional form to the bending ab initio potential energies. However, the bending case is complicated by the presence of the angular redundancy. In order to fully exploit the C3v symmetry during the t it was necessary to reintroduce the redundant angle into the expression X Vb1 ; b2 ; b3 ; f1 ; f2 ; f3 Ci;j;k;l;m;n f b1 i f b2 j f b3 k
i;j;k;l;m;n

f f1 l f f2 m f f3 n 3 where f (bn) = (bn bne) and f (fn) = (fn fne) (5) (4)

The functional type selected to t the stretching potential energies has the following form X Ci;j;k;l f r1 i f r2 j f r3 k f r4 l 1 Vr1 ; r2 ; r3 ; r4
i;j;k;l

where i, j, k and l give the degree of the expansion in each coordinate, with i, j and k being the three CH indices and l the index for the CF mode and f(rn) = 1 exp [an(rn rne)]. (2)

In tting the stretching PES, the Ci,j,k,l and an parameters were adjusted in the t, with the rne being held at the values corresponding to our calculated ab initio equilibrium geometry. Only symmetrically unique Ci,j,k,l coecients were included in the t. For example, in the most general case there are six symmetrically equivalent coecients that may be obtained by permutation of the i, j and k indices, but only one representative coecient was varied independently. The indices were varied to include all permutations for which i + j + k + l r 6, with the C0000 parameter xed at zero (taking the equilibrium as the zero of energy). Any adjustable parameters with a 95% condence value greater than that of the parameter itself were xed to zero before continuing with the t. Points with energy 429 000 cm1 above equilibrium were excluded because of the diculty reproducing such points accurately.
This journal is
 c

The Ci,j,k,l,m,n parameters were oated in the t, with bne and fne being held at our calculated ab initio equilibrium geometry. Inclusion of the redundant angle leads to complications in the tting procedure. For example, it can be shown that the potential is always invariant with respect to terms of the type Ca,a,a,b+1,b,b. Similarly, it can be shown that a number of terms are not linearly independent, for example, C1,1,0,3,0,0 and C1,1,0,2,1,0. In the case of these systematically correlated parameters, it was arbitrarily decided to remove the one with the highest individual index before proceeding with the t. As in the stretching case, parameters with a 95% condence limit greater than the value of the parameter itself were also removed before continuing with the t. The indices were chosen such that i + j + k r 10 and l + m + n r 6, although only two parameters with i + j + k 4 8 were included. In addition, due to the large number of possible parameters, the overall sum of the indices was restricted such that i + j + k + l + m + n r 12. The C000000 parameter was held xed at zero. Using 259 symmetrically unique adjustable parameters, the 1224 equally weighted data points were tted with an average deviation of 11.6 cm1. The bending PES is not tted as well as the stretching one. This is not unexpected: not only does the extra dimension in the bending surface make it more challenging but it also involves tting two dierent types of coordinate. The fact that a large number of Ci,j,k,l,m,n, parameters are required to t the surface to this level of accuracy indicates that the functions used in the t (eqn (4) and (5)) lack some of the exibility, and indeed basic shape, required to t this type
Phys. Chem. Chem. Phys., 2006, 8, 28552865 | 2857

the Owner Societies 2006

of surface to within a few cm1. It may be that the use of cosine or Legendre functions would prove to be a better choice for the b anglesthese functions having been used for similar angular coordinates of smaller molecules.43 However, the dierence in performance from the function in eqn (4) was not found to be signicant in lower dimensional test cases. For the f coordinates (unlike the stretching case), there appears no obvious suitable functional form to use. This is a problem that will require further investigation in order to t the bending surface with the same accuracy as the stretching one. Although the bending PES is not tted as accurately as the stretching PES, it is still well within the accuracy of, for example, the ammonia potential used in ref. 47. It is valid for use in a variational calculation, up to the 30 000 cm1 region that the ab initio points have been determined for. However, outside this region there are areas of the surface that behave unphysically. In particular, for areas of the potential corresponding to two or more atoms approaching each other: intuitively, such regions should have extremely large and positive energies. However, in our tted potential this is not the case. Consequently, if the potential is to be used in dynamical calculations it is necessary to take care in dealing with these unphysical regions. In the variational calculations reported in Section 3.3 below, a scheme of basis set contractions is used this has the eect of localising the basis functions used in the nal stage of the calculation, so as to avoid the unphysical regions. We note that for 1D cuts through the potential (where all other coordinates are held at their equilibrium values) the surface does behave physically realistically at highly distorted geometries. Nevertheless, an important piece of future work would be to consider re-tting the bending PES (and extending our angular grid) in an attempt to obtain a more accurate t and to also eliminate the unphysical regions.

corresponding rows of M are (g + e)mHj for the heavy atom and djk + mHje for the light atoms, where mC and mHj are the masses of the atoms. The total mass of the molecule is given by mT with g = [1 (mT/mC)]1/2]/SjmHj and e = (gmC + 1)/mT.48 Once the M matrix is constructed, it is straightforward to obtain its inverse and hence the V matrix to be used in eqn (6) of Part I41 to construct the internal coordinates. The V matrix (obtained using the following masses (in atomic mass units): mH = 1.007 825, mC = 12.0, mF = 18.998 403) that denes the Radau/Jacobi orthogonal coordinate system used in this work is (not to full precision and omitting the nal column, which locates the centre of mass) 0 1 0:964576 0:035424 0:035 424 0:067083 B 0:035424 0:964576 0:035 424 0:067083 C B C V B 0:035424 0:035424 0:964576 0:067083 C: B C @ 0:000000 0:000000 0:000000 1:000000 A 0:893728 0:893728 0:893 728 0:798750 6 The rst three rows correspond to the hydrogen atoms, the fourth to the uorine atom and the last to the carbon atom. Our internal coordinates are now dened as the four vector lengths, r1,. . .,r4, three angles between the Jacobi vector and each of the Radau vectors, b1,. . .,b3 and two angles between planes, f2 and f3. 3.2. Stretching vibrational energy levels

