Electronic properties of poly(thiophene-3-methyl acetate) Alex L. Gomes,1,2 Jordi Casanovas,3 Oscar Bertran,4 João Sinézio de C. Campos,2 Elaine Armelin,1,5,* and Carlos Alemán1,5,* 1 Departament d’Enginyeria Química, E.T.S d’Enginyers Industrials de Barcelona, Universitat Politècnica de Catalunya, Diagonal 647, Barcelona E-08028, Spain 2 Departamento de Tecnologia de Polímeros, Faculdade de Engenharia Química, Universidade Estadual de Campinas, Avenida Albert Einstein nº 500, Barão Geraldo, CEP 13083-852, Campinas-SP, Brazil 3 Departament de Química, Escola Politècnica Superior, Universitat de Lleida, c/Jaume II n°69, Lleida E-25001, Spain 4 Departament de Física Aplicada, E.U.E.T.I.I., Universitat Politècnica de Catalunya, Pça Rei 15, Igualada 08700, Spain 5 Center for Research in Nano-Engineering, Universitat Politècnica de Catalunya, Campus Sud, Edifici C′, C/Pasqual i Vila s/n, Barcelona E-08028, Spain * elaine.armelin@upc.edu and carlos.aleman@upc.edu 1 Abstract The electronic structure of poly(thiophene-3-methyl acetate) has been investigated using UV-vis absorption spectroscopy and quantum mechanical calculations. Experimental measures in chloroform solution indicate that the -conjugation length increases with the polymer concentration, which is reflected by the red shift of the absorbance peak of the -* transition. On the other hand, the energy required for the -* transition has been found to decrease with the volatility of the solvent for concentrated polymer solutions, even though the influence of the solvent is very small for dilute solutions. Quantum mechanical calculations indicate that the interactions between the -conjugated backbone and the methyl acetate side groups are very weak. On the other hand, the lowest energy transition predicted for an infinite polymer chain that adopts the anti-gauche and all-anti conformations is 2.8 and 1.9 eV, respectively. Finally, measurements on spin-casted nanofilms reflect that the -* transition energy increases with the thickness, which has been attributed to the distortion of the molecular conformation. In spite of this, the energy gap obtained for the thinnest film (1.52 eV) is significantly smaller than that determined for dilute and concentrated chloroform solutions (2.56 and 2.09 eV, respectively). Keywords: Conducting polymers; Polythiophene; UV-vis spectroscopy; Quantum chemistry; Conjugation 2 Introduction Organic semiconducting polymers have important applications in microelectronics, electrochemical switching, photovoltaics, light-emitting diodes, field effect transistors, electrochromic and electromechanical devices, chemical and physical sensing, etc [1-2]. Among these materials, polythiophene (PTh) derivatives are the most commonly used due to their high charge carrier mobility. In particular, polymers bearing substituents into the 3position of the thiophene are especially important because in some cases they overcome the solubility and processability problems intrinsically associated to PTh. Thus, the incorporation of long alkyl side chains increases the solubility in organic solvents [3-5], while hydrophilic substituents produce PTh derivatives soluble in water or polar solvents [6]. PTh derivatives bearing carboxylic acid groups have been found to be soluble in a wide variety of organic solvents [6-13]. Within this field, the work of Heeger and co-workers on poly(thiophenesulfonate)s was a pioneering contribution to water-soluble semiconducting polymers [6]. At the end of the nineties, electroactive polymers and copolymers with acetic acid, propinionic acid and octanoic acid linked to the thiophene ring were also investigated [8,9]. More recently, we reported the synthesis, structure and electronic characterization of PTh derivatives bearing malonic acid [10,11] and acrylic acid [12,13] as substituents. Among all these functionalized, soluble and electroactive PTh derivatives, poly(thiophene3-acetic acid), or PT3AA (Scheme 1), has attracted much attention because of its versatility to prepare layer-by-layer (LBL) self-assembled systems. Specifically, in the last few years multilayered LBL thin films obtained by combining PT3AA with organic polycations [e.g. poly(sodium-2-(4-methyl-3-thiehyloxy)ethane sulfonate) and chitosan] [14,15], small 3 surfactants (e.g. dialkyldimethylammonium) [16], metallic and inorganic particles (e.g. gold, TiO2 and zeolites) [17,18], other