You are on page 1of 9

WWW.Q-CHEM.

ORG

FULL PAPER

Intramolecular CH/p Interactions in Alkylaromatics: Monomer Conformations for Poly(3-alkylthiophene) Atomistic Models
Alberto Baggioli,[a] Stefano V. Meille,[a] Guido Raos,[a] Riccardo Po,[b] Martin Brinkmann,[c] and Antonino Famulari*[a]
In existing poly(3-alkylthiophenes) atomistic models, an extended conformation of the side chain is usually assumed. We report a first principle study of the side-chain energetics of 3-hexylthiophene, with the constraint of compatibility with crystal packing requirements. The first two torsion angles of the side chain closest to the ring were considered. Electron correlation is shown to be of great relevance in the assessment of the relative stability of folded conformers against extended ones. The roles of local charge-transfer, rehybridization, steric repulsion, and basis set superposition error, were all considered in the rationalization of our results. We extend our analysis to the thiophene/methane complex in order to elucidate the main differences between intermolecular and intramolecular CH/p phenomena. While in the noncovalent complex a single CAH bond mediates the interaction, folded arrangements of 3-alkylthiophenes require the collective effort C 2013 Wiley Periodicals, Inc. of several aliphatic bonds. V

DOI: 10.1002/qua.24472

Introduction
Due to their remarkable optoelectronic properties combined with ease of fabrication and low-cost manufacturing, semiconducting oligomers and polymers are promising materials for a variety of applications in organic electronics.[1,2] Among conjugated polymers, poly(3-hexylthiophene) (P3HT) is one of the most extensively used and investigated. The high carrier mobility coupled with solution processability has allowed its application in devices such as light-emitting diodes and thinfilm transistors.[3,4] In combination with TiO2, ZnO, carbon nanotubes, or phenyl-C61-butyric acid methyl ester, it is also a well-known donor material in photovoltaic devices.[5,6] Regioregular P3HT is a semicrystalline polymer, exhibiting at least three polymorphs (the so-called forms I, I, and II) with markedly different properties.[7,8] A knowledge of their precise three-dimensional structure, or at least a limit-ordered approximation of it, would be desirable in order to correlate these properties with supramolecular organization.[9,10] The literature contains several semiquantitative[7] and a couple of atomistically detailed models of P3HT crystal structures[8,11] but, surprisingly for many workers in the field, even the latter are not fully satisfactory. The determination of detailed polymer crystal structures typically resorts to the analysis of experimental data such as powder or fiber X-ray and electron diffraction patterns. However, these are difficult to interpret quantitatively without the support of additional information, such as that from molecular modeling or solid-state spectroscopies. In addition, the construction of polymer crystal structures is generally based on the equivalence postulate of monomeric units and on the minimum energy conformation principle.[12] The latter states that the packed macromolecules tend to adopt one of the stable
2154 International Journal of Quantum Chemistry 2013, 113, 21542162

conformations of the isolated polymer chain, very often that of the absolute minimum. The study of homologous oligomers, which can be conveniently obtained as single crystals, can also be extremely useful. The interpretation of their X-ray diffraction does not rely on any a priori assumptions, and the structures thus obtained may provide important insights. Several quantum chemical and simulation studies have been performed to help the construction of atomistically detailed models of P3HT.[1317] Many of them have been devoted to the problem of coplanarity of successive rings and to side chains effects,[1822] but even these resorted to some a priori assumptions. A typical one is the replacement of the hexyl by ethyl or even methyl side chains, in order to speed up the calculations and limit the size of the conformational problem. Rotation about the a torsion angle adjacent to the thiophene ring (see Fig. 1 for the definition of a, b, c,, f) was also investigated,[9,20] but to the best of our knowledge no further exploration of the side chain conformations can be found in the literature. This kind of investigation appears to be especially important when the side chains contain more than two
[a] A. Baggioli, S. V. Meille, G. Raos, A. Famulari Dipartimento di Chimica, Materiali e Ingegneria Chimica G. Natta Politecnico di Milano , Via L. Mancinelli 7, 20131 Milano, Italy E-mail: antonino.famulari@polimi.it [b] R. Po Research Center for Non-Conventional Energies, Istituto Donegani ENI S.p.A., Via Fauser 4, 28100 Novara, Italy [c] M. Brinkmann  de Strasbourg, 23 rue du Loess, Institut Charles Sadron, CNRS-Universite Strasbourg, 67034 France Contract grant sponsors: ENI SpA, CARIPLO (PLENOS, ref 2011-0349), Regione Lombardia, and CILEA/CINECA (Laboratory for Interdisciplinary Advanced Simulation Initiative).
C 2013 Wiley Periodicals, Inc. V

WWW.CHEMISTRYVIEWS.ORG

WWW.Q-CHEM.ORG

FULL PAPER

Figure 1. Definition of the torsion angles of the aliphatic moiety of 3-hexylthiophene; a is the dihedral formed by C(4)AC(3)AC(6)AC(7). [Color figure can be viewed in the online issue, which is available at wileyonline library.com.]

carbon atoms, that is, when they approach the length required to make poly(3-alkylthiophenes) easily soluble in organic solvents and processable around room temperature. Indeed, there are a few oligothiophene crystal structures in the Cambridge Structural Database (CSD)[23] which display a gauche conformation of b and the successive torsion angles, even though anti arrangements are generally dominant. The consequences of this observation have not yet been fully appreciated. It has been considered almost obvious that any deviation from the anti arrangement was caused by intermolecular packing effects. Instead, our present calculations on 3hexylthiophene demonstrate that this model system can have several low-energy folded states, which are characterized by specific intramolecular CH/p interactions and are compatible with crystal packing requirements.

geometries), MP2/aug-cc-pVTZ[30]//MP2, and CCSD(T)/6311G(d,p)//MP2 levels. DFT calculations were also carried out on 3-hexylthiophene in order to evaluate the performance of different functionals and the impact of the dispersion corrections, which are becoming increasingly popular. First of all, we used the well-known B3LYP hybrid functional,[31] without any dispersion correction (this is not implemented in Gaussian). Next, we considered some of the newer functionals, which were designed to provide an increasingly accurate description of van der Waals interactions.[32] The first one is a long-range corrected hybrid functional with damped atomatom dispersion (xB97XD).[33,34] The following ones are the double hybrid functionals B2PLYP-D[35,36] and mPW2PLYP-D.[36,37] These three functionals were also tested without dispersion correction (xB97X, B2PLYP, and mPW2PLYP). In all DFT calculations, we performed fullgeometry optimization with the 6-311G(d,p) basis set and a pruned 99,590 integration grid (ultrafine Gaussian keyword). The 6-3111G(2d,2p) basis set was also used in additional B3LYP calculations. Natural bond orbital (NBO) analysis were performed using the NBO 3.1 module[38,39] included into the Gaussian package.