3. Vibrational energy levels


3.1. Coordinate system In order to simplify the KEO (see Part I41), we have used coordinates based on orthogonal internal vectors. Three Radau vectors are used to represent the CH3 subunit (this allows the C3v symmetry of the molecule to be exploited). A Jacobi vector then connects the F atom with the centre of mass of the CH3 subunit. Our KEO is derived in terms of translationfree internal vectors dened by a matrix V.41 The V matrix for the present choice of orthogonal vectors is obtained from the inverse of the M matrix of mass factors, which itself is formed following the prescription of Schwenke48 as outlined below. The M matrix is a 5 5 matrix of mass factors that gives the Cartesian coordinates in terms of the internal coordinates (the inverse of the relationship in eqn (6) of Part I41). The Jacobi and Radau vectors are then constructed as follows. A Jacobi vector connects the centre of mass of object A to object B, where A and B are atoms or collections of atoms having masses mA and mB (here A is the CH3 subunit and B the F atom). In this case, the entries of the corresponding row of M are mB/(mA + mB) for the atoms in object A and mA/(mA + mB) for the atoms in object B.48 The three Radau vectors (numbered with an index k = 1, 2, 3) describe the relative positions of the heavy central C atom and the three light atoms Hj, j = 1,. . .,3. The entries of the
2858 | Phys. Chem. Chem. Phys., 2006, 8, 28552865

The variational calculations of the stretching energy levels of the C3v penta-atomic molecule CH3F make use of the vibrational KEO derived in Part I41 and the stretching ab initio PES described in Section 2.3. Although we discuss CH3F specically below, our variational method is applicable to any similar XY3Z system (including other methyl halides, and species such as CH3D, CHD3, SiH3D and SiH3F). 3.2.1. Hamiltonian. The stretching KEO is expressed in terms of the orthogonal coordinate system detailed in Section 3.1 above. For the stretching-only calculation, the bi and fj angles are held at their equilibrium values. From eqn (2) and Table 1 of Part I,41 the stretching KEO is given by ! X 2 X 2 h2 N 1 Vai @ 2 N 1 2Vai @ ^str  T 7 2 ma @r2 ma ri @ri i i i where there is an implied summation over a = 1,. . .,N, where N is the number of atoms. Using this KEO and adding the potential energy, the matrix elements of the Hamiltonian are calculated as R * str 2 ci Hij cjr1. . .r2 dr1. . .drN1. (8) N1 However, as is commonly undertaken in order to facilitate the evaluation of these integrals,4952 the radial part of the Jacobian is incorporated into the operator and basis functions c as follows: c 0 = r1r2r3r4c ^0 ^ T str r1 r2 r3 r4 T str   1 : r1 r2 r3 r4 (9) 10

This journal is

 c

the Owner Societies 2006

The resultant form of the KEO is  h ^0 T str 2


2 N 1 X i 2 Vai

@ ; ma @r2 i

1 jabcx; A2 i p jabcxi jacbxi jbacxi 6 11 jcabxi jbcaxi jcbaxi 1 jabcx; 1Ea i p 2jabcxi 2jacbxi jbacxi 12 jcabxi jbcaxi jcbaxi 1 jabcx; 1Eb i p jbacxi jcabxi jbcaxi jcbax 4 1 jabcx; 2Ea i p jbacxi jcabxi jbcaxi jcbaxi 4 1 jabcx; 2Eb i p 2jabcxi 2jabcxi jbacxi 12 jcabxi jbcaxi jcbaxi 14 Apart from the |abcx, 2Ebi function, which we nd takes the opposite sign, our symmetrised functions agree with Halonen and Child.54 This phase dierence was not signicant in the work presented in ref. 54 but matters here because of our method of computing the potential integrals, described in the following section. 3.2.3 Potential matrix elements. The integration of the potential matrix elements is carried out numerically in four dimensions because of our assumption of an arbitrary form for the stretching PES. We use the standard GaussLaguerre relation55 to evaluate the integrals as a sum of products of weights (wi) and the value of the function f at M quadrature points (yi) Z
0 1

with the same implied summation over a. Matrix elements are now evaluated via the integral R 0* 0str 0 (12) ci Hij cjdr1. . .drN1. 3.2.2. Basis functions, symmetry and kinetic energy matrix elements. We have used Morse-oscillator-like basis functions and evaluated the kinetic energy integrals analytically.53 We have made no assumption about the form of the potential energy and so it is necessary to evaluate the potential matrix elements numerically, as described in Section 3.2.3. A general 4D stretching basis function is denoted |abcxi, indicating x stretching quanta in the mode with vector length r4 (nominally the CF mode here) and a, b, c quanta in the three modes with vector lengths r1, r2 and r3 (nominally CH modes here). In order to symmetrise these functions, it is necessary to consider three distinct classes: Class I, where the number of quanta in all three CH modes is the same; Class II, where the number of quanta in two CH modes is the same; and Class III, where the number of quanta in all three CH modes is dierent. For the purpose of generating our basis set, we use the convention a Z b Z c and thus obtain an additional, but symmetrically identical, member of Class II. The symmetries spanned by the three distinct classes are Class I |aaaxi A1 Class IIa |abbxi IIb |aabxi A1 + E Class III |abcxi A1 + A2 + 2E. Symmetrisation of these basis functions can be achieved via the promotion operator technique, employed on a number of systems by Halonen and Child.54 A general symmetrised basis function for the present C3v system is given by ref. 54 X jabcx; Gi N 1=2 Ci jii: 13
i

ey ya f ydy

M X i1

wi f yi :

15

The points and weights are calculated using a modied version53 of the program given by Stroud and Secrest.55 Using eqn (15), the potential integrals for the four dimensional stretching problem have the form
M1 M1 M1 M2 XXXX i j k l

