You are on page 1of 8

2 Novemberc 2001

Chemical Physics Letters 348 (2001) 131±138


www.elsevier.com/locate/cplett

Hartree±Fock geometry optimisation of periodic systems


with the CR Y S T A L code
a,*
B. Civalleri , Ph. D'Arco b, R. Orlando c, V.R. Saunders d, R. Dovesi a,e

a
Dipartimento di Chimica IFM, Universit a di Torino, Via P. Giuria 7, 10125 Torino, Italy
b
Laboratoire de P etrologie, Mod elisation de Materiaux et Processus, Universit
e Pierre et Marie Curie,
4, Place Jussieu, 75232 Paris Cedex 05, France
c
Dipartimento di Scienze e Tecnologie Avanzate, Universit a del Piemont Orientale, C.so Borsalino 54, 15100 Alessandria, Italy
d
CLRC Daresbury Laboratory, Daresbury, WA4 4AD Warrington, UK
e
Unit
a INFM di Torino, Sezione F, Via P. Giuria 7, 10125 Torino, Italy
Received 9 August 2001; in ®nal form 6 September 2001

Abstract
Results are reported on the geometry optimisation of periodic systems with the Hartree±Fock analytical gradients
recently implemented in the CR Y S T A L code. Application to the structure optimisation of molecules, polymers, slabs
and crystals is presented. Ó 2001 Elsevier Science B.V. All rights reserved.

1. Introduction First-derivatives can be computed either nu-


merically or analytically. Nevertheless, analytical
The determination of equilibrium structure is of gradients have become an ever-increasing area in
primary importance in the modelling of chemical quantum chemistry and, when available, provide
systems and several methods for their geometry an important tool to facilitate the determination of
optimisation have been proposed [1]. All of them equilibrium structures.
try to locate minima and saddle points on the In the ®eld of quantum chemistry, the Hartree±
potential energy surface (PES). Basic tools to de- Fock approach still remains a fundamental theory
scribe the PES are ®rst- and second-derivatives. for both molecular and solid-state chemistry and is
First-derivatives permit to ®nd stationary points the basis for most of the wavefunction-based
on the PES whilst second-derivatives allow to methods. Analytical HF gradients are commonly
characterize them. Among geometry optimisation implemented in standard molecular codes and
techniques, gradient-based algorithms have been routinely used in molecular calculations, but this
indicated as the methods of choice for most levels does not apply for solid-state calculations al-
of theory [2]. Thus, gradients play a twofold cen- though, signi®cant work has been done for one-
tral role in the geometry optimisation of chemical dimensional periodic systems [3±5]. Recently, a
systems. general implementation of analytical Hartree±
Fock gradients has become available for zero, one,
two, and three-dimensional systems [6±8].
*
Corresponding author. Fax: +39-011-6707855. In this Letter we report on the geometry opti-
E-mail address: civalleri@ch.unito.it (B. Civalleri). misation of periodic systems based on these ana-
0009-2614/01/$ - see front matter Ó 2001 Elsevier Science B.V. All rights reserved.
PII: S 0 0 0 9 - 2 6 1 4 ( 0 1 ) 0 1 0 8 1 - 8
132 B. Civalleri et al. / Chemical Physics Letters 348 (2001) 131±138

lytical Hartree±Fock gradients [6±8] as imple- For periodic systems, a cell ®xed unconstrained
mented in the CR Y S T A L code [9]. Examples of optimisation of the atomic coordinates was carried
structure optimisation are reported ranging from out [11].
molecules to polymers, slabs and crystals. To the
best of our knowledge this is the ®rst time that
geometry optimisation of crystals adopting ana- 3. Results and discussion
lytical Hartree±Fock gradients is reported.
3.1. Preliminary calculations