Results and Discussion


The dihedral angles of the side chain of 3-hexylthiophene, labeled with Greek letters from a to f starting from the bond linking the thiophene ring to the alkyl group (see Fig. 1), were systematically explored through a series of fully relaxed geometry optimizations. A large number of folded conformations have been identified, for which a stabilizing interaction is established between the aliphatic chain and the p system of the thiophene ring (see Fig. 2 for some examples). This weak hydrogen bond, often referred to as CH/p interaction, appears to be only slightly affected by the number of aliphatic carbons interacting with the aromatic fragment, as also reported by Tsuzuki and coworkers[40] for 1:1 dimers of

Computational Details
The calculations were performed mainly using Gaussian 09 suite of programs.[24] In some cases, GAMESS-US code was used.[25] We used a combination of Mller-Plesset second-order perturbation theory (MP2),[26] Coupled Cluster with Single and Double excitations with noniterative Triple excitations correction (CCSD(T))[27] and Density Functional Theory with and without empirical Dispersion correction (DFT and DFT-D). The two wavefunction-based methods provide a solid ground for assessing the performance of the latter, in the perspective of prosecuting this study through the investigation of larger systems and the development of general-purpose molecular mechanics force fields for conjugated polymers. Calculations on 3-alkylthiophenes (all side chains from propyl to octyl, plus dodecyl), hexylbenzene, and n-hexane were carried out at the MP2/6-311G(d,p)[28] level, with full geometry optimizations. These represent the majority of our calculations and we will refer to this level of theory as MP2 for ease of discussion. Vibrational frequencies were computed to verify that the fully optimized structures indeed correspond to potential energy minima. Torsion potential energy curves were generated by a series of constrained geometry optimizations. One dihedral angle at a time was incremented in steps of 2 5 , depending on the desired level of detail. Additional calculations on 3-hexylthiophene were performed at MP2/ 6-3111G(2d,2p)[28,29] (again with full reoptimization of the

Figure 2. A selection of stable conformers of 3-hexylthiophene obtained at MP2/6-311G(d,p) level, along with the corresponding values of a, b, and c torsion angles. Energies relative to a. are given in brackets. [Color figure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

International Journal of Quantum Chemistry 2013, 113, 21542162

2155

FULL PAPER

WWW.Q-CHEM.ORG

Figure 3. Potential energy of 3-propylthiophene and 3-hexylthiophene as a function of the a angle (see Fig. 1). The continuous curves represent the MP2/6-311G(d,p) relative energies, the squares are the CCSD(T)/6-311G(d,p) energies at the MP2/6-311G(d,p) stationary points of 3-hexylthiophene. The roman numerals correspond to the conformations in Figure 5. [Color figure can be viewed in the online issue, which is available at wileyonlinelibrary. com.]

benzene with several small alkanes (see Figs. 2a and 2e2g). Our calculations indicate that any conformation of the side chain in which this interaction is absent (e.g., conformers c and d in Fig. 2) is destabilized with respect to the all-anti staggered arrangement (Fig. 2b), as one would expect for a short linear alkane. In addition, extremely folded conformers, for example, conformers 2e2f in Figure 2, characterized by several non-anti dihedral angles, would also hinder the assembly of the thiophene moieties in a stacked pp or in a herringbone fashion, key features of the crystal structures of many aromatic systems. For these reasons, we shall limit our discussion to the first two torsion angles of the aliphatic chain, a and b, while angles from c to f will be set in staggered arrangement (180 ) and then relaxed. Conformations defined by the angle a (see Fig. 1) were investigated first. The b angle was set to 180 at this stage, implying an all-anti arrangement of the side chain that did not change significantly during the subsequent optimizations. Potential energy curves as a function of a have been computed at the MP2 level for 3-propylthiophene and 3-hexylthiophene and they are shown in Figure 3. As can be seen, two equivalent out-of-plane conformations (I and I) are found at a 5 673 . There are also two in-plane conformations representing a local minimum (II, at 6180 ) and the absolute maximum, 0.8 and 1.8 kcal/mol, respectively, higher in energy. There is a remarkably low energy barrier for the interconversion of II into I. The MP2 potential energy curve for the internal rotation about the b angle of 3-hexylthiophene is given in Figure 4a. For comparison, we give also the analogous curves for hexylbenzene and n-hexane (for the latter, we consider the C(1)AC(2)AC(3)AC(4) torsion). The a angle of hexylthiophene and hexylbenzene corresponds to the out-of-plane conformation I, but was fully relaxed at every b (see Fig. 4b). As expected in the case of n-hexane, the anti arrangement (b 5 6180 ) represents the most stable conformation and the two equivalent gauche conformers (b 5 660 ) are 0.42 kcal/mol
2156 International Journal of Quantum Chemistry 2013, 113, 21542162

