An Introductory Study of The Molecular Structure and Properties of Oligothiadiazoles

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Journal of Molecular Structure (Theochem) 634 (2003) 6776 www.elsevier.

com/locate/theochem

An introductory study of the molecular structure and properties of oligothiadiazoles


n, Daniel Glossman-Mitnik* Norma Flores-Holgu
mica Computacional, Centro de Investigacio n en Materiales Avanzados (CIMAV), Miguel de Cervantes 120, LAQUICOMLaboratorio de Qu Complejo Industrial Chihuahua, Chihuahua, Chih. 31109, Mexico Received 7 February 2003; accepted 31 March 2003

Abstract Several potentially conducting polymers, optically nonlinear polymers, and biomaterials contain heterocyclic structures. Reduction of the energy band gap of a conjugated polymer is a topic of considerable interest due to the possible elimination of doping in the preparation of highly conductive polymers. Control of the energy gap value of a polymer by molecular design could modify its optical, electronic and optoelectronic properties. Thiadiazoles and their derivatives are the structural basis of some of these polymeric materials. The results of the calculation of the HOMO LUMO gap, the dipole moment and polarizability of thiadiazole oligomers in vacuo and in the presence of solvents are reported. The calculations are based on density functional theory using a specially tailored model chemistry. The potential utility of these materials for the development of chemical sensors is discussed. q 2003 Elsevier B.V. All rights reserved.
Keywords: Thiadiazole oligomers; Density functional theory; Molecular structure; HOMOLUMO gap; Dipole moment; Polarizability

1. Introduction Nonlinear optics is expected to play a major role in the technology of photonics. Photonics is emerging as a multidisciplinary new frontier of science and technology that is capturing the imagination of scientists and engineers worldwide because of its potential applications to many areas of present and future information and image processing technologies. Examples of nonlinear optical phenomena that are potentially useful in this context are the ability to
* Corresponding author. Tel.: 52-614-439-1151; fax: 52-614439-1112. E-mail address: danielglossman@cimav.edu.mx (D. GlossmanMitnik).

alter the frequency or color of light and to amplify one source of light with another, switch it, or alter its transmission characteristics through a medium, depending on its intensity. It is the potential for providing these functions in suitable materials that motivates much of the current fundamental and exploratory research in the eld of nonlinear optics [1]. The thiadiazoles have been broadly applied in the areas of pharmaceutical, agricultural, industrial and polymer chemistry. The thiadiazole ring has been incorporated into the backbone of a number of polymers with desirable chemical and thermal stability. Thiadiazole polyamides, polyesters, polyethers and polyhydrazides are useful as packaging

0166-1280/03/$ - see front matter q 2003 Elsevier B.V. All rights reserved. doi:10.1016/S0166-1280(03)00253-7

68

n, D. Glossman-Mitnik / Journal of Molecular Structure (Theochem) 634 (2003) 6776 N. Flores-Holgu

materials, bers, electrical insulators and membranes for reverse osmosis [2,3]. The molecular and electronic structure of monocyclic thiadiazoles is of considerable interest because they are isoelectronic with thiophenes, thiazoles, pyrazines and related ve- and six-membered heterocycles [4]. These are systems of growing interest in materials science in view of the potential technological applications in elds such as electronic, nonlinear optics, sensors or corrosion protection [5,6]. The ultimate goal of computational chemistry is to understand chemical reactivity and to predict the outcome of molecular interactions. The behavior of atoms and molecules is characterized by some parameters that chemists have extracted from their experience and used for prediction of chemical reactivity. Electron population analysis [7,8], electronic density maps [8], molecular electrostatic potentials [9], the electronegativity as a measure of the tendency of a species to attract electrons [10], global and local hardness and softness [11 13] and Fukui functions [13], the HOMO LUMO gap [14, 15], total dipole moments [16,17] and polarizabilities [16,17] are examples of those parameters. Moreover, it is of interest to understand the inuence of the surrounding medium (i.e. the solvent) on the behavior of these mentioned molecular properties. As a further step in the study of the chemical reactivity of the thiadiazoles and related compounds, some results are presented in this paper on the calculation of thiadiazole oligomers with N up to six units, following previous work along the same line [4, 18 25] using nonlocal density functional theory (DFT) calculations [13]. The molecular structures, HOMO LUMO gaps [14,15], dipole moments m [16, 17] and polarizabilities a [16,17] have been determined in vacuo and in the presence of 18 solvents of different dielectric constant. The results have been plotted as a function of the number of units in order to get the extrapolated values for large N belonging to the corresponding polymers.