conducting polymers (e.g. polyaniline) [19], conventional organic polymers [e.g. poly(4-vinylpyridine) and polyimide] [20] or even peptides [21] have been reported. CH2-COOH S n PT3AA Scheme 1 PT3AA presents high surface resistivity as compared with other PTh derivatives. For example, the surface resistivity reported by Zhang and Srinivasan for PT3AA-polyimide composites ranged from 106 to 108 ·sq-1 depending on the preparation of the sample [22]. Similarly, the surface resistivity of the undoped forms of both PT3AA and poly(3thiophene-acetic acid-co-3-hexylthiophene) was found to be 7·1010 ·sq-1 [23]. After doping with iodine, the surface resistivity of the latter copolymer decreases to 105 ·sq-1, while the homopolymer does not exhibit any change remaining almost identical to the value measured before doping [23]. On the other hand, the electrical conductivity reported for PT3AA-poly(ethyleneoxide) blends was also very low, i.e. around 10-6 S·cm-1 [24]. The high electrical resistivity of PT3AA was attributed to the presence of carboxylic side groups close to the -conjugated backbone [23]. Thus, this polar substituent was proposed to affect the electron transport through the -system of the polymer. Moreover, the poor doping behavior of PT3AA was explained by the dense packing of the chains with 4 CH2COOH pendant groups that interact through hydrogen bonds, precluding the diffusion of the dopant molecules through the polymeric matrix [23]. Furthermore, the electronic structure of PT3AA was examined by studying the pH and temperature dependence of the UV-Vis absorption spectrum [8]. An abrupt reversible increase of max was found when the pH varies from 5 to 6, which was attributed to a variation in the effective electronic length. Thus, the electrostatic repulsions between the dissociated carboxylic acid side groups were suggested to induce conformational changes in the polymer chains from the aggregated state, which is stabilized by hydrogen bonds, to the extended state, giving rise to an increased effective conjugation length [25]. On the other hand, the max after the pH transition (pH= 6.3) decreased with the increase in temperature, whereas that before the transition (pH= 5.2) increases slightly with increasing temperature. These thermal effects suggested that at the former pH the extended polymer chains evolve towards a disordered state as the temperature increases [8]. In opposition, at pH= 5.2 the intramolecular hydrogen bonds break upon the increase of temperature, PT3AA chains becoming more conjugated. In this work we examine the electronic structure of a PT3AA derivative in which both the ability to form hydrogen bonds and the pH dependence have been eliminated. More specifically, the electronic properties of poly(thiophene-3-methyl acetate), or PT3MA (Scheme 2), have been investigated in detail by combining both absorption spectroscopy and quantum mechanical calculations. Specifically, experimental investigations were performed considering dilute and concentrated PT3MA solutions in different organic solvents as well as nanofilms prepared by spin-coating, while theoretical studies were 5 carried out using Density Functional Theory (DFT) calculations in both gas-phase and solution environments. Methods Materials and Polymerization. 3-Thiophene acetic acid (T3AA) and anhydrous ferric chloride (FeCl3) were purchased from Sigma-Aldrich Química S.A. and were employed without further purification. All solvents were purchased from Panreac Química S.A. with ACS grade. 3-Thiophene methyl acetate (T3MA) was prepared and subsequently polymerized by oxidative coupling following the procedure described by Kim et al. [8], which is summarized in Scheme 2. The purified T3MA monomer was obtained with 82% of yield, while the yield of PT3MA after remove the residual oxidant and oligomers was 25%. Unfortunately, changes in the monomer:oxidant (T3MA:FeCl3) ratio did not improve the low yield of the oxidative coupling polymerization. The characterization of PT3MA using FTIR and 1H-NMR was previously reported [8], our results being in complete agreement with them. Table 1 describes the solubility of the resulting PT3MA in different organic and polar solvents at room (25 ºC) and high temperatures (60 ºC). 