Figure 4. a) Potential energies of 3-hexylthiophene, hexylbenzene and nhexane as a function of the b angle (see Fig. 1). The continuous curves represent the results of constrained MP2/6-311G(d,p) optimizations, the squares are the CCSD(T)/6-311G(d,p) energies at the MP2/6-311G(d,p) stationary points of 3-hexylthiophene. b) Values of a angle as a function of b. [Color figure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

higher in energy. In the case of hexylthiophene and hexylbenzene, three minima and three maxima are found as well. The minima correspond to an anti conformation (b 5 180 ), which is essentially I, and to two gauche conformations III and IV at b 5 660 . See Figure 5 for a definition of conformers I, II, III, and IV, and Table 1 for their relative energies. Maxima on the curves correspond to the syn (b 5 0 ) and to two anticlinal (b 5 6120 ) conformations. Note that the two gauche conformers of hexylbenzene are symmetry related, whereas those of hexylthiophene are not, so that III is the absolute minimum. The lower symmetry of the thiophene ring (compared to the benzene one) implies an asymmetry of the a vs. b curve shown in Figure 4b. A remarkable difference between n-hexane and the other two compounds lies in the relative stability of the gauche and anti conformers. Conformers III and IV of 3-hexylthiophene are more stable than conformer I by 1.01 kcal/mol and 0.58 kcal/ mol at MP2 level, respectively. Similarly, the antigauche gap for hexylbenzene is 0.71 kcal/mol. Interconversion barriers between minima are in the same range for all molecules (35 kcal/mol). The order of stability revealed for 3-hexylthiophene (III, IV, I, and then II) is also found for all the 3-alkylthiophenes. Figure 6 shows the MP2 energy gaps between these conformations
WWW.CHEMISTRYVIEWS.ORG

WWW.Q-CHEM.ORG

FULL PAPER

Figure 5. Minimum energy conformations of 3-hexylthiopene and corresponding values of a and b torsion angles. Interatomic distances are . [Color figure can be viewed in the online issue, which is reported in A available at wileyonlinelibrary.com.]

for the whole series. Longer chains behave exactly as the hexyl, whereas shorter ones display some slight differences. It is interesting to attempt a rationalization of the phenomenon. This is not straightforward, for at least two reasons: there is always a certain degree of arbitrariness in any decomposition of an ab initio energy and, furthermore, this energy has spurious contributions originating from the basis set superposition error (BSSE).[4143] The latter is expected to have a greater impact on the folded conformations than on the extended one due the lower mean internuclear distance causing larger overlap between orbitals. In order to rule out a significant intramolecular BSSE contribution, we have performed a single-point counterpoise correction at the MP2 level on the four stable conformers of 3-hexylthiophene. The intramolecular counterpoise procedure involves a partitioning of the molecule in separate segments, resulting in the virtual fracture of covalent bonds.[44,45] Several segmentation patterns have been tested in the literature for both aliphatic and aromatic moieties, but there is still no universally accepted method. To be conservative, we chose to apply four different schemes, segmenting only bonds between heavy atoms in the side chain (case A), in the aromatic ring (case B), in both of them (case C), or just the one between the ring and the chain (case D). The results are collected in Table 2. None of these corrections

produced relevant effects on the relative stabilities of the conformers. These outcomes, together with the results of calculations performed with larger basis sets, prove that the folded gauche arrangements of b are not artificially stabilized by the BSSE. We have also carried out a harmonic free energy calculation ( T 5 298 K) at MP2 level to investigate a possible vibrational stabilization of the anti against the gauche arrangements. We obtained the same order of stability: taking III as reference for the energy, we found IV at 0.38 kcal/mol, I at 0.66 kcal/mol and II at 0.95 kcal/mol. The results of our systematic investigation of the effects of the theory level (MP2, CCSD(T) and variants of DFT) and basis set are summarized in Table 1. All the calculations confirm the outlined order of conformational stability. The CCSD(T) results, in particular, show a IIII energy gap of 0.65 kcal/mol, notably lower than from the MP2 estimates, while the relative stability of the two gauche conformers III and IV remains almost unchanged. As DFT has become the standard computational chemistry method,[46,47] and acknowledging that the recent introduction of explicit dispersion terms has improved its ability to describe subtle nonbonded interactions,[32,48] we have carried out a full reoptimization of all four conformers of 3-hexylthiophene with a selection of density functionals. Doubly hybrid B2PLYP and mPW2PLYP, long-range corrected hybrid xB97X were used along with their dispersion-corrected implementation. Geometrical parameters obtained at all DFT levels correlate well with MP2 results, the only shortcoming being a slightly smaller a angle in the IV-like structures, this being found in the range of 9092 . Overall, dispersion-corrected functionals perform much better than the noncorrected counterparts when it comes to energy estimation. The data in Table 1 show that dispersioncorrected functionals yield the most consistent results, while the noncorrected ones predict a reverted IIV order of stability. All DFT methods underestimate the energy of conformer II. Table 1 also reports B3LYP results in combination with two basis sets of different size. Despite its reliability for general-purpose applications, several studies have highlighted problems in the B3LYP description of van der Waals and hydrogen-

Table 1. Relative energies (kcal/mol) of the four minimum energy conformations of 3-hexylthiophene computed at various levels of theory. Conformer Method MP2/6-311G(d,p) [5MP2] MP2/aug-cc-pVTZ//MP2 MP2/6-3111G(2d,2p) CCSD(T)/6-311G(d,p)//MP2 B2PLYP/6-311G(d,p) mPW2PLYP/6-311G(d,p) xB97X/6-311G(d,p) B2PLYP-D/6-311G(d,p) mPW2PLYP-D/6-311G(d,p) xB97X-D/6-311G(d,p) B3LYP/6-311G(d,p) B3LYP/6-3111G(2d,2p) I 1.01 0.85 0.79 0.65 0.14 0.18 0.29 0.64 0.54 0.67 0.00 0.00 II 1.84 1.27 1.33 1.25 0.73 0.72 0.46 1.07 0.96 0.94 0.51 0.46 III 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.25 0.42 IV 0.43 0.43 0.45 0.49 0.51 0.52 0.54 0.47 0.49 0.52 0.79 0.94