Where G is the symmetry species, N the normalisation constant, Ci the expansion coecient and |ii the basis member. The wavefunctions for each of the three symmetry classes are given below. | aaax,A1i = | aaaxi 1 jabbx; A1 i p jabbxi jbabxi jbbaxi 3 1 jabbx; Ea i p 2jabbxi jbabxi jbbaxi 6 1 jabbx; Eb i p jbabxi jbbaxi 2 1 jabcx; A1 i p jabcxi jacbxi jbacxi 6 jcabxi jbcaxi jcbaxi
This journal is
 c

habcx; GjVja0 b0 c0 x0 ; Gi

wri ; rj ; rk ; rl

cG ri ; rj ; rk ; rl cG0 b0 c0 x0 ri ; rj ; rk ; rl abcx a Vri ; rj ; rk ; rl : 16 Here, (ri, rj, rk, rl) denotes a point in the 4D quadrature grid, w (ri, rj, rk, rl) the corresponding product of weights and cG a symmetrised wavefunction. Calculating the potential matrix elements in this way is computationally demanding and inecient, as it involves repeated evaluation of the potential at points related by symmetry. For example, in the most general case where indices i, j, k correspond to the three CH modes, there are six points (corresponding to the permutations of the i, j, k indices) which return the same value of the potential. Consequently, we make
Phys. Chem. Chem. Phys., 2006, 8, 28552865 | 2859

the Owner Societies 2006

use of a slightly adapted form of the symmetrised quadrature expression for XY4 molecules given by Xie and Tennyson27
j M2 M1 i XXXX i j k l

habcx; GjVjja0 b0 c0 x0 ; Gi

w0 ri ; rj ; rk ; rl

Vri ; rj ; rk ; rl
h X a1

cGa ri ; rj ; rk ; rl cG0ab0 c0 x0 abcx a

ri ; rj ; rk ; rl : 17 The quantity h is the degree of degeneracy and the new weight is related to the unsymmetrised weight expression via w 0 (ri, rj, rk, rl) = w (ri, rj, rk, rl)ch1, (18)

where c is the number of equivalent geometries (1, 3 or 6). As a result of mixing the two components (Ga) of the degenerate E representation, the expression for the symmetrised quadrature is more complicated than eqn (16). The use of this symmetrised quadrature formula reduces not only the computational time required to perform the integration but also reduces the memory requirement, because it is no longer necessary to store the wavefunction at symmetrically related geometries. The extent of the saving is dependent on the problem under investigation, but the use of eqn (17) can produce up to a six-fold saving in computational demand. The basis set is characterised by the number N, which represents the maximum order of the basis function in any dimension. In addition, the constraint Sini r N is applied, where ni is the number of quanta in the ith mode. In the next two sections, CH3F stretching energy levels calculated using our 4D ab initio stretching PES are reported. 3.2.4. Computational details. Morse-oscillator-like functions may be variationally optimised for the system under investigation by altering the Morse parameters oe, re and De. The parameters oe, re and De are the harmonic frequency, the equilibrium radial coordinate and the dissociation energy of the corresponding Morse potential energy curve, respectively. For the calculations on CH3F reported here, the optimised Morse parameters are: oe = 2900 cm1, re = 1.130 A and 1 De = 35 000 cm for the CH modes; and oe = 1100 cm1, re = 1.431 A and De = 38 000 cm1 for the CF mode. Our vibrational variational procedure uses the orthogonal coordinates discussed above. However, our stretching potential is calculated and tted in terms of bond length coordinates (in each case the angular coordinates, b and f, are held xed at the appropriate equilibrium values). A problem arises in that it is not possible to convert between the two coordinate systems (either by retting the PES or by converting each 4D orthogonal grid point to the corresponding bond length, bond angle grid point) in the reduced dimensionality stretching-only problem. This is because converting between the two coordinate systems also changes the angular coordinates. Consequently, a full 9D bond length, bond angle PES would be required in order to correctly carry out the conversion.
2860 | Phys. Chem. Chem. Phys., 2006, 8, 28552865

Of course, we could have calculated our 4D stretching and 5D bending CH3F potentials using the same orthogonal coordinates used for vibrational energy level calculations. However, as already discussed, the approach was to determine reduced dimensionality PESs as an important rst step to obtaining an accurate, full 9D ab initio surface for CH3F. Consequently, the PESs were determined at geometries dened by bond length, bond angle coordinates because it was felt that such coordinates would be more generally applicable than the orthogonal coordinate system. This does mean that a calculation using the variational procedure described here and our present stretching PES is not self-consistent and can not be readily made so. However, we believe that the error introduced, especially for low lying levels, is comparable with the dierence that will arise between the calculated and observed energy levels as a result of a lack of stretchbend coupling. In fact, as has been similarly noted in recent stretch- or bend-only calculations of methane,2729 the importance of stretchbend coupling in CH3F means there is little to be gained from a straightforward comparison of a large number of calculated and observed energy levels. However, comparison of at least a few calculated and observed quantities (whilst remaining aware of the discrepancy caused by the lack of stretchbend coupling) is still useful in that it provides evidence that the variational code and the ab initio PES produce physically realistic energy levels. 3.2.5. Results. Table 1 contains CH3F stretching energy levels of all symmetries up to 10 000 cm1 above the ground state. All levels given are converged to 0.1 cm1 or better. The
Table 1 Vibrational stretching energies for CH3F (in cm1) calculated using orthogonal coordinates. Excited state energies are given relative to the ground state Symmetry Ground state I=1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 A1 A1 A1 E A1 A1 E A1 A1 E A1 A1 A1 E E A1 E A1 A1 A1 E E A1 E A1 A1 A1 E E Energy 4997.8 1059.3 2105.5 2879.3 2913.9 3138.7 3939.7 3973.9 4159.2 4986.9 5020.4 5167.1 5666.3 5671.1 5774.7 5814.2 6021.1 6053.3 6163.0 6727.8 6732.5 6836.0 6875.0 7042.5 7072.1 7147.6 7776.0 7780.6 7883.9 I = 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 Symmetry A1 E A1 A1 E A1 A2 E A1 E A1 A1 E E A1 E A1 A1 E A1 A2 A1 E A1 A1 E E A1 Energy 7922.1 8051.3 8074.0 8124.2 8346.6 8346.9 8534.3 8550.8 8588.1 8589.4 8695.6 8811.1 8815.5 8918.6 8955.3 9047.6 9051.2 9101.2 9409.4 9409.7 9597.0 9650.4 9651.6 9757.5 9833.2 9837.4 9939.9 9972.8