6 CH2-COOH CH2-COOCH3 FeCl 3 CH3OH S T3AA H2SO4 / 90ºC CH2-COOCH3 CHCl 3 / 0ºC S T3MA S n PT3MA Scheme 2 Spin-Casting. PT3MA solutions were prepared by dissolving the polymer in chloroform, concentrations ranging from 0.01 mg/mL to 5.00 mg/mL being considered. Nanofilms were obtained by spin-coating (WS-400B-NPP Laurell Technology Co.) about 18-20 drops of solution deposited on indium-tin oxide (ITO) glass slides, previously cleaned with ethanol, at 8000 rpm for 60 s with the acceleration set to be of 15 s. It is well known that the film thickness depends on the concentration and the speed of spinning. Previous studies on other polymeric systems reported film thickness ranging from 10 to 80 nm for concentrations similar to those used in this work [26,27]. Due to the poor mechanical properties of P3TMA, the thickness of the nanofilms was estimated using a calibration curve for 50:50 mixtures of this PTh derivative and poly(tetramethylene succinate), the latter being added to impart mechanical consistency. The thickness of the nanofilms obtained using spinning speeds ranging from 1500 to 12000 rpm for 60 s were determined by scratch AFM. According to the curve thickness versus speed, the thickness of the nanofilms obtained at 8000 rpm for 60 s was comprised between 20 and 30 nm. Absorption Spectroscopy. The absorption spectra were obtained with a Shimadzu 3600 spectrophotometer equipped with a tungsten halogen visible source, a deuterium arc UV source, a photomultiplier tube UV-vis detector, and a InGaAs photodiode and cooled PbS photocell NIR detectors. Spectra were recorded in the absorbance mode using the 7 integrating sphere accessory (model ISR-3100), the wavelength range being 200-1200 nm. The interior of the integrating sphere is coated with a highly diffuse BaO reflectance standard. Spectra were obtained for both polymer solutions and nanofilms. Single-scan spectra were obtained at a scan speed of 60 nm/min using the UVProbe 2.31 software. Quantum Mechanical Calculations. Complete and partial geometry optimization of oligomers containing n repeating units (n-T3MA) with n ranging from 1 to 10 were performed using Density Functional Theory (DFT) calculations at the B3LYP [28,29] level combined with the 6-31+G(d,p) basis set [30]. Calculations on n-T3MA were performed considering different arrangements for the methyl acetate side groups, which were selected from previous studies on polythiophene derivatives bearing carboxylic acid substituents [10,31]. The all-anti conformation, in which all the inter-ring dihedral angles (; defined by the S-C-C-S sequence) adopt a value of 180.0º, was used as starting structure in all cases. The Koopman’s theorem [32], which according to Janak’s theorem can be applied to DFT calculations [33], was used to estimate the ionization potentials (IPs) and electron affinities (EAs). Accordingly, the IP and the EA were taken as the negative of the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) energies, respectively, i.e. IP= -EHOMO and EA= -ELUMO. The IP and EA indicate if a given acceptor (p-type dopant) and donor (n-type dopant) is capable of ionizing, at least partially, the molecules of the compound, respectively. The -* lowest transition energy (Eg) was approximated as the difference between the HOMO and LUMO energies, i.e. Eg= EHOMO ELUMO. Levy and Nagy evidenced that g can be rightly approximated using this procedure in DFT calculations [34]. 8 Results and Discussion Dilute Solutions (0.01 mg/mL to 0.05 mg/mL). Dried PT3MA (5 mg) was used to prepare a chloroform solution with concentration 2.6·10-3 mol·L-1. In order to achieve the best solubility conditions, chloroform anhydrous than contains 0.5-1.0% ethanol was used as solvent. This is because some residual oxidant particles coming from synthesis were found to precipitate when no stabilizer was contained in the solvent, while such residuals disappear in presence of a small fraction of ethanol (i.e. hereafter, chloroform refers to chloroform stabilized with ethanol). Exact volumes (1-5 mL) of such solution were transferred to 10 mL volumetric flasks and diluted with chloroform to prepare dilute solutions with various known concentrations. Dilute solutions show an absorbance peak in the UV-vis range at max= 400 nm (Figure 1), which correspond to the -* transition of the thiophene ring of PT3MA [35,36]. Furthermore, the absorbance of the solutions scaled with concentrations linearly. These UV-vis absorbance (A) data (peak at 400 nm) were used to develop the Beer’s law calibration curve (Figure 2), for which each measurement was repeated three times and the average values were recorded. The linear portion was plotted and the molar absorptivity () was obtained from the slop of the straight line. The Beer’s law plot for PT3MA was expressed as the equation: A= 6193.7c + 0.1038. Accordingly, considering A= ·b·c, where b is the path length of the sample (b= 1 cm for a standard cuvette), the molar absorptivity of PT3MA (max= 400 nm) in chloroform is = 6193.7 mol·g-1·cm-1, while 0.1038 is typically attributed to systematic experimental errors (cuvettes, air, solvent, etc). 