These are intermediate between being strongly bonded between nearest-neighbor atoms and being free-electron-like. Therefore, they respond relatively easy to external perturbations, for instance, due to structural or compositional changes [26]. Another area where these properties have become of great interest is the eld of nonlinear optics [1,27 30]. Any material that is exposed to an external electromagnetic eld will respond to it. The response may be quantied through either the dipole moment or the total energy. For the former one obtains X X X 0 mi m aij Ej bijk Ej Ek i
j jk jkl

gijkl Ej Ek El ;

where mi is the ith component of the dipole moment, 0 Ei that of the electromagnetic eld, m i is the static dipole moment, a is the polarizability, and b; g; are the rst, second, hyperpolarizabilities. For the total energy, one gets [26] X 0 Etot Etot 2 mi Ei ; 2
i

tot being the total energy in the absence of with the external eld. For free electrons, only a will be nonzero, whereas for strongly bonded electrons, all polarizabilities and hyperpolarizabilities will be very small [17]. The intermediate characteristics of the p electrons of conjugated polymers make both the polarizabilities and hyperpolarizabilities large compared with the values of other materials [26]. The DFT [13] provides a convenient theoretical framework for calculating global and local indices that quantitatively describe the inherent reactivity of chemical species. The dipole moment can be extracted from the output of any standard electronic structure program. Many experiments are done on isotropic systems (gases, neat liquids, and solutions) where the invariant vector and scalar components of a are measured. For example, the isotropic (or average) polarizability is given by [16] kal
1 3

0 Etot

2. Theory and computational details The interesting semiconducting properties of conjugated polymers can be ascribed to the p electrons.

axx ayy azz :

The other components of the polarizability (axy ; axz ; etc.) are not needed to obtain the isotropic quantity, saving considerable computing time.

n, D. Glossman-Mitnik / Journal of Molecular Structure (Theochem) 634 (2003) 6776 N. Flores-Holgu

69

The inuence of the solvent on the mentioned properties have been considered by resorting to the Self-Consistent Reaction Field (SCRF) methods. These methods model the solvent as a continuum of uniform dielectric constant e : the reaction eld. The solute is placed into a cavity within the solvent. SCRF approaches differ in how they dene the cavity and the reaction eld [31]. In particular, we have considered the Tomasis polarized continuum model which denes the cavity as the union of a series of interlocking atomic spheres. The effect of the polarization of the solvent continuum is represented numerically: it is computed by numerical integration rather by an approximation to an analytical form [32 39]. The solvents considered in this work include: heptane e 1:92; cyclohexane e 2:023; CCl4 e 2:228; benzene e 2:247; toluene e 2:379; diethylether e 4:335; chloroform e 4:9; chlorobenzene e 5:621; aniline e 6:89; THF e 7:58; dichloromethane e 8:93; dichloroethane e 10:36; acetone e 20:7; ethanol e 24:55; methanol e 32:63; acetonitrile e 36:64; nitromethane e 38:2; DMSO e 46:7 and water e 78:39: As a high accuracy is needed for a good description of properties that depend on the electronic structure, the HOMO LUMO gaps, dipole moments and polarizabilities were calculated with nonlocal DFT using the 3-21G* basis set through single point calculations on the fully optimized at the same level neutral molecule. The justication for choosing a medium size basis set was threefold: (1) Previous calculations by one of the authors [40] showed that the use of a 3-21G* basis set in conjunction with hybrid density functionals gives accurate results for the molecular structure and electronic properties of thiadiazole single molecules; (2) When doing calculations in the presence of a solvent, the different models lack analytical derivatives. Thus, the use of larger basis sets is more costly and computationally prohibitive; (3) Some authors [41 45] have shown recently that dipole moments calculated with the hybrid functionals are usually in good agreement with high level ab initio calculations and with experiment. The nonlocal density functional calculations have been performed by resorting to the B3LYP functional, which is formed by the Becke three parameter hybrid

functional (B3) [46] for the exchange part and the Lee, Yang and Parr (LYP) [47] for the correlation part. All the calculations were performed with the GAUSSIAN 98W series of programs [48].