This journal is

 c

the Owner Societies 2006

Table 2 Vibrational stretching energies for CH3F (in cm1) calculated using bond-length coordinates and the N = 11 basis. Excited state energies are given relative to the ground state Symmetry Ground state I=1 2 3 4 5 6 7 8 A1 A1 A1 A1 E A1 A1 E A1 Energy 5208.4 1124.4 2234.1 2941.4 3032.1 3329.2 4066.9 4158.0 4409.8 I=9 10 11 12 13 14 15 16 17 Symmetry A1 E A1 A1 E A1 E A1 E Energy 5177.6 5269.0 5475.9 5825.5 5877.2 6003.0 6046.9 6273.6 6365.4

Table 3 Comparison of calculated results using orthogonal coordinates (Table 1) and bond-length coordinates (Table 2) with selected observable values (deperturbed for stretchbend Fermi resonance) for CH3F. All values are in cm1 Assignment n1(A1) n3(A1) n4(E) 2n3(A1) 3n3(A1) 2n42 (E) x33a x33b
a

Observed 2919.57 1048.6164 2998.971 2081.3864 3098.4464 6001.865 7.9264 7.9064


1

Orthogonal calc. 2913.90 1059.31 2879.32 2105.48 3138.71 5671.08 6.57 6.54
b

Bond length calc. 2941.43 1124.44 3032.09 2234.14 3329.21 5877.25 7.37 7.35 This value for x33

This value for x33 is calculated using n3 and 2n3. is calculated using n3 and 3n3.

A1 symmetry calculations used N = 21 (giving a Hamiltonian matrix of order 2678), N = 17 was used for the E levels (matrix order 1974) and N = 15 for the A2 levels (matrix order 441). For the A2 and E symmetry types, convergence similar to the A1 case is achieved with a smaller basis set as a result of there being a lower density of states than in the A1 case. The complete calculation took 14.4 h on a Compaq Alpha XP1000/667 workstation. Results are given in Table 2 for a calculation carried out using the conventional bond length, bond angle approach. This has been achieved by making slight modications to the stretching variational procedure already described. The major change is the use of a stretching KEO derived in terms of bond length radial vectors. This KEO may be readily obtained because, as detailed in Part I,41 our approach to deriving the operator was based on the use of a general set of radial vectors. The only change to the KEO from that of eqn (11) is the addition of terms involving second order derivative operators coupling the radial vectors to one other.53 This change to the KEO and the use of a bond length V matrix (see eqn (8) of Part I41) are the only alterations that need be made to the computer code. This allows the stretching ab initio PES to be tested in a variational calculation with the same coordinates. It also gives an estimate of the error introduced in our orthogonal coordinate calculations by using a PES tted in a dierent coordinate system. Fewer levels are presented in Table 2 than for the orthogonal coordinate calculation because the main purpose of the calculation is to obtain an estimate of the stretchbend coupling, and this may be achieved by comparison of only a few energy levels. All the energy levels presented in Table 2 are converged to 0.1 cm1 or better, the calculation is performed using a N = 11 basis for the A1 and the E symmetry blocks. Note that the N = 11 basis is sucient to obtain the smaller number of levels in Table 2 to the same degree of convergence (better than 0.1 cm1) as those presented in Table 1. We compare the results calculated with orthogonal and bond-length coordinates with a small number of observables in Table 3. The importance of stretchbend coupling in molecules such as CH3F means that we would not expect the calculated and observed results to be in particularly good agreement. Similar observations have also been made in stretching-only calculations of methane.27 Additionally, for the stretching-only calculation, the calculated results obtained using the dierent coordinates systems would not be expected to agree.
This journal is
 c

Nevertheless, the results in Table 3 demonstrate that the stretching variational codes and the CH3F ab initio stretching PES produce physically realistic vibrational energy levels. The pattern of levels and their symmetries are correct, while the reproduction of the x33 anharmonicity constant is encouraging. In order to investigate the anharmonic CH stretching part of the potential, we have retted our 1D CH ab initio points to a Taylor expansion and used the resultant force constants to obtain a value for the local mode CH anharmonicity, xCH (using second order perturbation theory56). The value of the anharmonicity constant obtained is 62.5 cm1. This compares well with the experimental value of Law6: 61.0 cm1. In addition, we have carried out some large A1 symmetry calculations in order to compare calculated and observed high energy CH overtone transitions. The calculated values are obtained using the bond length variational code described above. The calculation uses a basis of NCH = 24 in the CH modes, with NCF = 21 for the CF mode (the total number of quanta in all 4 modes is r24), this results in a matrix of order 4218. Using this basis, it has been possible to converge the results presented in Table 4 to 1 cm1 or better. The results in the Table demonstrate that for 36 quanta of excitation, the calculated vibrational energy levels obtained are consistently too high. For the higher observed states in CH3F, the motions are normally considered as isolated CH stretching vibrations, nearly free of the eects of stretchbend Fermi resonance coupling. The overestimation of the vibrational energies together with the value for the anharmonicity constant given above would suggest that stretchbend coupling is having a signicant eect on

Table 4 Comparison of selected CH stretching results (calculated using the bond-length coordinates and NCH = 24, NCF = 21 basis) with corresponding observable values for CH3F; for each manifold with n quanta of total CH stretching excitation, the lowest A1 state is given. The n = 1 observed value is deperturbed for stretchbend Fermi resonances. All values are in cm1 No. CH quanta (n) 1 2 3 4 5 6 Obs. 2920 58006 85356 11 1356 13 6176 15 97266
1