2. Computational details Preliminarily, a detailed investigation has been


performed on the geometry optimisation of mo-
The calculation of the forces on the atoms has lecular structures. The main goal is to compare our
been performed by using the recent implementa- implementation with that of molecular ab initio
tion of analytical HF energy gradients [6±8] in the codes. We refer to the performance test on the
CR Y S T A L program [9]. These have been used to determination of equilibrium molecular structures
relax the atoms to equilibrium using a modi®ed presented by Baker [12], where the geometry op-
conjugate gradient algorithm as proposed by timisations in Cartesian coordinates and natural
Schlegel [10]. The optimisation algorithm evalu- internal coordinates of 30 molecules at the HF
ates the gradients each time the energy is com- level and by using a STO-3G basis set are com-
puted and the second derivative matrix is updated pared. In particular, in the present work, we use as
employing the gradients. At each step, a one- a reference the set of results obtained in [12] for the
dimensional minimisation using a quadratic geometry minimisation in cartesian coordinates
polynomial is carried out, followed by an with a unit Hessian as initial guess.
n-dimensional search using the Hessian matrix. The geometry optimisation has been carried out
In the present work, a unit matrix is used as an by adopting almost the same computational con-
initial guess for the Hessian and symmetry is fully ditions as Baker. Results are reported in Table 1.
exploited. Convergence is tested on the root- Initial and ®nal energies of the molecules di€er by
mean-square (RMS) and the absolute value of the less than 1  10 5 Hartree from Baker's determi-
largest component of both the gradients and the nations, showing that the optimised geometries are
estimated displacement of nuclei. In the present essentially the same in both cases. From the
calculations the threshold for the maximum and number of iterations one can conclude that e-
the RMS forces and the maximum and the RMS ciency is comparable for the two implementations.
atomic displacements on all the atoms have been Although a few critical cases are observed such as
set to (in a.u.) 0.00045, 0.00030 and 0.00180, benzidine and menthone, the global sum of the
0.00120, respectively. The optimisation is consid- number of iterations needed for this set of mole-
ered complete when all four conditions are sat- cules is quite similar: 750 in our case and 765 in
is®ed. Baker's. This small di€erence is probably due to
The choice of these conditions enables a tight slightly di€erent computational conditions and the
comparison with standard ab initio molecular di€erent optimisation algorithm adopted to per-
codes (vide infra) For the same reason the toler- form the geometry relaxation. Nevertheless, results
ances that rule the calculation of the Coulomb and in Table 1 clearly show that the present imple-
exchange integrals were set to TOL ˆ 15 15 15 15 mentation is reliable and of comparable eciency
20 in the molecular cases [9]. For periodic calcu- as standard molecular codes.
lations the default CR Y S T A L tolerances [9] were The application of this procedure to periodic
adopted, instead. The threshold for stopping the systems requires a careful preliminary analysis of
SCF cycle was 1  10 8 and 1  10 6 Hartree for the in¯uence of computational conditions, such as
molecular total energy and in the case of periodic the truncation of the Coulomb and exchange series
calculations, respectively. and the de®nition of the sampling net in the irre-
B. Civalleri et al. / Chemical Physics Letters 348 (2001) 131±138 133

Table 1
Comparison between present work (pw) and previous data on Baker's [12] set of molecules
Molecule Number of at- Symmetry Number of Number of cycles
oms variables pw Ref. [12]
Water 3 C2v 2 5 5
Ammonia 4 C3v 2 6 7
Ethane 8 D3d 3 7 7
Acetylene 4 D1h 2 6 7
Allene 7 D2d 3 7 10
Hydroxysulphane 4 C1 6 26 21
Benzene 12 D6h 2 5 6
Methylamine 7 Cs 10 8 10
Ethanol 9 Cs 13 19 18
Acetone 10 C2v 8 15 22
Disilyl ether 9 C2v 7 20 27
1,3,5-trisilacyclohexane 18 C3v 11 23 36
Benzaldehyde 14 Cs 25 14 17
1,3-di¯uorobenzene 12 C2v 11 7 8
1,3,5-tri¯uorobenzene 12 D3h 4 7 7
Neopentane 17 Td 3 6 10
Furan 9 C2v 8 12 10
Naphthalene 18 D2h 9 10 11
1,5-di¯uoronaphthalene 18 C2h 17 14 16
2-hydroxybicyclopentane 14 C1 36 44 45
ACHTAR10 16 C1 42 47 51
ACANIL01 19 Cs 34 35 36
Benzidine 26 D2 18 39 26
Pterin 17 Cs 31 20 23
Difuropyrazine 16 C2h 15 18 21
Mesityl oxide 17 Cs 28 38 38
Histidine 20 C1 54 92 102
Dimethylpentane 23 C1 63 31 29
Ca€eine 24 Cs 42 42 39
Menthone 29 C1 81 127 100