3. Results and discussion The molecular structure of 1,3,4-thiadiazole hexamer and 1,2,5-thiadiazole hexamer are shown in Figs. 1 and 2, respectively. Every subunit of the oligomer chain retains the values for bond lengths and bond angles that were present in the original monomeric unit [4,19]. The subunits are planar on both molecules and in all the oligomers, with average , C N 1.31 A , and N values of S C 1.72 A , for the 1,3,4-thiadiazole oligomers, and N 1.35 A , C N 1.33 A , and C C 1.42 A , S N 1.64 A for the 1,2,5-thiadiazole oligomers. The main difference between 1,3,4-thiadiazole oligomers and 1,2,5thiadiazole oligomers is that the former is a linear and planar structure, while the second presents an entangled structure. This entangling is larger as the number of monomeric units grows. On the contrary, the 1,3,4-thiadiazole oligomers remain linear and planar even for a large number of units. This can be related to aromaticity, and in turn, to chemical reactivity. As has been already shown [25], the monomers are aromatic substances. Nevertheless, probably only the 1,3,4-thiadiazole oligomers will retain its aromatic character, while for the 1,2,5thiadiazole oligomers, the lost of planarity will result in a decrease of aromaticity. The HOMO LUMO gap, that is, the difference between the energy values of the HOMO and LUMO orbitals, can be considered a measure of chemical hardness [13], and it is a crucial result in order to understand the chemical reactivity of a molecule. Moreover, the HOMO LUMO gap of a polymer gives an indication of its possible use as a semiconductor. The HOMO LUMO gaps as a function of the number of monomeric units for the 1,3,4thiadiazole and 1,2,5-thiadiazole oligomers are presented in Figs. 3 and 4, both in vacuo and in the presence of solvents. We can see that the behavior is almost independent of the surrounding media. Although in this work, we have, in general, studied oligomers of up to six units, we have extended

70

n, D. Glossman-Mitnik / Journal of Molecular Structure (Theochem) 634 (2003) 6776 N. Flores-Holgu

Fig. 1. Molecular structure of the 1,3,4-thiadiazole hexamer.

Fig. 2. Molecular structure of 1,2,5-thiadiazole hexamer.

n, D. Glossman-Mitnik / Journal of Molecular Structure (Theochem) 634 (2003) 6776 N. Flores-Holgu

71

Fig. 3. HOMOLUMO gap of 1,3,4-thiadiazole oligomers as a function of the number of monomeric units.

Fig. 4. HOMOLUMO gap of 1,2,5-thiadiazole oligomers as a function of the number of monomeric units.

72

n, D. Glossman-Mitnik / Journal of Molecular Structure (Theochem) 634 (2003) 6776 N. Flores-Holgu

the range for the in vacuo studies up to 12 units. The usual way to study the behavior of oligomers with increasing number of units is to plot the studied property as an inverse function of the unit number. Notwithstanding, we have found that a better description is obtained when the results are tted to a decreasing exponential function. The resulting equations are DE 3:626 4:597 exp2N =1:755 for the 1,3,4-thiadiazole oligomers, and, DE 4:768 8:248 exp2N =0:559 5 4