Calc. 2941 5826 8620 11 288 13 834 16 266

(Obs. Calc.)/[n(n + 1)] 10.5 4.3 7.1 7.6 7.2 7.0

the Owner Societies 2006

Phys. Chem. Chem. Phys., 2006, 8, 28552865 | 2861

the anharmonicity of the CH stretching energy levels. Considering the higher CH stretching vibrations as isolated Morse oscillators, the eective CH stretching anharmonicity constant contributes a term n (n + 1)xCH to these observed transitions. Hence, column 4 of Table 4 shows the eective anharmonicity is consistently too small in magnitude for the present calculations. For the orthogonal coordinate calculation, it can be seen from Table 3 that certain levels agree closely with the observed values. However, the inconsistency in the use of coordinates for this calculation may result in a fortunate cancellation of errors. As with the bond length calculation, and as expected, the discrepancies between the calculated and observed levels become larger for the overtone levels. For the orthogonal calculation, the error introduced by the inconsistency in the use of coordinates makes further comparison of higher energy overtone levels redundant. 3.3. Bending

The variational calculations of the bending energy levels of CH3F use the vibrational KEO derived in Part I41 and the bending ab initio PES described in Section 2.4. Again our method is general for any similar XY3Z system. The bending vibrational problem is more challenging than the stretching one for several reasons. Firstly, there are ve bending degrees of freedom compared to just four in the stretching case. Thus, the density of states is considerably higher and the multidimensional quadrature used to evaluate potential energy matrix elements is much more demanding. Also, depending upon the choice of angular coordinates, calculating the bending energy levels of a centrally-connected penta-atomic system may be further complicated by the presence of the angular redundancy. This is seen in the work of Xie and Tennyson28 and Mladenovic57 but our choice of coordinates makes the redundancy relatively straightforward to deal with. 3.3.1. Hamiltonian and basis functions. For the bendingonly calculation, the radial coordinates are held at their equilibrium values. The bending KEO is constructed using eqn (2) of Part I41 and by selecting all gij and hi terms, in coordinates b and f only, from Table 1 of Part I.41 This gives a KEO consisting of twenty separate terms. Unlike the stretching case, no part of the Jacobian is incorporated into the operator or basis functions. The bending Hamiltonian is therefore integrated over the following volume element and integration limits: Z p Z p Z p Z 2p Z 2p sin b1 db1 sin b2 db2 sin b3 db3 df2 df3
0 0 0 0 0

However, there are technical disadvantages associated with the use of a coupled angular basis.51 For example, if a basis set contraction approach is being used then a very large number of eigenfunctions must be stored at each contraction stage. Also, for XY3Z-type molecules, such as the methyl halides, even at very high energies the singular points (b = 0, b = p) will never be probed. Consequently, we can make use of the direct product basis described above. Additionally, this leaves open the possibility of a straightforward transformation to a DVR representation, since the FBR approach will undoubtedly prove too computationally-demanding when applied to the full 9D stretchbend vibrational problem. Note that, although the singular points lie far from the equilibrium geometry, the nature of the Legendre functions means a quadrature based only on points far from either of the integration limits (b = p, b = 0, the singular points) is unlikely to be accurate. The diculties associated with the primitive Legendre functions are overcome by carrying out a basis set contraction in each b coordinate. It is of course only necessary to perform this for one b mode if all three are related by symmetry. Using the primitive Legendre basis functions, we diagonalise a 1D matrix in b, leaving all other coordinates xed at their equilibrium values. The coecients of the eigenvectors are then used to form the 1D contracted b functions. For example, in b1 Fa1 b1
NP1 X j1 1

Pj1 cos b1 Cj1 a1 :

20

19 The basis functions were chosen to be Legendre functions,58 Pj (cos b), to describe the b motion and sin (kf) and cos (kf) functions for the f motion. However, inspection of Table 1 of Part 141 and eqn (19) shows that with this choice of basis the cot2 and csc2 terms in the KEO give rise to singularities with innite integrals. The most rigorous solution to this problem is to use instead a coupled angular basis (associated Legendre functions).51 This results in a cancellation of these singular terms.
2862 | Phys. Chem. Chem. Phys., 2006, 8, 28552865

where NP1 is the number of primitive b functions, a1 runs from 1 to the number of contracted functions used and Cj1a1 is an eigenvector coecient. As b tends to zero (corresponding to the overlap of two vectors) the potential becomes highly repulsive while as b tends to p (corresponding to a linear XCH geometry in the methyl halides) the potential becomes highly attractive. Consequently, the new optimised functions rapidly approach zero at these points and the integrals hFai|cot2bi|Faii, hFai|csc2bi|Faii may therefore be evaluated accurately. The newly optimised functions are also more physically suitable than the primitive Legendre functions because they more closely resemble the true eigenfunctions of the system. This reduces the size of the matrix required to converge the full 5D bending calculation. A 1D f contraction is also performed in order to obtain a set of f functions that are more physically realistic than the primitive sin (kf) and cos (kf) functions. The 1D f contraction is carried out with b = p/2 for the KE contribution but at b = be for the potential integrals. This is necessary at this stage for two reasons. Firstly, because of the singular terms in the KEO and secondly, because isolated terms in the KEO are non-Hermitian in our chosen basis. This approximation will not aect the accuracy of our nal result. Again, it is only necessary to perform the contraction once as the f modes are related by symmetry. The contracted f functions, for example in f2, are given by Fb2 f2
1 X k2 MOP2

sink2 f2 Ck2 b2 ;