International Journal of Quantum Chemistry 2013, 113, 21542162

2157

FULL PAPER

WWW.Q-CHEM.ORG

Figure 6. Relative energies of stable conformations of a series of n-alkylthiophene molecules computed at MP2/6-311G(d,p) level as a function of the number of carbon atoms in the side chain. [Color figure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

bonded complexes.[49,50] Indeed, also our tests return a different order of relative stabilities, with conformer I as the absolute minimum. Increasing the number of basis functions brings conformers II and III very close in energy, further aggravating the discrepancy with all the other methods. At the same time, the B3LYP prediction of the relative energies of the n-hexane conformations is qualitatively correct, resulting in a slightly overestimated gaucheanti gap of 0.84 and 0.89 kcal/mol with 6-311G(d,p) and 6-3111G(2d,2p) basis sets, respectively. This confirms that the main deficiency of B3LYP is in the estimation of the dispersive component. Based on the energetic aspects discussed above, we can conclude that the relative stability observed for 3-alkylthiophenes (and alkylbenzene) stems primarily from dispersion contributions involving the aromatic ring, and is not to be considered an artificial result of the computational method. We further investigated this topic by performing a NBO[39] analysis on conformers I, III, and IV. NBOs are a set of localized orbitals designed to provide an accurate description of the Lewis electronic structure of a molecular system. In other words, it returns a valence bond-type representation of the

Table 2. Results of the intramolecular counterpoise analysis on the conformers of 3-hexylthiophene. Segmentation pattern Reference Has the thiophene ring been segmented? Has the side chain been segmented? Fragments composing the thiophene ring Fragments composing the side chain Total number of fragments Relative energy of conformer I Relative energy of conformer II Relative energy of conformer III Relative energy of conformer IV No No 1 1.01 1.84 0.00 0.43 A No Yes 1 6 7 0.49 1.34 0.00 0.40 B Yes No 5 1 6 0.40 1.20 0.00 0.44 C Yes Yes 5 6 11 0.68 1.62 0.00 0.43 D No No 1 1 2 0.22 0.94 0.00 0.44

All calculations are carried out at MP2/6-311G(d,p) level. Corrected relative energies in kcal/mol.

wavefunction, formally independent from the basis set used to describe it. The so called Natural Population Analysis produced is also found to be virtually unaffected by the extension of the basis set, and its electron population sums properly to the actual total number of electrons in the system, leading to a univocal set of net atomic charges. In addition, the Rayleigh dinger perturbation principle applied to the Fock matrix Schro in the NBO basis allows the estimation of the deviation from the ideal Lewis electronic structure, in terms of electron density delocalization from all possible high-occupancy donor and low-occupancy acceptor pairs of NBOs. The magnitude of each delocalization event is computed as a function of the corresponding off-diagonal element of this effective Hamiltonian, of the occupancy of the donor orbital, and of the reciprocal of the donoracceptor NBOs energy gap. The formation of this CH/p interaction upon torsion of the b angle leads to some important geometrical deformation, moving from conformer I to conformers III and IV. In particular, relevant differences are observed for C(3)AC(6)AC(7) and C(6)AC(7) AC(8) angles (u1 and u2 hereafter), and all bonds involving these nuclei and the hydrogen atoms bound to them. Angles u1 and u2 in conformer III are slightly reduced with respect to I by a few tenths of a degree. In conformer IV the u2 angle is increased by the same order of magnitude. From these observations we can assume that the steric repulsion between aromatic ring and side chain in IV is somewhat stronger than that in III. Changes in the bond lengths among the methylene groups involved in the gauche conformation of the side chain are also observed. In particular, using conformer I as a reference, the C(8)AHa bonds in III and IV are contracted by as , while the C(8)AHb bonds are elongated by much as 0.002 A 0.001 A. In addition, the C(6)AC(7) bond in both gauche con with respect to I. formers is elongated by almost 0.003 A Changes of covalent-bond lengths are well known and well documented phenomena associated with the formation of hydrogen-bonding interactions.[5159] For a generic XAHY hydrogen bond, a spectral shift is expected for the XAH IR stretching frequency, either blue (bond contraction, higher frequencies) or red (bond elongation, lower frequencies).[60] Furthermore, covalent bonds in remote regions of the donor species have been reported to exert appreciable spectral shifts.[51] There has been considerable debate over the past decades concerning the underlying physical phenomena involved in the formation of hydrogen-bonded complexes,[39,51,53,56,6164] and special attention has been paid to the origin of these bond-length changes.[5155,57,6265] In spite of the initial controversies, a general consensus has been reached about the ascription to hydrogen bond of several interactions involving chemical species other than highly polarized XAH proton-donors and lone pair electrons on a proton-acceptor, including the CH/p interaction.[66] As of bond-length changes, Alabugin et al. indicated the subtle balance between nY!r*XH hyperconjugation and rXH rehybridization as a key factor in the rationalization of the phenomenon.[57] Electron density transfer to the antibonding orbital of the proton-donor is known to weaken the bond, causing its elongation.[67] Through extensive investigation of
WWW.CHEMISTRYVIEWS.ORG