9 Concentrated Solutions (0.03 mg/mL to 2.5 mg/mL). The evidence of chemical reaction between the dopant and the polymer comes from the changes observed in the UV-vis spectra. Figure 3 compares the UV-vis spectra of PT3MA in chloroform solution with concentrations ranging from 0.03 mg/mL to 2.5 mg/mL. The -* transition of the thiophene ring shows a red shift with respect to the spectra recorded for dilute solutions (Figure 1) indicating that the -conjugation length increases with the concentration of polymer (i.e. the energy required for the -* transition decreases). Thus, the max increases from 400 nm (dilute solutions) to 496 nm (highest concentration in Figure 3). Furthermore, the band observed at 744 nm, which is not present in dilute solutions, can be attributed to the formation of polarons and bipolarons in polymer chains evidencing the influence of the dopant agent (i.e. FeCl3) [37-39]. Effect of the solvent. Changing from chloroform to more polar and less volatile solvents in dilute solutions, has a negligible effect in the -* absorption band of PT3MA. Thus, the max is 400, 407 and 400 nm for a 0.01 mg/mL solution in chloroform, chlorobenzene and cyclohexanone, respectively (spectra not shown). The optical -* lowest transition energy (Eg) or band gap energy was determined from the onset wavelength (onset) of the UV-vis spectra recorded in these environments, i.e. Eg= 1240/onset eV [40], the resulting values being 2.56, 2.48 and 2.48 eV (Table 2), respectively. These results indicate that the influence of the solvent in the -electron delocalization is very small (< 0.1 eV) at the molecular level. Thus, although the dielectric constant of these solvents ranged from 3.7 (chloroform) to 15.6 (cyclohexanone), the nature of the non-specific solute···solvent 10 interactions (i.e. specific hydrogen bonds between these solvents and P3TMA are not possible) is relatively similar in all cases minimizing the effects. In opposition, the solvent plays a very important role in the electronic structure of concentrated solutions, which is reflected in Figure 4 for the 5.0 mg/mL solutions in the same three solvents. As it can be seen, the solvent affects the peak position of the -* transition, the max obtained for the -* transition in chloroform, cyclohexanone and chlorobenzene being 508, 522 and 537 nm, respectively. Thus, red shift occurs when the volatility of the solvent decreases, which results in a reduction of the energy required for the -* transition (Table 2). Thus, in chloroform (vapor pressure 160 mmHg at 20º) the value of Eg is 2.09 eV, decreasing to 1.91 and 1.89 eV in cyclohexanone (vapor pressure 3.4 mmHg at 20ºC) and chlorobenzene (vapor pressure 11.8 mmHg at 25ºC), respectively. These results indicate that the intermolecular interactions among the polymer molecules are affected by the molecular mobility of the different solvents in the medium, which in turn vary with their volatility. Another interesting feature is that the oxidation band observed at 744 nm in chloroform solution is not detected in chorobenzene and cyclohexanone confirming that this band corresponds to the FeCl3 complex. Effect produced by the addition of other dopants. Polystyrensulfonic acid (PSSA) and dodecylbenzenesulphonic acid (DBSA) were dropped in dilute (0.001, 0.005 and 0.01 mg/mL) and concentrated (5.0 mg/mL) chloroform solutions of PT3MA. The spectra absorption recorded for PT3MA doped with PSSA and DBSA were very similar to those obtained with the FeCl3 coming from polymerization (data not shown). For example, the max and Eg obtained for PSSA in chloroform solution (0.01 mg/mL) were 403 nm and 2.53 11 eV, respectively, whereas they were 402 nm and 2.54 eV for DBSA. The poor doping behavior of PT3MA is fully consistent with that