for the 1,2,5-thiadiazole oligomers, where DE is the extrapolated HOMO LUMO gap and N is the number of monomeric units in both cases. Thus, the extrapolated values for N ! 1 are 3.626 eV for the 1,3,4-thiadiazole polymer and 4.768 eV for the 1,2,5-thiadiazole polymer. While this is in line with the calculated and experimental results for similar polymers (i.e. polythiophene) [6], it would be interesting for technical applications to decrease these values. This could be probably achieved by replacing the terminal hydrogens with electron-acceptor and electron-donor groups, as in push pull polymers [6]. The molecular dipole moment is perhaps the simplest experimental measure of charge density in a molecule. The accuracy of the overall distribution of electrons in a molecule is hard to quantify, since it involves all of the multipole moments. For the oligomers considered in this work, the study can be done by separating the oligomers with an even number of units from those with an odd number of monomeric units. When the number is even the symmetry of the molecules results in a zero value for the dipole moment in all cases. When the number is odd, there is a different behavior for the 1,3,4thiadiazole oligomers and the 1,2,5-thiadiazole oligomers. For the former, the dipole moment goes from 3.37 D for the monomer, to 3.30 D for the trimer, to 8.96 D for the pentamer. In contrast, for the 1,2,5thiadiazole oligomers present a dipole moment of 1.50 D for the monomer, 1.26 D for the trimer and 0.96 D for the pentamer. This behavior is probable related to the different molecular structures and can be related with the possible nonlinear optical properties for these molecules.

Fig. 5. Molecular structure of degemol (bis-2,5-1,3,4-thiadiazolyl1,3,4-thiadiazole).

n, D. Glossman-Mitnik / Journal of Molecular Structure (Theochem) 634 (2003) 6776 N. Flores-Holgu

73

One of the most interesting results of this work and which leads to striking conclusions, arises from the study of the polarizability of the oligomers as function of the dielectric constant of the different solvents. That is, a differential behavior of that property for any of the oligomers in any of the solvents will paved the way for the making of a chemical sensor for testing the presence of that solvent. We have considered the behavior of all the oligomers with the 18 solvents and for the sake of brevity they cannot be reproduced here. We will show only some representative results. The most interesting molecule is the all-trans1,3,4-thiadiazole trimer, or bis-2,5-1,3,4-thiadiazolyl-1,3,4-thiadiazole, which we have called degemol for short. The molecular structure of degemol is shown in Fig. 5. It is a linear and planar molecule and it will be probable aromatic. In Fig. 6, there is a representation of the polarizability of degemol in the presence of different solvents. It can be seen from Fig. 6 that the polarizability is the same in all the solvents, but for cyclohexane, where it presents a sharp minimum. That means, that degemol could be a material for making a sensor to detect cyclohexane. Moreover, the dielectric constant of cyclohexane is e 2:023; which is almost the same as the average dielectric constant for gasoline. Thus, degemol could be the basis for a sensor for the detection of gasoline leaks. Other thiadiazole oligomers show a similar behavior. For example, a representation of the polarizability of the 1,3,4-thiadiazole dimer given in Fig. 7 shows a shallow minimum in the presence of nitrometane e 38:2 and DMSO e 46:7: If we consider a similar picture for the 1,2,5-thiadiazole

Fig. 7. Polarizability of the 1,3,4-thiadiazole dimer as a function of the dielectric constants of the surrounding solvents.

pentamer, shown in Fig. 8, we can see that there is a sharp maximum of the polarizability in the presence of aniline e 6:89: Thus, any of these materials could have potential uses for making chemical sensors for the detection of nitromethane and DMSO, in the rst case, and aniline, in the second. However, it is not completely clear from a theoretical point of view, that the oligomer polarizabilities can be treated as a function solely of the solvent dielectric constants. Figs. 6 8 show that the solvents in which deviant polarizabilities are found have dielectric constants very similar to others in which the polarizability is not deviant. Thus, the deviant polarizabilities can be due to other extra factors, which investigation is now being pursued in our laboratory. If we think of possible technological applications of these materials, we should consider the processability

Fig. 6. Polarizability of degemol as a function of the dielectric constant of the surrounding solvents.

Fig. 8. Polarizability of the 1,2,5-thiadiazole pentamer as a function of the dielectric constants of the surrounding solvents.

74

n, D. Glossman-Mitnik / Journal of Molecular Structure (Theochem) 634 (2003) 6776 N. Flores-Holgu

Fig. 9. Solvation energies of 1,3,4-thiadiazole oligomers as a function of the number of monomeric units for the different solvents considered in this work.