2158

International Journal of Quantum Chemistry 2013, 113, 21542162

WWW.Q-CHEM.ORG

FULL PAPER

various XAH hydrogen-bond donors (X@H,C,N,Si,P), paired with a selection of several n and a few p hydrogen-bond acceptors, the authors demonstrated that for weak r-acceptors, such as CAH but also NAH and OAH under particular circumstances, the rehybridization of the X-hybrid orbital in the XAH bond is mandatory in order to have its contraction and the blue-shift of its IR stretching frequency. Results from NBO analysis show negligible occupancy differences in the antibonding orbitals of C(8)AHa and C(8)AHb bonds between the anti and gauche conformers. Reconfiguration of the C(8)-hybrid orbitals toward Ha and Hb occurs instead, moving from sp3.71 to sp3.62 for the former and to sp3.75 for the latter. The increased (decreased) s-character of these hybrid orbitals, coupled with very little contribution from p!r*CH hyperconjugation, is associated with the contraction (elongation) of the corresponding CAH bonds. It is also worth to mention that, with I as a reference, conformers III and IV show an augmented (negative) charge on C(8) and C(3), by 20.010e and 20.008e, respectively (see Figs. 1 and 5), and an augmented charge on Ha (10.015e).[68,69] The enhanced attractive force between these partial charges, which can be portrayed as the weak interaction occurring between a soft acid and a soft base as first suggested by Nishio and coworkers,[70,71] is another well-known peculiarity of the CH/p interaction. The results presented so far are consistent with Bents rule,[72] according to which an increase in the s-character of the X-hybrid orbital of the XAH bond is to be expected upon formation of a XAHY complex, as H becomes more electropositive. No significant change, on the other hand, is observed for r* orbital occupancy and hybridization of the C(6)AC(7) bond, whose elongation can be explained by considering the effect of steric repulsion between the ring and the side chain. To get a different perspective on the subject, we investigated the methane/thiophene van der Waals dimer. Its singlepoint counterpoise-corrected MP2 interaction energy is about 0.56 kcal/mol, which is expected to be almost entirely of dispersive nature, given a near-zero electrostatic contribution and the small total charge transfer between molecules. The latter was estimated to be roughly 0.0033e (from thiophene to methane) from the results of an NBO analysis. The same analysis demonstrates a small polarization contribution, associated with a slight increase of the charge of the hydrogen (HA hereafter) pointing toward thiophene (0.010e above that in the isolated methane). Similarly to conformers III and IV of 3-hexylthiophene, reorganization of the C-hybrid orbitals of methane occurs. The augmented s-character of the CAHA C-hybrid orbital (from sp3.00 in the isolated molecule to sp2.87) contraction of the corresponding is accompanied by a 0.001 A bond, while increase of p-character (to sp3.03) and bond elon ) is obtained for the remaining CAH bonds. gation (0.0005 A Although the CAHA bond in the thiophene/methane complex and CAHa moieties in the gauche conformers of 3hexylthiophene share several features, these are the result of slightly different phenomena. In the former case in fact, the CAHA bond is oriented almost orthogonally to the thiophene ring plane, while in the latter the CAHa bonds lie on planes

Figure 7. Graphical representation of the 2s extravalence NBO of HA (top) and of the CAHA antibonding NBO (bottom) of methane in the thiophene/ methane complex. Orbital occupancies are given in italics in both cases. [Color figure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

which are parallel to that of the aromatic ring. Orbitals on methane and the p system of thiophene are thus expected to overlap to different extents, resulting in a less efficient delocalization interaction in the parallel case. The second-order perturbation theory analysis of the Fock matrix in the NBO basis shows for the orthogonal orientation of the van der Waals complex a set of delocalization events resulting in electron density transfer from all rCC, rCS, and p orbitals of thiophene to the 2s orbital of HA, amounting 0.0028e. A small p!r*CH contribution to the CAHA bond is also observed. See Figure 7 for the graphical representation of these orbitals. The same kind of delocalizations toward Ha in the parallel arrangement of 3-hexylthiophene gauche conformers is expected to be significantly weaker. On the bond-length reduction of CAH proton-donors in intermolecular hydrogen bonds, observations by Schlegel and coworkers[54] regarding a steric repulsion contribution to the contraction, by Dannenberg and coworkers[52] concerning electron density transfer from HA to the corresponding rCH orbital for weak electric fields, and by Joseph and Jemmis[65] relating to an increase of electron density in the CAHA bond region, are not in conflict with our observations for the thiophene/methane complex. In the 3-hexylthiophene intramolecular hydrogen-bonding frame, on the other hand, Schlegels explanation is not applicable due to the orientation of the
International Journal of Quantum Chemistry 2013, 113, 21542162 2159

FULL PAPER

WWW.Q-CHEM.ORG

CAHa bond with respect to the aromatic ring. Occupancy of the 2s orbitals on Ha atoms do not change significantly between anti and gauche conformers, making also not applicable the theory proposed by Dannenberg and coworkers. A distinctive characteristic of intramolecular noncovalent interactions which is of course not present in the intermolecular case, is that acceptor and donor fragments only have access to a finite range of coordination schemes. Covalent spacers hinder in fact the approach of the two moieties, so that the allowed reciprocal orientations are not always optimal for an efficient interaction. In our case, C(6)H2 and C(7)H2 groups limit the conformational freedom of C(8)H2. The different configuration of these three methylene groups between anti and gauche arrangements has two major effects: the reorganization of the hyperconjugation frame among covalent bonds of the side chain and the formation of new remote interactions with the aromatic ring. The first effect described follows from a switch of the antiperiplanarity of CAC and CAH bonds of the side chain. In the anti arrangement all CAC bonds of the side chain are antiperiplanar to each other, while this is not true for gauche conformers, in which C(3)AC(6) and C(7)AC(8) bonds form a 660 angle, as for the very definition of b. As a consequence, a few rCC!r*CC and rCH!r*CH interactions are disrupted and as many rCC!r*CH and rCH!r*CC interactions are introduced. Hyperconjugation events limited to the thiophene ring or involving any methylene group beyond C(8)H2 remain mostly unchanged. We observe the reorganization of geminal and vicinal donoracceptor pairs accordingly to the new reciprocal orientation of the NBOs in the gauche conformers with respect to the anti one. However, the magnitude of these hyperconjugative interactions is roughly equivalent among all conformers. Geminal interactions are reoriented with almost no variations of strength, and vicinal rCC and rCH donors produce very similar stabilization energies whether they are paired with a r*CC or a r*CH acceptor NBO, balancing out each other over the whole molecules. The second effect, concerning the formation of new remote interactions, depends on reciprocal orientation of aromatic and aliphatic fragments in gauche conformers. These are in fact favored over the anti one by remote hyperconjugative interactions between the thiophene ring and C(7)H2 and C(8)H2 methylene groups. Indeed, a number of donoracceptor interactions in conformers III and IV do not find an equivalent counterpart in conformer I, including rCH!r* from the two C(8)AHa bonds, rCC!n* and rCH!n* from C(7)AC(8) and C(7)AH to extravalence orbitals on C(3), rCC!n* from thiophene to extravalence orbitals on C(7), C(8), and Ha, and p!r*CC to the C(8)AC(9) bonds. Although 3-alkylthiophene conformers featuring more twisted side chains (see Fig. 2) will experience an even larger number of such remote delocalization effects, this aspect alone does not guarantee an equally larger energy stabilization. The orientation of r and r* aliphatic orbitals with respect to the p system will be crucial in those cases, for the magnitude of the stabilizing interactions depend on the extent of orbitals overlap. In addition, such configurations of the side
2160 International Journal of Quantum Chemistry 2013, 113, 21542162