found for PT3AA [23], which was attributed to the short distance between the conjugated chain and the quenching carboxylate group. Modeling of the electronic properties in the gas-phase: influence of the molecular conformation. Quantum mechanical calculations on 4-T3MA were performed to determine the conformational preferences of the methyl acetate side group. Thus, different conformations, which only differ in the disposition of the methyl acetate side groups, were constructed considering an anti arrangement for all the inter-ring dihedral angles (= 180º). All these structures were used as starting points for geometry optimizations at the B3LYP/6-31+G(d,p) level. The two arrangements of lower energy, which are almost isoenergetic, are depicted in Scheme 3. As it can be seen, in the lowest energy one (I) the side groups of all repeating units retain the same relative orientation, while in the second one (II) the side groups show an alternated disposition, the latter being only 0.1 kcal/mol less stable than the former. In both structures, all the inter-ring dihedral angles evolved toward an anti-gauche conformation with ranging from 130º to 132º. H3COOC H3COOC S H3COOC S S S S H3COOC H3COOC S S H3COOC S H3COOC I H3COOC II 12 Scheme 3 In order to evaluate the influence of the side groups in electronic properties (IP, EA and Eg) of PT3MA, quantum mechanical calculations on n-T3MA oligomers with n ranging from 2 to 10 were performed considering the arrangements I and II for each value of n. The most favored arrangement was I in all cases, the maximum destabilization obtained for II being only 0.3 kcal/mol. Furthermore, the backbone adopted an anti-gauche conformation in all cases with varying between 128º and 140º. These results clearly indicate that, independently of the number of repeating units, the interaction between the methylacetate side groups and the polythiophene backbone is very weak. On the other hand, the all-anti arrangement has been also considered for the modeling of the electronic properties. Thus, each n-T3MA oligomer was optimized considering initially the side groups as in arrangement I and fixing the inter-ring dihedral angles at 180º (restricted geometry optimizations). As it was expected, the all-anti conformation was less stable than the antigauche one, this effect increasing with n, i.e. the relative energy increases from 1.9 to 12.8 kcal/mol when n grows from 2 to 10. Figure 5 displays the variation of the IP, EA and Eg calculated in the gas-phase with the inverse of the number of repeating units for the anti-gauche (I and II) and all-anti arrangements. Linear regression analyses, which are also displayed in Figure 5, allowed extrapolate the IP, EA and Eg values for an infinite chain of PT3MA. As it was expected, the IP predicted for two anti-gauche arrangements, which only differ in the relative disposition of the side groups, were similar due to the relatively poor influence of the backbone – side chains interactions (Figure 5a), i.e. 5.25 and 5.37 eV, respectively. 13 However, oxidation of PT3MA becomes an easier process when the thiophene rings are obligated to adopt a planar disposition, the IP for the all-anti conformation being 4.74 eV. The latter value is almost identical to those calculated for the all-anti conformation of unsubstituted polythiophene (4.75 eV) [41] and poly(2-thiophen-3-yl-malonic acid dimethyl ester) (PT3MDE, 4.71 eV) [10], a similar polythiophene derivative that contains two carboxilate groups per repeating unit (Scheme 4), using similar theoretical methods. CH-(COOCH3)2 S n PT3MDE Scheme 4 Figure 5b, which displays the variation of the EA against 1/n for the three investigated arrangements, indicates that energy of the LUMO is around 0.4 eV larger for the all-anti than for I and II. The influence of the backbone conformation on the IP and EA has a significant impact on the Eg, the value predicted for an infinite of PT3MA being 2.78, 2.93 and 1.89 for the I, II and all-anti conformations, respectively (Figure 5c). As it was expected, the g value decreases when the planarity of the backbone increases, which must be attributed to the enhancement of the -conjugation produced when the inter-ring dihedral angles evolve towards 180º. The Eg value determined experimentally for a dilute chloroform solution, which is the solvent with the lowest polarity among those used in this work, is 