Fig. 10. Solvation energies of 1,2,5-thiadiazole oligomers as a function of the number of monomeric units for the different solvents considered in this work.

n, D. Glossman-Mitnik / Journal of Molecular Structure (Theochem) 634 (2003) 6776 N. Flores-Holgu

75

and solubility of these oligomers. While the actual solubility of a material depends on several factors, a brief indication can be given by its free energy of solvation DGs : A representation of the DGs of the oligomers as a function of the number of monomeric units and the dielectric constant of the solvent are presented in Figs. 9 and 10, for the 1,3,4-thiadiazole oligomers and 1,2,5-thiadiazole oligomers, respectively. A DGs lower than zero, would be an indication of the solubility of that material. It can be seen from the gures, that while degemol will be soluble in most of the solvents considered here, the 1,2,5-thiadiazole oligomers will be less soluble or insoluble at all, in most of the solvents, specially in aniline. This should be taken into account when considering any of these oligomers as starting materials for the making of chemical sensors. 4. Conclusions In this work, we have shown calculations of 1,3,4thiadiazole oligomers and 1,2,5-thiadiazole oligomers with up to six units performed using hybrid density functional methods (B3LYP) at the 3-21G* basis set level. We have studied the properties of these oligomers in vacuo and in the presence of 18 different solvents. The molecular structures predicted with the B3LYP functional are an indication that the 1,3,4thiadiazole oligomers are linear, planar and possibly aromatic materials, while the 1,2,5-thiadiazole oligomers are entangled molecules, which can cage metal atoms in its structures. The HOMO LUMO gap has been studied in vacuo and in solution as a function of the number of monomeric units. The values has been tted to a decreasing exponential function and the extrapolated results indicate a DE of 3.63 eV for the 1,3,4thiadiazole polymer and a DE of 4.77 eV for the 1,2,5-thiadiazole polymer. These values could be probably lowered by replacing the terminal hydrogens with electron-acceptor and electron-donor groups, as in push pull polymers. The dipole moment of the oligomers can be explained in terms of the symmetry of the molecules, where even-numbered oligomers have zero dipole moment, while odd-numbered oligomers have nonzero dipole moments, with values ranging from 0.96 to 8.96 D.

The polarizability a of all the molecules has been studied as a function of the dielectric constant of the solvents. Of special interest is the all-trans-1,3,4thiadiazole trimer or degemol, which presents a sharp minimum in the polarizability in the presence of cyclohexane. As the dielectric constant of cyclohexane is similar to that of gasoline, this molecule could be the starting material for making chemical sensors useful for the detection of gasoline leaks. Other interesting materials found in this study are the 1,3,4-thiadiazole dimer, which could be useful for sensing nitromethane and DMSO, and 1,2,5-thiadiazole pentamer, which is potentially a sensor for aniline. However, as it is not clear that the oligomer polarizabilities can be treated as a function solely of the solvent dielectric constants, there is a need for additional research in order to understand if the behavior of the deviant polarizabilities can be due to other extra factors. This investigation is now being pursued in our laboratory. The free energies of solvation DGs of the oligomers, which in absence of other factors can be considered as an indication of solubility, show that the 1,3,4-thiadiazole oligomers are soluble in most of the solvents, while the 1,2,5-thiadiazole oligomers are partially soluble, or, specially the largest, insoluble in most of the solvents. Acknowledgements This work has been supported by Consejo Nacional a (CONACYT, Mexico) under de Ciencia y Tecnolog Grant No. 34697-U/2000. DGM is a researcher of CONACYT and CIMAV. NFH gratefully acknowledges a fellowship from CONACYT. References
[1] P. Prasad, D. Wiliams, Introduction to Nonlinear Effects in Molecules and Polymers, Wiley, New York, 1991. [2] L.M. Weinstock, I. Shinkai, in: K.T. Potts (Ed.), Comprehensive Heterocyclic Chemistry, vol. 6, Pergamon Press, Oxford, UK, 1984. n, P. Goya, C. Ochoa, Adv. Heterocycl. Chem. 44 [3] V. Ara (1988) 81. [4] D. Glossman, J. Mol. Struct. (Theochem) 330 (1995) 385. [5] P. Strohriegl, J. Grazulevicius, in: H. Nalwa (Ed.), Handbook