chain will also lead to larger repulsive (destabilizing) contributions. We can only assume the net balance between remote ring/chain hyperconjugation and steric repulsion to be the ruling factor in the energy assessment of twisted 3-alkylthiophene conformational isomers. From the results outlined earlier, we may claim that the conformations of 3-alkylthiophenes are strongly affected by a weak CH/p blue-shifting interaction[51] between CAHa moieties of the alkyl chain and the aromatic ring. The importance of a proper treatment of dispersion forces, CAHa C-hybrid orbitals reorganization, and remote hyperconjugation effects between thiophene and aliphatic fragments, was discussed. The participation of several CAC and CAH moieties of the aliphatic fragment to the latter hyperconjugation interactions, other than just CAHa bonds, make the stabilization of gauche over anti conformers a real group effort. We expect these aspects to play a crucial role also for any generic aromatic or unsaturated aliphatic system with an alkyl substituent.

Conclusions
We have used a range of quantum chemical methods to determine the stable conformations of 3-hexylthiophene and its analogues, containing both shorter and longer side chains. Our work has focused on the two torsion angles closest to the ring, as these are the ones which are more likely to play a role in crystalline oligo- and poly(alkylthiophenes). Hexylbenzene and n-hexane have also been discussed for comparison. Our study demonstrates for both alkylaromatic species the preferential stability of gauche conformers III and IV of the side chain with respect to the anti arrangement I. We have also identified a fully planar conformation II at higher energies. Intramolecular CH/p interactions can be invoked to rationalize these results. Electron correlation is shown to be of great importance in the assessment of the relative stability of these isomers. The roles of local charge-transfer, rehybridization, and repolarization of covalent bonds, steric repulsion between electron clouds, and BSSE contributions, were all considered in the rationalization of our results. We present a similar analysis for the thiophene/methane van der Waals dimer in order to elucidate the main differences between intermolecular and intramolecular CH/p phenomena. While in the noncovalent complex the CAHA bond of methane aimed toward the thiophene ring is in fact the only mediator of the interaction, folded side chains of 3-alkylthiophenes require the collective effort of several CAC and CAH bonds in order to achieve the degree of stabilization observed at highly correlated levels of theory. Although the wide majority of the crystal structures of alkylthiophene oligomers present anti arrangements of the b torsion angles, some cases[7378] in which gauche conformations are preferred have been reported. As for polymers, the only crystal structure in which a gauche conformation has been determined so far is poly(3-(S)-2-methylbutylthiophene),[79] which, however, has a branched side chain. Our study suggests that the prevalence of anti conformations for
WWW.CHEMISTRYVIEWS.ORG

WWW.Q-CHEM.ORG

FULL PAPER

the b torsion angle is largely due to packing effects, contrary to what previously assumed. The results presented can influence many aspects of our understanding of thiophene-based materials. In particular, the existence of accessible conformations other than the linear allanti one can affect estimates of the intramolecular and intermolecular components of the internal energy in crystals, side chain disorder and side chain melting,[80] dynamical behavior and charge transport phenomena,[81] thermal behavior and phase transitions,[82] as well as the choice of the most appropriate molecular modeling approach (including force field parameterization) for substituted aromatics. Work is in progress in our group to explore some of these issues through structural and modeling studies of 3-alkylthiophene oligomers and polymers. Keywords: hydrogen bonds poly(3-hexylthiophene) (P3HT) alkylaromatics ab initio calculations density functional calculations

[25]

[26]

[27]

How to cite this article: A. Baggioli, S. V. Meille, G. Raos, R. Po, M. Brinkmann, A. Famulari, Int. J. Quantum Chem. 2013, 113, 21542162. DOI: 10.1002/qua.24472

[28]

[29] [30]

[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24]

A. Heeger, J. Phys. Chem. B 2001, 105, 8475. M. L. Chabinyc, A. Salleo, Chem. Mater. 2004, 16, 4509. H. Sirringhaus, N. Tessler, R. H. Friend, Science 1998, 280, 1741. B. Gburek, V. Wagner, Org. Electron. 2010, 11, 814. S. Gunes, N. S. Sariciftci, Inorg. Chim. Acta 2008, 361, 581. C. Y. Ma, X. Gong, K. Lee, A. Heeger, Adv. Funct. Mater. 2005, 15, 1617. T. J. Prosa, M. J. Winokur, R.D. McCullough, Macromolecules 1996, 29, 3654. N. Kayunkid, S. Uttiya, M. Brinkmann, Macromolecules 2010, 43, 4961. S. Dag, L. -W. Wang, J. Phys. Chem. B 2010, 114, 5997. S. B. Darling, J. Phys. Chem. B 2008, 112, 8891. D. Dudenko, A. Kiersnowski, J. Shu, W. Pisula, D. Sebastiani, H. W. Spiess, M. R. Hansen, Angew. Chem. Int. Ed. 2012, 51, 11068. G. Natta, P. Corradini, P. Ganis, J. Polym Sci. 1962, 58, 1191. A. Maillard, A. Rochefort, Phys. Rev. B 2009, 79, 115207. J. E. Northrup, Phys. Rev. B 2007, 76, 245202. D. L. Cheung, D. P. McMahon, A. Troisi, J. Phys. Chem. B 2009, 112, 9393. C. Melis, L. Colombo, A. Mattoni, J. Phys. Chem. C 2011, 115, 576. A. Famulari, G. Raos, A. Baggioli, M. Casalegno, R. Po, S. V. Meille, J. Phys. Chem. B 2012, 116, 14504. G. Raos, A. Famulari, V. Marcon, Chem. Phys. Lett. 2003, 379, 364. G. Raos, A. Famulari, S. V. Meille, M. C. Gallazzi, G. Allegra, J. Phys. Chem. A 2004, 108, 691. M. Moreno, M. Casalegno, G. Raos, S. V. Meille, R. Po, J. Phys. Chem. B 2010, 114, 1591. K. H. DuBay, M. L. Hall, T. F. Hughes, C. Wu, D. R. Reichman, R. A. Friesner, J. Chem. Theory Comput. 2012, 8, 4556. R. Colle, G. Grosso, A. Ronzani, C. M. Zicovich-Wilson, Phys. Status Solidi B 2011, 248, 1360. F. H. Allen, Acta Crystallogr. Sect. B 2002, 58, 380. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J.