ducible Brillouin zone (IBZ) [9], on the accuracy of grals screening (TOL ˆ 6 6 6 6 12), [9] we have
geometry optimisation. progressively increased the number of sampling
On this purpose, we performed a systematic in- points (NKP) in IBZ (see cases 1, 3, 8, and 9) and
vestigation on the geometry optimisation of crys- tested the convergence on Ei . Results in Table 2
talline urea, as a test case. The experimental show that with 18 k-points (case 3) Ei is already
structure determined by neutron low temperature stable. As regards the dependence of Ei on TOL
di€raction data [13] has been used as initial ge- (see cases 2, 3, 4, 5, 6, and 7) one can see that Ei is
ometry. All the optimisations have been carried out still varying in the explored range of TOL, though
by employing a STO-3G basis set and a unit matrix by less than 1 mHartree. However, the number of
as initial Hessian. Results are shown in Table 2. iterations is de®nitely insensitive to the computa-
In Table 2, the number of optimisation cycles, tional conditions, being always about 15 cycles
the initial geometry total energy, Ei and the energy and it is worth noting that the gain in the com-
gain at convergence, DE, are reported as well a the puted total energy is almost constant in all cases,
largest (Max DX10 ) and (RMS DX10 ) di€erence indicating that the ®nal structure obtained in dif-
between the optimised cartesian coordinates for ferent conditions are all very similar. This is also
each case and case 10, considered as a reference. con®rmed when considering Max DX10 and RMS
Starting from the default tolerances for the inte- DX10 which are in general below 1  10 3 and
134 B. Civalleri et al. / Chemical Physics Letters 348 (2001) 131±138

Table 2
Dependence of the geometry optimisation of crystalline urea on the number of sampling points in the IBZ, NKP, and on the integrals
screening tolerances, TOL [9]
Case NKP TOL Number Ei a DEa Maxb RMSb CPU
of cycles DX10 DX10 timec
1 6 6 6 6 6 12d 15 442:069367 0.015228 0.001021 0.000435 179
2 18 5 5 5 5 10 15 442:069659 0.015213 0.001310 0.000512 110
3 6 6 6 6 12d 15 442:069584 0.015265 0.000257 0.000083 180
4 7 7 7 7 14 15 442:069929 0.015326 0.000222 0.000118 299
5 8 8 8 8 16 15 442:070365 0.015330 0.000468 0.000185 443
6 5 5 5 10 10 14 442:069901 0.015291 0.000701 0.000262 132
7 6 6 6 12 12 14 442:069759 0.015330 0.000254 0.000108 228
8 40 6 6 6 6 12d 15 442:069584 0.015265 0.000653 0.000287 181
9 75 6 6 6 6 12d 14 442:069584 0.015265 0.000649 0.000253 185
10 8 8 8 8 16 16 442:070365 0.015330 0.000000 0.000000 439
a
Ei : initial geometry total energy; DE: gain in total energy at convergence (atomic units).
b
Max DX10 and RMS DX10 : maximum and RMS di€erence between the optimised cartesian coordinates of the current case and case 10,
as the reference.
c
Average CPU time for optimisation cycle in seconds.
d
Default values in CR Y S T A L .