MOP2 X k2 0

cosk2 f2 Ck2 b2 ; 21

This journal is

 c

the Owner Societies 2006

where MOP2 is the maximum order of primitive f functions, b2 runs from 1 to the number of contracted 1D f functions used and Ck2b2 is an eigenvector coecient. 3.3.2. Symmetry and evaluation of matrix elements. Symmetrisation of the bending basis is more complicated than for the stretching case discussed above. There are now two types of coordinate and implications of the angular redundancy to be considered. Due to the redundancy, we do not use the coordinate f1. This means it is not possible, in our chosen basis, to carry out a direct symmetrisation of the contracted 1D f functions in an analogous manner to the stretching case. Instead it is necessary, as shown by Handy et al. in the case of ammonia,59 to exploit the correlation between the C3v group and its Cs subgroup. The procedure involves carrying out a further 2D f basis set contraction using the 1D f contracted functions already obtained. The 2D f contraction is carried out in Cs symmetry, producing A 0 and A00 sets of eigenvalues and eigenfunctions. It is noted that, as in the 1D contraction step, an approximate form of the KEO must be used (that is, with bi = p/2) but again this will have no eect on the accuracy of the nal answer. Inspection of the C3v/Cs correlation table60 shows how the individual symmetry contracted functions may be isolated. The eigenvalues and eigenvectors of the rst contraction (A 0 ) are stored, then after the second contraction (A00 ) the two sets of eigenvalues are compared. States that agree to within a degeneracy threshold of 104 cm1 are identied as E type contracted functions. The remaining A 0 and A00 contracted functions may then be immediately identied as A1 and A2 functions, respectively. The full symmetrised 5D basis is then constructed by combining these 2D f functions with symmetrised 3D b functions (which are symmetrised in exactly the same way as the three symmetrically equivalent stretching modes above) using the appropriate vector coupling coecients:61 A1 :A1 bA1 f A2 bA2 f 1 p Ea bEa f Eb bEb f 2 A2 :A1 bA1 f A2 bA2 f 1 p Ea bEb f Eb bEa f 2 Ea :A1 bEa f A2 bEb f Ea bA1 f Eb bA2 f 1 p Ea bEa f Eb bEb f 2
This journal is
 c

Eb :A1 bEb f A2 bEa f Ea bA2 f Eb bA1 f 1 p Ea bEb f Eb bEa f: 2 All the matrix elements for the full 5D bending calculation are evaluated using these symmetrised 5D basis functions. For the KE matrix elements, integration over the f modes is carried out analytically,62 while for the b modes it is performed using an appropriate Gaussian quadrature scheme.55 The factorisability of the KEO means these numerical integrals are inexpensive and can therefore be rapidly calculated to very high accuracy. At each stage of contraction, the KE integrals are stored for reuse in the full 5D calculation. The integration of the potential matrix elements is carried out numerically in ve dimensions because of our assumption of an arbitrary form for the bending PES. For the b modes, we use a GaussLegendre quadrature scheme. Integration over the f modes is performed using a trapezoid rule with quadrature points and weights   2p 1 2p fk k ; wk ; nf 2 nf 22

where nf is the number of quadrature points and k = 1,. . .,nf. The 5D quadrature formula is given by ha1 a2 a3 c23 ; GjVja01 a02 a03 c023 ; Gi
p p p q q XXXXX i j k l m

wbi ; bj ; bk ; fl ; fm cG1 a2 a3 c23 bi ; bj ; bk ; fl ; fm a cG1 a2 a3 c23 bi ; bj ; bk ; fl ; fm a Vbi ; bj ; bk ; fl ; fm : 23 Where (bi,bj,bk,fl,fm) denotes a point on the 5D quadrature grid, w(bi,bj,bk,fl,fm) the corresponding product of weights and cG a symmetrised wavefunction comprised of products of 1D b contracted functions (labelled by ai) and a 2D f contracted function (labelled by c23). Calculating the potential energy matrix elements in this manner is extremely computationally-demanding. Unfortunately, with the present choice of angular coordinates and quadrature points for the f modes, it is not possible to exploit the full C3v symmetry molecular symmetry of an XY3Z system. This is because not every point of our quadrature grid is mapped onto the grid by the symmetry operations of the C3v group. However, this can be remedied63 within the framework of our general approach. In the present work we have only implemented a symmetry saving for the bending potential integrals using the Cs group. This does not complicate the
Phys. Chem. Chem. Phys., 2006, 8, 28552865 | 2863

the Owner Societies 2006

expression given in eqn (23), as happened in the stretching case, because the components of the degenerate representations are not mixed. Rather, eqn (23) is only slightly modied such that the nal summation runs to l rather than q. The bending basis set is characterised using the notation: {b,t,p}, where b gives the number of contracted b functions in each mode, the constraint Si3bi r t is applied and p gives the number of contracted 2D f functions. 3.3.3. Results. Again, the bending-only vibrational calculations use coordinates based on the orthogonal vectors discussed above. The bending potential is, however, calculated and tted in terms of coordinates based on bond length vectors. As above, a problem arises in that it is not possible in the present work to convert between the two coordinate systems, because this requires use of a full 9D PES. Again, at least for low lying energy levels, the error introduced by this inconsistency is comparable with the dierence that will arise between the calculated and observed energy levels as a result of the lack of stretchbend coupling. Table 5 contains CH3F bending energy levels of all symmetries up to just above 5000 cm1 above the ground state. All levels given are converged to 0.3 cm1 or better. For A1 and A2 symmetries, the calculation is carried out using a {9,12,21} basis, producing an A1 matrix of order 803, and an A2 matrix of order 716. The E symmetry calculation is carried out using a {9,11,19} basis, producing a matrix of order 1036. The complete calculation took three weeks on a Compaq Alpha XP1000/667 workstation. The majority of this time was spent performing the numerical integration of the potential energy matrix elements. 3.3.4. Discussion. The importance of stretchbend coupling in molecules such as CH3F means that we would not
Table 5 Calculated vibrational bending energies for CH3F (in cm1). Excited state energies are given relative to the ground state Symmetry A1(ground state) E A1 E E A1 E A1 E A1 A2 E E A1 E A1 A2 E A1 E E E A1 E A1 Energy 3331.0 1173.6 1331.2 1543.4 2320.4 2328.2 2517.6 2673.8 2683.5 2697.7 2757.0 2885.7 3056.1 3067.0 3484.9 3598.1 3608.5 3675.6 3683.5 3832.9 3894.1 4009.8 4025.8 4037.3 4085.0 Symmetry A2 E A1 A2 A1 A2 E A1 A2 E E A1 E E A1 E E E A1 A2 A1 E E A1 E Energy 4111.6 4167.0 4168.6 4168.6 4182.2 4183.4 4219.0 4241.4 4244.3 4259.0 4408.1 4420.5 4499.5 4581.7 4659.5 4806.7 4852.7 4967.9 4971.0 4983.9 5003.4 5042.9 5065.1 5070.6 5196.9