[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]

Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannen Farkas, J. B. Foresman, J. V. Ortiz, J. berg, S. Dapprich, A. D. Daniels, O. Cioslowski, D. J. Fox, Gaussian 09, Revision C.01; Gaussian, Inc.: Wallingford, CT, 2009. (a) M. W. Schmidt, K. K. Baldridge, J. A. Boatz, S. T. Elbert, M. S. Gordon, J. H. Jensen, S. Koseki, N. Matsunaga, K. A. Nguyen, S. Su, T. L. Windus, M. Dupuis, J. A. Montgomery, J. Comput. Chem. 1993, 14, 1347; (b) M. S. Gordon, M. W. Schmidt, Theory and Applications of Computational Chemistry, the first forty years; C. E. Dykstra, G. Frenking, K. S. Kim, G. E. Scuseria, Eds.; Elsevier: Amsterdam, 2005; pp. 11671189. (a) C. Mller, M. S. Plesset, Phys. Rev. 1934, 46, 618; (b) M. Head-Gordon, J. A. Pople, M. J. Frisch, Chem. Phys. Lett. 1988, 153, 503; (c) S. f, Chem. Phys. Lett. 1989, 154, 83; (d) M. J. Frisch, M. Saeb, J. Almlo Head-Gordon, J. A. Pople, Chem. Phys. Lett. 1990, 166, 275; (e) M. J. Frisch, M. Head-Gordon, J. A. Pople, Chem. Phys. Lett. 1990, 166, 281; (f ) M. Head-Gordon, T. Head-Gordon, Chem. Phys. Lett. 1994, 220, 122. (a) J. A. Pople, R. Krishnan, H. B. Schlegel, J. S. Binkley, Int. J. Quantum Chem. 1978, 14, 545; (b) R. J. Bartlett, G. D. Purvis III, Int. J. Quantum Chem. 1978, 14, 561; (c) G. D. Purvis, III, R. J. Bartlett, J. Chem. Phys. 1982, 76, 1910; (d) J. A. Pople, M. Head-Gordon, K. Raghavachari, J. Chem. Phys. 1987, 87, 5968; (e) G. E. Scuseria, C. L. Janssen, H. F. Schaefer, III, J. Chem. Phys. 1988, 89, 7382. (a) R. Krishnan, J. S. Binkley, R. Seeger, J. A. Pople, J. Chem. Phys. 1980, 72, 650; (b) A. D. McLean, G. S. Chandler, J. Chem. Phys. 1980, 72, 5639; (c) M. J. Frisch, J. A. Pople, J. S. Binkley, J. Chem. Phys. 1984, 80, 3265. T. Clark, J. Chandrasekhar, G. W. Spitznagel, P. v. R. Schleyer, J. Comput Chem. 1983, 4, 294. (a) T. H. Dunning, Jr., J. Chem. Phys. 1989, 90, 1007; (b) R. A. Kendall, T. H. Dunning, Jr., R. J. Harrison, J. Chem Phys. 1992, 96, 6796; (c) D. E. Woon, T. H. Dunning, Jr., J. Chem. Phys. 1993, 98, 1358. (a) A. D. Becke, J. Chem. Phys. 1993, 98, 5648; (b) C. Lee, W. Yang, R. G. Parr, Phys. Rev. 1988, 37, 785; (c) S. H. Vosko, L. Wilk, M. Nusair, Can. J. Phys. 1980, 58, 1200; (d) P. J. Stephens, F. J. Devlin, C. F. Chabalowski, M.J. Frisch, J. Phys. Chem. 1994, 98, 11623. J. Klime s, A. Michaelides, J. Chem. Phys. 2012, 137, 120901. J.-D. Chai, M. Head-Gordon, Phys. Chem. Chem. Phys. 2008, 10, 6615. J. -D. Chai, M. Head-Gordon, J. Chem. Phys. 2008, 128, 84106. S. Grimme, J. Chem. Phys. 2006, 124, 34108. T. Schwabe, S. Grimme, Phys. Chem. Chem. Phys. 2007, 9, 3397. T. Schwabe, S. Grimme, Phys. Chem. Chem. Phys. 2006, 8, 4398. E. D. Glendening, A. E. Reed, J. E. Carpenter, F. Weinhold, NBO 3.1; Theoretical Chemistry Institute, University of Wisconsin, Madison, 1996. A. E. Reed, L. A. Curtiss, F. Weinhold, Chem. Rev. 1988, 88, 899. A. Fujii, H. Hayashi, J.W. Park, T. Kazama, N. Mikami, S. Tsuzuki, Phys. Chem. Chem. Phys, 2011, 13, 14131. S. F. Boys, F. Bernanrdi, Mol. Phys. 1970, 19, 553. A. Famulari, E. Gianinetti, M. Raimondi, M. Sironi, Int. J. Quantum Chem. 1998, 69, 151. M. Sironi, A. Famulari, Theor. Chem. Acc. 2000, 113, 417. D. Asturiol, M. Duran, P. Salvador, J. Chem. Phys. 2008, 128, 144108. R. M. Balabin, J. Chem. Phys. 2008, 129, 164101. K. Burke, J. Chem. Phys. 2012, 136, 150901. anchez, W. Yang, Chem. Rev. 2012, 112, 289. A. J. Cohen, P. Mori-S  ka, P. Hobza, Chem. Rev. 2010, 110, 5023. K. E. Riley, M. Piton ak, P. Jurec thi, J. Chem. Phys. 2001, 114, 3949. S. Tsuzuki, H. P. Lu Y. Zhao, O. Tishchenko, D. G. Truhlar, J. Phys. Chem. B 2005, 109, 19046. P. Hobza, Z. Havlas, Chem. Rev. 2000, 100, 4253. A. Masunov, J. J. Dannenberg, R. H. Contreras, J. Phys. Chem. A 2001, 105, 4737. (a) K. Hermansson, J. Chem. Phys. 1993, 99, 861; (b) K. Hermansson, Int. J. Quantum Chem. 1993, 45, 747; (c) K. Hermansson, J. Phys. Chem. A 2002, 106, 4695. X. Li, L. Liu, H. B. Schlegel, J. Am. Chem. Soc. 2002, 124, 9639. W. Qian, S. Krimm, J. Phys. Chem. A 2002, 106, 6628. E. Cubero, M. Orozco, P. Hobza, F. Luque, J. Phys. Chem. A 1999, 103, 6394.