5  10 4 A, respectively. In particular, for case 3 in the geometry optimisation of periodic systems,
their values are 0.000257 and 0.000083 A,  well we de®ned an appropriate test suite of 19 systems.
below the accuracy of an experimental structure General information on the adopted set of test
determination. cases is presented in Table 3.
In summary, good equilibrium structures can be Systems are grouped on the basis of their peri-
obtained by adopting well-converged HF solutions odicity: polymers, slabs, and crystals. For a sys-
and default integrals screening tolerances, as in tematic assessment, periodic systems are selected
case 3. This also leads to a signi®cant saving in to cover a wide range of structural types and
computational time as shown in the last column of symmetries. They represent most of the di€erent
Table 2. Indeed, the cost of case 10 is about 2.5 kinds of chemical bonds that can be found in solid-
times larger than that of case 3. state chemistry. Thus, ionic, semi-ionic, covalent
As a further information on the cost of the and molecular periodic systems are included.
calculations, a few details on the CPU time (on an Furthermore, the examined systems play an im-
IBM RS/6000 43P) for case 3 follow. For a con- portant role in materials science and most of them
ventional SCF, i.e., storing all of the integrals on have been the subject of previous studies in our
disk the average CPU time for an optimisation laboratory [14±24]. In particular, examples have
step takes 180 s, i.e., 37 s for the SCF cycle and 143 been reported on: oxides and their surface relax-
s for the gradient evaluation. For a direct SCF ation (a-Al2 O3 and a-Cr2 O3 from bulk to a slab
calculation, the average CPU time for an optimi- model [14,15]), silicates and zeolites [17±20], ad-
sation step is 346 s where the SCF cycle and the sorption on surfaces (MgO(0 0 1)/CO [21]) and in
gradient computation take 203 and 143 s, respec- zeolites (H±CHA(3:1)/NH3 [20]), defective systems
tively (it must be noticed that the analytical gra- (carbon substitution in bulk silicon, C@Si [22])
dient evaluation is always performed by means of and molecular crystals (e.g., ice [23] and urea [24]).
a direct algorithm). Initial starting structures were taken from well-
de®ned sources either from experimental data or
3.2. Periodic test cases from previous quantum or molecular mechanics
calculations. For the sake of brevity, we do not
On the basis of the previous calculations and in report all the references and structural details here,
order to illustrate the performances of the method however, full information about the starting ge-
B. Civalleri et al. / Chemical Physics Letters 348 (2001) 131±138 135

Table 3
General information on the periodic systems adopted as test cases in the present work
Dim. System Typea Symmetryb Number Basis set
of atoms
One-dimensional In®nite water chain P P1 (1) 6 STO-3G
Polyglycine (C5 conformer) P P1m1 (2) 14 STO-3G
Polytetra¯uoroethylene P Pmcm (8) 6 STO-3G
Two-dimensional a-Al2 O3 slab (6 atomic layers) S P3 (6) 10 STO-3G
a-Cr2 O3 slab (6 atomic layers) S P3 (6) 10 Cr,O [14,15]
MgO(0 0 1)/CO (3 atomic layers) Ad P4=mmm (16) 10 Mg,O [21]
Three-dimensional a-Al2 O3 IC R3c (12) 10 STO-3G
a-Cr2 O3 IC R3c (12) 10 Cr,O [14,15]
a-Quartz IC P32 21 (6) 9 STO-3G
a-Boron CC R3m (12) 12 STO-3G
C@Si (32 atoms super-cell) CC I43m (24) 32 STO-3G
Ferroelectric ordered ice (Ice XI) MC Cmc21 (4) 12 STO-3G
a-Oxalic acid dihydrate crystal MC P21 =n (4) 28 STO-3G
Urea crystal MC P421 m (8) 16 STO-3G
Silica sodalite SOD(Si) Z P43m (24) 36 STO-3G
Silica faujasite FAU(Si) Z Fd3m (48) 144 STO-3G
Titano chabazite Ti±CHA(1:1)c Z R3 (6) 36 Si,O,Ti [18]
Acidic chabazite H±CHA(3:1)d Z R3 (3) 39 STO-3G
H±CHA(3:1)/NH3 d Ad R3 (3) 51 STO-3G
a
Type: P ˆ polymer, S ˆ surface, Ad ˆ adsorption, IC ˆ ionic/semi-ionic crystal, CC ˆ covalent crystal, MC ˆ molecular crystal,
Z ˆ zeolite.
b
Number of symmetry operations in parentheses.
c
Si:Al ˆ 1:1
d
Si:Al ˆ 3:1