Table 6 Comparison of calculated bending vibrational energy levels with selected observable values (deperturbed for stretchbend Fermi resonance) for CH3F. All values are in cm1 Assignment n2(A1) n5(E) n6(E) 2n2(A1) 2n50 (A1) n2 + n5(E) 2n52 (E) Observed 1459.4 1467.8 1182.7 2914.2 2921.7 2923.7 2932.7
1 1 67 1 1 1 1

Calculated 1331.2 1543.4 1173.6 2697.7 3067.0 2885.7 3056.1

expect the calculated and observed results to be in particularly good agreement. Similar observations having also been made in a recent bending-only calculation of methane.29 Nevertheless, the variational code developed in this work produces the correct pattern of levels with the correct symmetries. The results in Table 6 also demonstrate that the bending variational code and the CH3F ab initio bending surface produce physically realistic vibrational energy levels. The use of contracted basis functions as a way of overcoming singularities in the KEO means that the bending code could not be used to calculate all bound bending levels of a XY3Z system. However, the very high energy of the singular points means that the code may be successfully used to calculate a very large number of high energy bending vibrational levels.

4. Conclusion
We have calculated ab initio 5D bending and 4D stretching potential energy surfaces for CH3F as an important rst stage towards a full 9D potential energy surface for this molecule. Using these surfaces, we have calculated variationally highly excited stretching and (separately) bending vibrational energy levels of CH3F using a nite basis representation method. The method uses the exact vibrational kinetic energy operator derived in Part I41 and the full C3v symmetry is used to form the nal 4D and 5D basis functions. The computer codes are designed to use an arbitrary potential energy function. Ultimately, these results will be used to form a compact basis for fully coupled stretchbend calculations of the vibrational energy levels of the CH3F system. If we retain the assumption of an arbitrary form for the potential energy surface then the cost of numerical quadrature will become prohibitive for full 9D vibrational energy level calculations. Recent full-dimensional variational calculations for methane have made use of the discrete variable representation (DVR) for some or all of the vibrational modes.31,32 In the DVR basis, the potential energy matrix is diagonal so that multidimensional quadrature over the internal coordinates is not required. We have already computed accurate stretching vibrational energy levels of CH3F using the DVR and potential-optimised DVR approaches (obtaining good agreements with the FBR results reported above) and are currently developing computer codes to tackle the full 9D vibrational problem. This work will also include rotational motion. Further consideration will be given to the choice of internal coordinates. The latter will determine the degree of separation of vibrational and rotational motions.33 The maximum
This journal is
 c

2864 | Phys. Chem. Chem. Phys., 2006, 8, 28552865

the Owner Societies 2006

separation is critical to successful solution of the full rovibrational Schrodinger equation. The overall approach taken here and in Part I41 will allow considerable freedom in the ultimate choice of internal coordinates.

Acknowledgements
We thank the UK Engineering and Physical Sciences Research Council (EPSRC) for access to the CSAR facility via the ChemReact Consortium. We thank the Carnegie Trust for the Universities of Scotland for supporting this work via the awards of studentships to Steven Manson and Ian Atkinson. We also thank a number of people for helpful discussions, including Jonathan Tennyson, Junkai Xie, Mirjana Mladeno vic and Igor Kozin. Finally, we thank the referees for comments on the manuscript.

References
1 M. Badaoui and J. P. Champion, J. Mol. Spectrosc., 1985, 109, 402. 2 K. M. Dunn, J. E. Boggs and P. Pulay, J. Chem. Phys., 1987, 86, 5088. 3 D. Luckhaus and M. Quack, Chem. Phys. Lett., 1992, 190, 581. 4 T. K. Ha, D. Luckhaus and M. Quack, Chem. Phys. Lett., 1992, 190, 590. 5 J. Lummila, L. Halonen, I. Merke and J. Demaison, J. Mol. Spectrosc., 1996, 179, 125. 6 M. M. Law, J. Chem. Phys., 1999, 111, 10021. 7 D. Papousek, P. Pracna, M. Winnewisser, S. Klee and J. Demaison, J. Mol. Spectrosc., 1999, 196, 319. 8 I. A. Atkinson and M. M. Law, J. Mol. Spectrosc., 2001, 206, 135. 9 A. Baldacci, R. Visinoni and G. Nivellini, Mol. Phys., 2002, 100, 3577. 10 A. Baldacci, R. Visinoni and G. Nivellini, Mol. Phys., 2004, 102, 1731. 11 S. Kondo, Y. Koga and T. Nakanga, J. Chem. Phys., 1984, 81, 1951. 12 M. M. Law, J. L. Duncan and I. M. Mills, J. Mol. Struct., 1992, 260, 323. 13 W. Schneider and W. Thiel, Chem. Phys., 1992, 159, 49. 14 J. Demaison, J. Breidung, W. Thiel and D. Papous ek, Struct. Chem., 1999, 10, 129. 15 I. A. Atkinson and M. M. Law, Spectrochim. Acta, Part A, 2002, 58, 873. 16 G. F. Bauerfeldt and H. Lischka, J. Phys. Chem. A, 2004, 108, 3111. 17 L. Wang, V. V. Kislov, A. M. Mebel, X. Yang and X. Wang, Chem. Phys. Lett., 2005, 406, 60. 18 D. Luckhaus, M. Quack and J. Stohner, Chem. Phys. Lett., 1993, 212, 434. 19 M. M. Law and J. L. Duncan, Chem. Phys. Lett., 1993, 212, 172. 20 R. P. A. Bettens, J. Am. Chem. Soc., 2003, 125, 584. 21 P. Botschwina, M. Horn, S. Seeger and R. Oswald, Ber. BunsenGes. Phys. Chem., 1997, 101, 387. 22 L. Sun, K. Song, W. L. Hase, M. Sena and J. M. Riveros, Int. J. Mass Spectrom., 2003, 227, 315. 23 J. M. Gonzales, W. D. Allen and H. F. Schaefer, J. Phys. Chem. A, 2005, 109, 10613. 24 D. W. Schwenke and H. Partridge, Spectrochim. Acta, Part A, 2001, 57, 887. 25 D. W. Schwenke, Spectrochim. Acta, Part A, 2002, 58, 849. 26 R. Marquardt and M. Quack, J. Phys. Chem. A, 2004, 108, 3166. 27 J. Xie and J. Tennyson, Mol. Phys., 2002, 100, 1615. 28 J. Xie and J. Tennyson, Mol. Phys., 2002, 100, 1623. 29 X.-G. Wang and T. Carrington, Jr, J. Chem. Phys., 2003, 118, 6946.