2.56 eV indicating an overestimation of 0.22 and 0.37 eV for arrangements I and 14 II, respectively. In contrast, the Eg of the all-anti conformation is underestimated by 0.67 eV. Accordingly, the backbone of the PT3MA molecules in dilute chloroform solution adopts an intermediate conformation between the anti-gauche and the all-anti, even though it is closer to the former than to latter. In contrast, the Eg determined by UV-vis spectroscopy for a concentrated chloroform solution (5 mg/mL) is closer to the Eg predicted for the all-anti conformation than for the anti-gauche one, i.e. the difference is 0.20 and 0.69 eV, respectively. Accordingly, PT3MA chains tend to adopt planar arrangements in concentrated solutions, which is fully consistent with the red-shift observed for max (i.e. the -conjugation length increases in polymer chains). Moreover, this planarity increases when the volatility of the solvent decreases, the Eg predicted for a concentrated chlorobenzene solution being identical to that calculated for the ideal all-anti conformation. Figure 6 shows the contour plots HOMO and LUMO computed for n-T3MA with n= 3 and 10. As it can be seen, the HOMO presents a major bonding character delocalized along the backbone. For the shortest oligomer such delocalization is uniform, whereas for the largest one this effect involves the eight central repeating units. Furthermore, in both cases the HOMO resides in the conjugated backbone, no participation of the carbonyl side groups being detected. This is fully consistent with the weak interaction found between the methylacetate side groups and the polythiophene backbone. Similar distributions are predicted for the LUMO of 3-T3MA and 10-T3MA, i.e. complete and partial delocalization, respectively, which mainly resides in the inter-ring bonds. Nanofilms. Figure 7 shows the absorption spectra of ultra-thin films of PT3MA. Films were prepared by spin-casting from chloroform solutions using the same speed and time of 15 spinning in all cases, their thickness being varied by changing the concentration of polymer in the solution (see Methods section). Although the absorption profiles recorded in the solid state looks similar to those obtained in solution, the thickness of the films affects both the peak position and the absorbance. A blue shift is detected with increasing the concentration of PT3MA in the solution used for the spin-casting, the max of the films cast using 0.01, 1.00 and 5.00 mg/mL in chloroform being 482, 433 and 419 nm, respectively. Furthermore, the Eg gap increases with the concentration, growing from 1.52 to 2.17 eV when the concentration increases from 0.01 to 5.00 mg/mL (Table 2). These results indicate that the more effective packing of the -conjugated backbone is obtained for the nanofilm samples prepared using the more dilute solutions. Furthermore, the very low Eg values determined for nanofilms produced using dilute solutions indicate that intermolecular -conjugation effects are very important in the solid state, as was recently predicted from sophisticated ab initio calculations [42]. Thus, the Eg measured for the nanofilms casted from the 0.01 mg/mL solution is 0.37 eV smaller than that predicted for an isolated polymer chain arranged in an ideal planar conformation. On the other hand, it is worth noting that the Eg measured for the film cast from the 5.00 mg/mL solution (2.17 eV) is higher than those determined in the solution state for the same concentration, independently of the solvent (1.89-2.09 eV, in Table 2). Thus, the reduction of the gap produced by -stacking interactions is not enough to compensate the negative conformational effects in the molecular -conjugation length. Accordingly, spincasting of concentrated solutions produce distortions in the shape of the polymer chains, which are not able to readapt their conformations due to the packing of the neighboring molecules. 