ometries can be found at the CR Y S T A L web site the studied systems are formed by three-dimen-
[25]. All optimisations were carried out employing sional networks of H-bonds. For instance, in the
the STO-3G basis set, except for a few cases ex- a-oxalic acid dihydrate crystal, each molecule is
plicitly indicated in Table 3, and a unit matrix has surrounded by six water molecules which are then
been used as initial guess for the Hessian. involved in three hydrogen bonds with nearby
Results are shown in Table 4. For each system, oxalic acid molecules (Fig. 1). The crystal structure
the number of variables to be optimised and the results in a low symmetry complex network of
number of optimisation cycles are reported as well H-bonds. In spite of such a complicated structure,
as energetic data, namely: initial and ®nal energies, the geometry optimisation of the oxalic acid crystal
and the gain in total energy at convergence. (21 degrees of freedom) is achieved after 49 cycles
Results in Table 4 show that geometry optimi- only, i.e., about twice the number of variables. In
sation performs equally well for systems with any general, results show that, even for weakly bound
periodicity: polymers, slabs and crystals. The systems, optimisation performs quite well.
quality of geometry optimisation appears to be Zeolites are another important class of materi-
independent of the type of structure and symmetry als. Because of their relevance in materials science,
and, as a general trend, the number of iterations a few cases have been included in the test suite.
ranges from one to ®ve times the number of vari- Among them, the geometry optimisation of H±
ables to be optimised. CHA(3:1)/NH3 (Fig. 2) is worth being remarked
Indeed, molecular crystals are important test upon. In fact, among the test cases here consid-
cases for geometry optimisation due to the pres- ered, H±CHA(3:1)/NH3 is the one with the largest
ence of soft modes and weak interactions. All of number of degrees of freedom and the speci®c Si/
136 B. Civalleri et al. / Chemical Physics Letters 348 (2001) 131±138

Table 4
Convergence data on the test suite of periodic systemsa
System Number of Number of cycles Ei Ef DE
variables
In®nite water chain 15 15 149:945729 149:960431 0.014702
Polyglycine 24 19 408:218296 408:268540 0.050243
Polytetra¯uoroethylene 3 6 466:968596 466:970283 0.001687
a-Al2 O3 slab 5 10 1399:799902 1400:114819 0.314917
a-Cr2 O3 slab 5 18 4622:595509 4622:823020 0.227511
MgO(0 0 1)/CO 4 12 1046:361591 1046:393148 0.031557
a-Al2 O3 2 4 1400:392826 1400:395110 0.002284
a-Cr2 O3 2 9 4623:109982 4623:110039 0.000057
a-Quartz 4 10 1300:382505 1300:383282 0.000777
a-Boron 4 5 292:609911 292:615488 0.005577
C@Si (SC32) 5 10 8892:597943 8892:655827 0.057884
Ice XI 10 15 299:957883 299:965335 0.007452
a-Oxalic acid 21 49 1042:387770 1042:637879 0.250109
Urea 8 15 442:069367 442:084595 0.015228
SOD(Si) 3 13 5201:458862 5201:472245 0.013383
FAU(Si) 10 23 20805:841975 20805:857697 0.015722
Ti±CHA(1:1) 18 39 8620:992411 8620:993034 0.000623
H±CHA(3:1) 38 70 5063:591308 5063:616329 0.025021
H±CHA(3:1)/NH3 50 169 5229:832874 5230:147140 0.314266
a
Ei and Ef : initial and ®nal total energies; DE: gain in total energy at convergence. For the initial geometry see [25] (atomic units).