30 X.-G. Wang and T. Carrington, Jr, J. Chem. Phys., 2003, 119, 94. 31 X.-G. Wang and T. Carrington, Jr, J. Chem. Phys., 2004, 121, 2937. 32 H.-G. Yu, J. Chem. Phys., 2004, 121, 6334. 33 J. Tennyson, Variational Calculations of Rotation-Vibration Spectra, in Computational Molecular Spectroscopy, ed. P. Jensen and P. R. Bunker, Wiley, Chichester, 2000, pp. 305323. 34 S. Dressler and W. Thiel, Chem. Phys. Lett., 1997, 273, 71. 35 S. Carter, I. M. Mills and N. C. Handy, J. Chem. Phys., 1993, 99, 4379. 36 S. Carter and W. Meyer, J. Chem. Phys., 1994, 100, 2104. 37 S. M. Colwell, S. Carter and N. C. Handy, Mol. Phys., 2003, 101, 523. 38 S. Zou, J. M. Bowman and A. Brown, J. Chem. Phys., 2003, 118, 10012. 39 I. N. Kozin, M. M. Law, J. Tennyson and J. M. Hutson, Comput. Phys. Commun., 2004, 163, 117. 40 I. N. Kozin, M. M. Law, J. Tennyson and J. M. Hutson, J. Chem. Phys., 2005, 122, 064309. 41 S. A. Manson and M. M. Law, Phys. Chem. Chem. Phys.DOI: 10.1039/b603106d. 42 R. D. Amos, A. Bernhardsson, A. Berning, P. Celani, D. L. Cooper, M. J. O. Deegan, A. J. Dobbyn, F. Eckert, C. Hampel, G. Hetzer, P. J. Knowles, T. Korona, R. Lindh, A. W. Lloyd, S. J. McNicholas, F. R. Manby, W. Meyer, M. E. Mura, A. Nicklass, P. Palmieri, R. Pitzer, G. Rauhut, M. Schutz, U. Schumann, H. Stoll, A. J. Stone, R. Tarroni, T. Thorsteinsson and H.-J. Werner, MOLPRO, a package of ab initio programs designed by H.-J. Werner and P. J. Knowles, Version 2002.1, 2002. 43 G. Tarczay, A. G. Csaszar, O. L. Polyansky and J. Tennyson, J. Chem. Phys., 2001, 115, 1229. 44 O. L. Polyansky, A. G. Csaszar, S. V. Shirin, N. F. Zobov, P. Barletta, J. Tennyson, D. W. Schwenke and P. J. Knowles, Science, 2003, 299, 539. 45 M. M. Law and J. M. Hutson, Comput. Phys. Commun., 1997, 102, 252. 46 R. Jaquet, Lect. Notes Chem., 1999, 71, 97. 47 C. Leeonard, N. C. Handy, S. Carter and J. M. Bowmam, Spectrochim. Acta, Part A, 2002, 58, 825. 48 D. W. Schwenke, J. Phys. Chem., 1996, 100, 2867. 49 B. T. Sutclie and J. Tennyson, Mol. Phys., 1986, 58, 1053. 50 N. C. Handy, Mol. Phys., 1987, 61, 207. 51 M. J. Bramley, W. H. Green and N. C. Handy, Mol. Phys., 1991, 73, 1183. 52 T. J. Lukka, J. Chem. Phys., 1995, 102, 3945. 53 J. Tennyson and B. T. Sutclie, J. Chem. Phys., 1982, 77, 4061. 54 L. Halonen and M. S. Child, J. Chem. Phys., 1983, 79, 4355. 55 A. H. Stroud and D. Secrest, Gaussian Quadrature Formulas, Prentice-Hall, Englewood Clis, 1966. 56 I. M. Mills, in MolecularSpectroscopy: Modern Research, ed. K. Narahari Rao and C. W. Mathews, Academic Press, New York, 1972. 57 M. Mladenovic, J. Chem. Phys., 2003, 119, 11513. 58 G. B. Arfken and H. J. Weber, Mathematical Methods for Physicists, 5th edn, Academic Press, 2001. 59 N. C. Handy, S. Carter and S. M. Colwell, Mol. Phys., 1999, 96, 477. 60 E. B. Wilson, J. C. Decius and P. C. Cross, Molecular Vibrations, McGraw-Hill, New York, 1955. 61 J. S. Grith, The Theory of Transition Metal Ions, Cambridge University Press, Cambridge, 1961. 62 I. S. Gradshtein and I. M. Ryzhik, Tables of Integrals, Series and Products, 4th edn, Academic Press, New York, 1980. 63 X.-G. Wang and T. Carrington, Jr, J. Chem. Phys., 2004, 123, 154303. 64 G. Graner, J. Phys. Chem., 1979, 83, 1491. 65 E. W. Jones, R. J. L. Popplewell and H. W. Thompson, Proc. R. Soc. London, Ser. A, 1966, 290, 490. 66 J. S. Wong and C. B. Moore, J. Chem. Phys., 1982, 77, 603. 67 E. Hirota, J. Mol. Spectrosc., 1978, 70, 469.

This journal is

 c

the Owner Societies 2006

Phys. Chem. Chem. Phys., 2006, 8, 28552865 | 2865

Вам также может понравиться