16 Conclusions The electronic properties of PT3MA have been examined in different environments using UV-vis absorption spectroscopy and quantum mechanical calculations. The absorption of the thiophene -* transition is not affected by the chemical nature of the solvent in dilute polymer solutions. In contrast, this transition shows a red-shift for concentrated solutions, which produces a reduction of the Eg. On the other hand, quantum mechanical calculations indicate that the anti-gauche arrangement is more stable in the gasphase than the all-anti by 3 kcal/mol per repeating unit. Comparison of the Eg values calculated for such two conformations with those determined experimentally allows conclude that in dilute solution PT3MA chains are closer to the anti-gauche than to the allanti, whereas in concentrated solutions polymer chains prefer the latter. The Eg values measured for nanofilms are significantly smaller than that predicted for an ideally planar molecule, the more effective packing of the chains being obtained for the films prepared using the more dilute solutions. Acknowledgements This work has been supported by MICINN and FEDER (Grant MAT2009-09138), by the Generalitat de Catalunya (research group 2009 SGR 925 and XRQTC). Computer resources were generously provided by the “Centre de Supercomputació de Catalunya” (CESCA). A.G. acknowledges financial support from the Euro Brasilian Windows agency (Grant No. 41309-EM-1-2008-PT-ERAMUNDUS-ECW-L16) for his 6-month stay at the 17 UPC. Support for the research of C.A. was received through the prize “ICREA Academia” for excellence in research funded by the Generalitat de Catalunya. References 1. Skotheim TA, Reynolds JR (2007). Handbook of Conducting Polymers, 3rd ed. CRC Press, Taylor and Francis Group, Boca Raton, FL. 2. Wallace GG, Spinks GM, Kane-Maguire LAP, Teasdale PR (2009). Conductive Electroactive Polymers, 3rd ed. CRC Press, Taylor and Francis Group, Boca Raton, FL. 3. Chen TA, Wu X, Rieke RD (1995) J Am Chem Soc 117:233-244. 4. Yang C, Orfino FP, Holdcroft S (1996) Macromolecules 29:6510-6517. 5. McCullough RD (1998) Adv Mater 10:93-116. 6. Patil OA, Ikenoue Y, Wudl F, Heeger AJ (1987) J Am Chem Soc 109:1858-1859. 7. Chayer M, Faïd K, Leclerc M (1997) Chem Mater 9:2902-2905. 8. Kim B, Chen L, Gong J, Osada Y (1999) Macromolecules 32:3964-3969. 9. Visy K, Kanlare J, Kriván E (2000) Electrochim Acta 45: 3851-3864. 10. Armelin E, Bertran O, Estrany F, Salvatella R, Alemán C (2009) Eur Polym J 45: 2211-2221. 11. Bertran O, Armelin E, Estrany F, Gomes A, Torras J, Alemán C (2010) J Phys Chem B 114:6281-6290. 12. Bertran O, Pfeiffer P, Torras J, Armelin E, Estrany F, Alemán C (2007) Polymer 48: 6955-6964. 18 13. Bertran O, Armelin E, Torras J, Estrany F, Codina M, Alemán C (2008) Polymer 49:1972-1980. 14. Takeoka Y, Iguchi M, Rikukawa M, Sanui K (2005) Synth Met 154:109-112. 15. Constantine CA, Mello SV, Dupont A, Cao X, Santos Jr D, Oliveira Jr ON, Strixino FT, Pereira EC, Cheng TC, Defrank JJ, Leblanc RM (2003) J Am Chem Soc 125: 1805-1809. 16. Yoon YS, Park K-H, Lee J-C (2009) Macromol Chem Phys 210:1510-1518. 17. Kim Y-G, Kim J, Ahn H, Kang B, Sung C, Samuelson LA, Kumar J (2003) J Macromol Sci Pure Appl Chem A40:1307-1333. 18. Thuwachaowsoan K, Chotpattananont D, Sirivat A, Rujiravanit R, Shwank JW (2007) Mater Sci Eng B 140:23-30. 19. Zhang Z, Wang L, Deng J, Wan M (2008) React Funct Polym 68:1081-1087. 20. Jiang Y, Wu P (2008) Appl Spectrosc 62:207-. 21. Nilsson KP, Rydberg J, Baltzer L, Inganäs O (2004) Proc Natl Acad Sci 101:1119711202. 22. Zang F, Srinivivasan MP (2005) Thin Solid Films 479:95-102. 23. Zang F, Srinivivasan MP (2008) Macromol Chem Phys 112, 223-224. 24. Hsieh KH, Ho KS, Wang YZ, Ko SD, Fu SC (2001) Synth Met 123:217-224. 25. Faid K, Leclerc M (1998) J Am Chem Soc 120:5274-5278. 26. Watanabe H, Kunitake T (2007) Adv Mater 19:909-912. 27. Jaczewska J, Budkowski A, Bernasik A, Moons E, Rysz J (2008) Macromolecules 41:4802-4810. 28. Becke AD (1993) J Chem Phys 98:1372-1377. 29. Lee C, Yang W, Parr RG (1988) Phys Rev B 37:785-789. 19 30. Frisch MJ, Pople JA, Binkley JS (1984) J Chem Phys 80:3265-3269. 31. Casanovas J, Zanuy D, Alemán C (2005) Polymer 46:9452-9460. 32. Koopmans T (1934) Physica 1:104-113. 33. Janak JF (1978) Phys Rev B 18:7165-7168. 34. Levy M, Nagy A (1999) Phys Rev A 59:1687-1689. 35. de Souza JM, Pereira EC (2001) Synth Met 118:167-170. 36. Jiang Y, Shen Y, Wu P (2008) J Col Interf Sci 319:398-405 37. Aaron JJ, Fall M (2000) Spectrochim Acta Part A 56:1391-1397. 38. Calado HDR, Matencio CL, Donnci LA, Cury LA, Rieumont J, Pernaut JM (2008) Synth Met 158:1037-1042 39. Alves MRA, Calado HDR, Donnici CL, Matencio T (2010) Synth Met 160:22-27. 40. Meng H, Zheng J, Lovinger AJ, Wang B-C, Van Patten PG, Bao Z (2003) Chem Mater 15:1778-1787. 41. Casanovas JC, Alemán C (2007) J Phys Chem C 111:4823-4830. 