76

n, D. Glossman-Mitnik / Journal of Molecular Structure (Theochem) 634 (2003) 6776 N. Flores-Holgu of Organic Conductive Molecules and Polymers, Wiley/VCH, Chichester, UK, 1997. K. Mullen, G. Wegner (Eds.), Electronic Materials: The Oligomer Approach, Wiley/VCH, Weinhein, FRG, 1998. (a) R.S. Mulliken, J. Chem. Phys. 23 (1955) 1833. (b) R.S. Mulliken, J. Chem. Phys. 23 (1955) 1841. (c) R.S. Mulliken, J. Chem. Phys. 23 (1955) 2338. S.M. Bachrach, in: K.B. Lipkowitz, D.B. Boyd (Eds.), Reviews in Computational Chemistry, vol. 5, Wiley/VCH, New York, 1994. P. Politzer, D.G. Truhlar (Eds.), Chemical Applications of Atomic and Molecular Electrostatic Potentials, Plenum Press, New York, 1981. R.S. Mulliken, J. Chem. Phys. 2 (1934) 782. R.G. Pearson, J. Am. Chem. Soc. 85 (1963) 3533. R.G. Pearson, Science 151 (1966) 172. R.G. Parr, W. Yang, W. Density, Functional Theory of Atoms and Molecules, Oxford University Press, New York, 1989. Z. Zhou, R.G. Parr, Tetrahedron Lett. 29 (1988) 4843. R.G. Pearson, J. Chem. Educ. 64 (1987) 561. H. Kurtz, D. Dudis, in: K.B. Lipkowitz, D.B. Boyd (Eds.), Reviews in Computational Chemistry, vol. 14, Wiley/VCH, New York, 1998. K. Bonin, V. Kresin, Electric-Dipole Polarizabilities of Atoms, Molecules and Clusters, World Scientic, Singapore, 1997. sico-Qu mica Orga nica M.D. Glossman, Atualidades de F 1995, Universidade Federal de Santa Catarina Press, Flor polis, Brazil, 1996. iano M.D. Glossman, J. Mol. Struct. (Theochem) 390 (1997) 67. co, S.L. Aimone, E.J. E.E. Castellano, O.E. Piro, M.V. Mir Vasini, M.D. Glossman, J. Phys. Org. Chem. 11 (1995) 91. co, J.A. Caram, J.A.D. GlossmanS.L. Aimone, M.V. Mir Mitnik, O.E. Piro, E.E. Castellano, E.J. Vasini, Tetrahedron Lett. 41 (2000) 3531. rquez-Lucero, Int. J. Quant. D. Glossman-Mitnik, A. Ma Chem. 81 (2001) 105. co, S.L. (a) E.E. Castellano, O.E. Piro, J.A. Caram, M.V. Mir rquez-Lucero, D. GlossmanAimone, E.J. Vasini, A. Ma Mitnik, J. Mol. Struct. 562 (2001) 157. (b) E.E. Castellano, co, S.L. Aimone, E.J. O.E. Piro, J.A. Caram, M.V. Mir rquez-Lucero, D. Glossman-Mitnik, J. Mol. Vasini, A. Ma Struct. 597 (2001) 163. (c) E.E. Castellano, O.E. Piro, J.A. co, S.L. Aimone, E.J. Vasini, A. Ma rquezCaram, M.V. Mir Lucero, D. Glossman-Mitnik, J. Mol. Struct. 604 (2002) 195. rquez-Lucero, J. Mol. Struct. (a) D. Glossman-Mitnik, A. Ma (Theochem) 535 (2001) 39. (b) D. Glossman-Mitnik, A. rquez-Lucero, J. Mol. Struct. (Theochem) 536 (2001) 41. Ma rquez-Lucero, J. Mol. Struct. (c) D. Glossman-Mitnik, A. Ma (Theochem) 538 (2001) 201. (d) D. Glossman-Mitnik, A. rquez-Lucero, J. Mol. Struct. (Theochem) 548 (2001) 153. Ma D. Glossman-Mitnik, J. Mol. Struct. (Theochem) 549 (2001) 285. a, in: F. M. Springborg, K. Schmidt, H. Meider, L.D. Mar Farchioni, G. Grosso (Eds.), Organic Electronic Materials Conjugated Polymers and Low Molecular Weight Organic