Fig. 1. a Oxalic acid dihydrated structure.

Al loading of the framework considered allows the


interaction of three molecules of ammonia with the
acidic Brùnsted sites per unit cell of the zeolite.
Hence, geometry optimisation is inherently di-
Fig. 2. Structure of H±CHA(3:1) acidic zeolite supercage
cult due to structural complexity and presence of showing the interaction of the Brùnsted sites with NH3 . Atoms
hard and soft mode coupling. Nevertheless, the are displayed.
B. Civalleri et al. / Chemical Physics Letters 348 (2001) 131±138 137

number of iterations required (169 cycles) appears 4. Conclusions


as reasonable.
It has been shown [12,29] that eciency in ge- Application to the structure optimisation of
ometry optimisation can be improved by adopting periodic systems at the Hartree±Fock level has
a good initial guess for the Hessian and an ap- been reported. We have shown that geometry
propriate coordinate system. On the one hand, minimisations are now feasible for molecules,
Baker [12] addressed that a reduction by a factor polymers, slabs, and crystals. Further, the imple-
up to six can be obtained for geometry optimisa- mentation of an automated geometry optimisation
tion in cartesian coordinates provided that a reli- in a standard periodic ab initio code like CR Y S T A L
able estimated Hessian is available (e.g., from [9] provides an important tool for solid-state
molecular mechanics calculations or model Hes- computational chemistry, as it has been demon-
sians [26±28]). On the other hand, the choice of a strated for systems of interest in many ®elds of
proper coordinate system, like natural internal materials science.
coordinates, has been shown to be very useful [29]
to deal with dicult structure optimisations. In-
ternal coordinates reduce the coupling between Acknowledgements
hard and soft modes facilitating the location of
energy minima and could be very advantageous The present work is part of a project coordi-
for complex systems like H±CHA(3:1)/NH3 . In nated by A. Zecchina and co-®nanced by the
this respect, geometry optimisation in either re- Italian MURST (Co®n98, A03). The authors are
dundant internal coordinates [30] or delocalised grateful to P. Ugliengo, C. Roetti and C.M. Zi-
internal coordinates [31], has been recently pro- covich-Wilson for suggestions and helpful discus-
posed for periodic systems. Implementation of sions.
such techniques is under study in our laboratory.
The use of cartesian coordinates, however, al-
lows partial geometry optimisation to be readily References
carried out. Thus, geometry optimisation can be
limited to an atomic fragment instead of the whole [1] H.B. Schlegel, Adv. Chem. Phys. 687 (1987) 249.
[2] H.B. Schlegel, Geometry optimisation: 1, in: Encyclopedia
system. Accordingly, the symmetrised cartesian
of Computational Chemistry, Wiley, New York, 1998.
coordinates are then generated on the basis of a list [3] H. Terame, T. Yamabe, A. Imamura, J. Chem. Phys. 81
of atoms to be optimised. This procedure leads in (1984) 3564.
general to a reduction in the number of cycles and [4] J.-Q. Sun, R.J. Bartlett, J. Chem. Phys. 104 (1996) 8553.
is useful in the case of large systems where a small [5] D. Jacquemin, J.-M. Andre, B. Champagne, J. Chem.
Phys. 111 (1999) 5324.
atomic region, only, plays the major role in de-
[6] K. Doll, N.M. Harrison, V.R. Saunders, Int. J. Quantum
termining the chemical activity (e.g., acidic zeo- Chem. 82 (2001) 1.
lites). As a test case, a fragment geometry [7] K. Doll, Comp. Phys. Comm. 137 (2001) 74.
optimisation has been carried out for H± [8] R. Orlando, V.R. Saunders, R. Dovesi, unpublished.
CHA(3:1). By starting from the same initial ge- [9] V.R. Saunders, R. Dovesi, C. Roetti, M. Causa N.M.
Harrison, R. Orlando, C.M. Zicovich-Wilson, CR Y S T A L -
ometry as in the full geometry optimisation, but
98 User's Manual (Universita di Torino, Torino, 1999).
limiting the optimisation to the Brùnsted site [10] H.B. Schlegel, J. Comp. Chem. 3 (1982) 214.
(BSiO(H)AlB) only, the geometry relaxation takes [11] The generalisation of the code to gradients with respect to
20 cycles instead of 70. The ®nal geometry of the cell parameters is in progress.
Brùnsted site is very similar in both cases, the [12] J. Baker, J. Comp. Chem. 14 (1993) 1085.
[13] S. Swaminathan, B.M. Craven, R.K. McMullan, Acta
main structural features are (fragment optimisa-
Crystallogr. B 40 (1984) 300.
tion results in parentheses): rSiOb ˆ 1:740 (1.746); [14] M. Catti, G. Sandrone, G. Valerio, R. Dovesi, J. Phys.
rAlOb ˆ 1:835 (1.849); rOb H ˆ 0:975 (0.976); Chem. Solids 57 (1996) 1735.
\SiOb Al ˆ 140:0 (141.2); sSiOb (H)Al ˆ 179.3 [15] M. Catti, G. Valerio, R. Dovesi, M. Causa, Phys. Rev. B
( 179:7). 49 (1994) 14179.
138 B. Civalleri et al. / Chemical Physics Letters 348 (2001) 131±138