International Journal of Quantum Chemistry 2013, 113, 21542162

2161

FULL PAPER

WWW.Q-CHEM.ORG

[57] I. V. Alabugin, M. Manoharan, S. Peabody, F. Weinhold, J. Am. Chem. Soc. 2003, 125, 5973. [58] P. Hobza, V.  Spirko, H. L. Selzle, E. W. Schlag, J. Phys. Chem. A 1998, 102, 2501. [59] E. Mr azkov a, P. Hobza, J. Phys. Chem. A 2003, 107, 1032. [60] A. Allerhand, P. v. R. Schleyer, J. Am. Chem. Soc. 1963, 85, 1715. [61] C. A. Coulson, Research 1957, 10, 149. [62] Y. Gu, T. Kar, S. Scheiner, J. Am. Chem. Soc. 1999, 121, 9411. [63] S. Scheiner, S. J. Grabowski, T. Kar, J. Phys. Chem. A 2001, 105, 10607. [64] S. Scheiner, T. Kar, J. Phys. Chem. A 2002, 106, 1784. [65] J. Joseph, E. D. Jemmis, J. Am. Chem. Soc. 2007, 129, 4620. [66] M. Nishio, M. Hirota, Y. Umezawa, The CH/p interaction: Evidence, Nature, and Consequencies; Wiley-VCH: New York, 1998. [67] A. E. Reed, F. Weinhold, J. Chem Phys. 1985, 83, 1736. [68] U. Kock, P. L. A. Popelier, J. Phys. Chem. 1995, 99, 9747. [69] P. L. A. Popelier, J. Phys. Chem. A 1998, 102, 1873. [70] Y. Kodama, K. Nishihata, M. Nishio, N. Nakagawa, Tetrahedron Lett. 1977, 24, 2105. [71] M. Nishio, Phys. Chem. Chem. Phys. 2011, 13, 13873. [72] H. A. Bent, Chem. Rev. 1961, 61, 275. [73] G. Fuhrmann, T. Debaerdemaeker, P. B auerle, Chem. Commun. 2003, 8, 948. DOI: 10.1039/b300542a. [74] J. B. Benedict, W. Kaminsky, C. J. Tonzola, Acta Crystallogr. Sect. E 2004, 60, 530. [75] L. Liu, W. Xu, X.-W. Xiao, D.-B. Zhu, Acta Crystallogr. Sect. E 2007, 63, 3961.

[76] M. Williams-Harry, A. Bhaskar, G. Ramakrishna, T. Goodson, III, M. Imamura, A. Mawatari, K. Nakao, H. Enozawa, T. Nishinaga, M. Iyoda, J. Am. Chem. Soc. 2008, 130, 3252. [77] G. J. McEntee, P. J. Skabara, F. Vilela, S. Tierney, I. D. W. Samuel, S. Gambino, S. J. Coles, M. B. Hursthouse, R. W. Harrington, W. Clegg, Chem. Mater. 2010, 22, 3000. [78] M. Iyoda, P. Huang, T. Nishiuchi, M. Takase, T. Nishinaga, Heterocycles 2011, 82, 1143. [79] P. Arosio, A. Famulari, M. Catellani, S. Luzzati, L. Torsi, S. V. Meille, Macromolecules 2007, 40, 3. [80] M. D. Curtis, J. I. Nanos, H. Moon, W. S. Jahng, J. Am. Chem. Soc. 2007, 129, 15072. [81] D. L. Cheung, D. P. McMahon, A. Troisi, J. Am. Chem. Soc. 2009, 131, 11179. [82] (a) M. Catellani, S. Luzzati, F. Bertini, A. Bolognesi, F. Lebon, G. Longhi, S. Abbate, A. Famulari, S. V. Meille, Chem. Mater. 2002, 14, 4819; (b) M. Brinkmann, P. Rannou, Adv. Funct. Mater. 2007, 17, 101; (c) P. Arosio, M. Moreno, A. Famulari, G. Raos, M. Catellani, S. V. Meille, Chem. Mater. 2009, 21, 78; (d) Y. Yuan, J. Zhang, J. Sun, J. Hu, T. Zhang, Y. Duan, Macromol. 2011, 44, 9341; (e) M. Casalegno, A. Baggioli, A. Famulari, S. V. Meille, T. Nicolini, R. Po, G. Raos, EPJ WEB OF CONFERENCES 2012, 22, 020021

Received: 12 March 2013 Revised: 23 April 2013 Accepted: 25 April 2013 Published online on 21 May 2013

2162

International Journal of Quantum Chemistry 2013, 113, 21542162

WWW.CHEMISTRYVIEWS.ORG

You might also like