Solids, Springer, Berlin, 2001. [27] J. Messier, F. Kajzar, P. Prasad, P.D. Ulrich, Nonlinear Optical Effects in Organic Polymers, Kluwer, Dordrecht, 1989. das, R. Chance, Conjugated Polymeric Materials: [28] J. Bree Opportunities in Electronics and Molecular Electronics, Kluwer, Dordrecht, 1990. [29] J. Messier, F. Kajzar, P. Prasad, Organic Molecules for Nonlinear Optics and Photonics, Kluwer, Dordrecht, The Netherlands, 1991. [30] S. Karna, A. Yeates, Nonlinear Optical Materials: Theory and Modelling, American Chemical Society, Washington, DC, 1996. [31] J. Foresman, A. Frisch, Exploring Chemistry with Electronic Structure Methods, Gaussian Inc, Pittsburgh, 1996. [32] S. Miertus, E. Scrocco, J. Tomasi, J. Chem. Phys. 55 (1981) 117. [33] S. Miertus, J. Tomasi, Chem. Phys. 65 (1982) 239. [34] M. Cossi, V. Barone, R. Cammi, J. Tomasi, Chem. Phys. Lett. 255 (1996) 327. [35] M. Cances, B. Mennucci, J. Tomasi, J. Chem. Phys. 107 (1997) 3032. [36] V. Barone, M. Cossi, B. Mennucci, J. Tomasi, J. Chem. Phys. 107 (1997) 3210. [37] M. Cossi, V. Barone, B. Mennucci, J. Tomasi, Chem. Phys. Lett. 286 (1998) 253. [38] V. Barone, M. Cossi, J. Tomasi, J. Comp. Chem. 19 (1998) 404. [39] V. Barone, M. Cossi, J. Phys. Chem., A 102 (1998) 1995. [40] D. Glossman-Mitnik, J. Mol. Struct. (Theochem) (2003) in press. [41] (a) F. DeProft, J. Martin, P. Geerlings, Chem. Phys. Lett. 256 (1996) 393. (b) F. DeProft, J. Martin, P. Geerlings, Chem. Phys. Lett. 256 (1996) 400. [42] F. DeProft, F. Tielens, P. Geerlings, J. Mol. Struct. (Theochem) 506 (2000) 1. [43] B. Jursic, J. Mol. Struct. (Theochem) 507 (2000) 11. [44] N. Jayakumar, P. Kolandaivel, N. Kuze, T. Sakasumi, O. Ohashi, J. Mol. Struct. (Theochem) 465 (1999) 197. [45] P. Kolandaivel, N. Jayakumar, J. Mol. Struct. (Theochem) 507 (2000) 197. [46] A. Becke, J. Chem. Phys. 98 (1993) 5648. [47] C. Lee, W. Yang, R.G. Parr, Phys. Rev., B 37 (1988) 785. [48] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, V.G. Zakrzewski, J.A. Montgomery, Jr., R.E. Stratmann, J.C. Burant, S. Dapprich, J.M. Millam, A.D. Daniels, K.N. Kudin, M.C. Strain, O. Farkas, J. Tomasi, V. Barone, M. Cossi, R. Cammi, B. Mennucci, C. Pomelli, C. Adamo, S. Clifford, J. Ochterski, G.A. Petersson, P.Y. Ayala, Q. Cui, K. Morokuma, P. Salvador, J.J. Dannenberg, D.K. Malick, A.D. Rabuck, K. Raghavachari, J.B. Foresman, J. Cioslowski, J.V. Ortiz, A.G. Baboul, B.B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. Gomperts, R.L. Martin, D.J. Fox, T. Keith, M.A. Al-Laham, C.Y. Peng, A. Nanayakkara, M. Challacombe, P.M.W. Gill, B. Johnson, W. Chen, M.W. Wong, J.L. Andr/es, C. Gonz/alez, M. HeadGordon, E.S. Replogle, J.A. Pople, J.A. Gaussian, 98, Revision A.11, Gaussian, Pittsburgh, 2001.

[6] [7]

[8]

[9]

[10] [11] [12] [13] [14] [15] [16]

[17]

[18]

[19] [20] [21]

[22] [23]

[24]

[25] [26]

You might also like