[16] M. Catti, B. Civalleri, P. Ugliengo, J. Phys. Chem. B 104 [24] R. Dovesi, M. Causa, R. Orlando, C. Roetti, V.R.
(2000) 7259. Saunders, J. Chem. Phys. 92 (1990) 7402.
[17] B. Civalleri, C.M. Zicovich-Wilson, P. Ugliengo, V.R. [25] Notes on the starting geometry and the CRYSTAL input
Saunders, R. Dovesi, Chem. Phys. Lett. 292 (1998) 394. decks can be found at the CRYSTAL home page: http://
[18] C. Zicovich-Wilson, R. Dovesi, J. Phys. Chem. 102 (1998) www.ch.unito.it/ifm/teorica/supplement/index.html.
1411. [26] H.B. Schlegel, Theor. Chim. Acta 66 (1984) 333.
[19] P. Ugliengo, B. Civalleri, C.M. Zicovich-Wilson, R. [27] T.H. Fisher, J. Alml of, J. Phys. Chem. 96 (1992)
Dovesi, Chem. Phys. Lett. 318 (2000) 247. 9768.
[20] E.H. Teunissen, A.P.J. Jansen, R.A. van Santen, R. [28] R. Lindh, A. Bernhardsson, G. Karlstr  Malmq-
om, P.-A
Orlando, R. Dovesi, J. Chem. Phys. 101 (1994) 5865. vist, Chem. Phys. Lett. 241 (1995) 423.
[21] A. Damin, R. Dovesi, A. Zecchina, P. Ugliengo, Surf. Sci. [29] C. Peng, P.Y. Ayala, H.B. Schlegel, M.J. Frisch, J. Comp.
479 (2001) 255. Chem. 17 (1996) 49.
[22] R. Orlando, P. Azavant, M.D. Towler, R. Dovesi, C. [30] K.N. Kudin, G. Scuseria, H.B. Schlegel, J. Chem. Phys.
Roetti, J. Phys.: Cond. Matter 8 (1996) 1123. 114 (2001) 2919.
[23] S. Casassa, P. Ugliengo, C. Pisani, J. Chem. Phys. 106 [31] J. Andzelm, R.D. King-Smith, G. Fitzgerald, Chem. Phys.
(1997) 8030. Lett. 335 (2001) 321.

You might also like