42. Rodríguez-Ropero F, Casanovas J, Alemán C (2008) J Comput Chem 29:69-, J. Comput. Chem. 2008, 29, 69-78. 20 Table 1. Basic solubility data of PT3MA in different solvents at room and high temperature. Solvent 25 ºC 60 ºC Water (pH= 7.0) NaOH aq. Acetone Δ Δ CH2Cl2 CHCl3 THF DMSO Δ DMF Δ TFA Xylene Methanol Ethanol Toluene Hexane Acetonitrile Diethyl ether - Ciclohexanone Chlorobenzene = soluble; Δ = partialy soluble; = insoluble 21 Table 2. Values of max, onset and Eg determined experimentally and theoretically for PT3MA in different conditions. max (nm) onset (nm) Eg (eV) Dilute chloroform sol. (0.01 mg/mL) 400 485 2.56 Dilute chlorobenzene sol. (0.01 mg/mL) 407 500 2.48 Dilute cyclohexanone sol. (0.01 mg/mL) 400 499 2.48 Concentrated chloroform sol. (5.00 mg/mL) 508 592 2.09 Concentrated chlorobenzene sol. (5.00 mg/mL) 537 657 1.89 Concentrated cyclohexanone sol. (5.00 mg/mL) 522 647 1.92 DFT calculations, gas-phase, arrangement I - - 2.78 DFT calculations, gas-phase arrangement II - - 2.93 DFT calculations, gas-phase, all-anti - - 1.89 Nanofilm – spin casting (0.01 mg/mL) 482 814 1.52 Nanofilm – spin casting (1.00 mg/mL) 433 626 1.98 Nanofilm – spin casting (5.00 mg/mL) 419 572 2.17 Conditions 22 Captions to Figures Figure 1. UV-vis absorption spectra of PT3MA in chloroform solutions with different concentrations used to develop the calibration curve. Figure 2. Calibration curve used to develop the Beer’s law slope for PT3MA in dilute chloroform solution. Figure 3. UV-vis absorption spectra of concentrated PT3MA solutions in chloroform. Figure 4. UV-vis absorption spectra of concentrated PT3MA solutions (5.0 mg/mL) in chloroform, cyclohexanone and chlorobenzene. Figure 5. Variation of the calculated IP (a), EA (b) and Eg (c) against 1/n, where n is the number of repeat units for PT3MA arranged in the following conformations: antigauche with I (black line, black triangles), anti-gauche with II (grey line, grey squares) and all-anti (dashed line, empty circles). The lines correspond to the linear regressions used to obtain the values of these electronic properties for infinite chain systems. Figure 6. Comparison of the contour plots of the HOMO and LUMO for n-T3MA with (a) n= 3 and (b) n= 10. Figure 7. UV-vis absorption spectra of PT3MA nanofilms obtained by the solution spin-casting process using polymer solutions with concentrations ranging from 0.01 to 5.00 mg/mL. The values of max are indicated for each concentration. 23 2.0 Absorbance 1.8 2.64 x 10-4 mol/L 2.11 x 10-4 mol/L -4 1.58 x 10 mol/L 1.06 x 10-4 mol/L -4 0.53 x 10 mol/L 1.5 1.3 1.0 0.8 0.5 0.3 0.0 297 347 397 447 Wavelength (nm) Figure 1 24 497 547 1.9 Absorbance 1.7 y = 6193.7x + 0.1038 2 R = 0.9921 1.5 1.3 1.1 0.9 0.7 0.5 0.3 0.00004 0.0001 0.00016 0.00022 Concentration (mol/L) Figure 2 25 0.00028 Absorbance 3.0 2.50 mg/mL 1.25 mg/mL 0.625 mg/mL 0.3125 mg/mL 0.03125 mg/mL 2.0 1.0 0.0 437 562 687 812 Wavelength (nm) Figure 3 26 937 1062 6.0 Absorbance 5.0 chlorobenzene 4.0 cyclohexanone 3.0 chloroform 2.0 1.0 0.0 437 562 687 812 Wavelength (nm) Figure 4 27 937 1062 (a) 6.0 IP (eV) 5.8 Arrangement II IP = 1.40(1/n ) + 5.37 R2 = 0.9800 Arrangement I IP = 1.62(1/n ) + 5.25 R2 = 0.997 5.6 5.4 All-anti IP = 2.08(1/n ) + 4.74 2 R = 0.999 5.2 5.0 4.8 4.6 0.0 (b) 0.1 0.2 0.3 0.4 0.5 0.6 0.5 0.6 0.5 0.6 1/n 2.9 EA (eV) 2.7 All-anti EA = -2.40(1/n ) + 2.85 R2 = 0.999 2.5 2.3 2.1 1.9 1.7 1.5 Arrangement I EA = -2.13(1/n ) + 2.47 R2 = 0.995 Arrangement II EA = -2.10(1/n ) + 2.44 2 R = 0.993 1.3 0.0 g (eV) (c) 0.1 0.2 0.3 0.4 1/n Arrangement II eg = 3.51(1/n ) + 2.93 R2 = 0.990 4.6 4.2 3.8 Arrangement I eg = 3.76(1/n ) + 2.78 R2 = 0.996 3.4 3.0 All-anti eg = 4.48(1/n ) + 1.89 2 R = 0.999 2.6 2.2 1.8 0.0 0.1 0.2 0.3 1/n Figure 5 28 0.4 (a) LUMO HOMO (b) HOMO LUMO Figure 6 29 Absorbance (%) 0.30 0.25 5 mg/mL 1mg/mL 0.5 mg/mL 0.1 mg/mL 0.01 mg/mL 419 nm 0.20 0.15 433 nm 432 nm 0.10 471 nm 0.05 0.00 300 482 nm 350 400 450 500 Wavelength (nm) Figure 7 30 550 600 Text for the Table of Contents The -conjugation length and the energy required for the -* transition of poly(thiophene3-methyl acetate) have been examined in different environments (i.e. gas-phase, dilute and concentrated solutions considering solvents with different polarity and volatility, and spincasted nanofilms) using a combination of UV-vis spectroscopy and quantum mechanical calculations. Graphic for the Abstract 31