π-Conjugated Fluorescent Azomethine Copolymers: Opto-Electronic

Jan 24, 2012 - Centre-ville, Montréal, Québec, H3C 3J7, Canada ... For a more comprehensive list of citations to this article, users are encouraged ...
0 downloads 0 Views 379KB Size
Article pubs.acs.org/Macromolecules

π-Conjugated Fluorescent Azomethine Copolymers: Opto-Electronic, Halochromic, and Doping Properties Satyananda Barik, Thomas Bletzacker, and W. G. Skene* Laboratoire de caractérisation photophysique des matériaux conjugués, Département de Chimie, Pavillon JA Bombardier, Université de Montréal, CP 6128, succ. Centre-ville, Montréal, Québec, H3C 3J7, Canada S Supporting Information *

ABSTRACT: A series of new conjugated copolyazomethines consisting of alternating fluorene and thiophenes were prepared. Different thiophenes were incorporated for examining their effect on the polymer’s opto-electronic properties. Blue, red, and green emissions were possible contingent on the thiophene comonomer and the placement of the azomethine nitrogen. Most notably, the polyazomethines were fluorescent both in solution and thin films with measured absolute fluorescent quantum yields (Φfl) ca. 10%. Their fluorescence could be enhanced to near unity at 77 K. High Φfl were also measured with the addition of trifluoroacetic acid (TFA). The polyazomethines’ color could also be reversible altered with protonation/neutralization with TFA/ triethylamine. Similar to the halochromic effect, the polyazomethines could be reversibly oxidized/neutralized with FeCl3/hydrazine, resulting in reversible color changes between red and blue. The oxidation potentials and the reversibility of the anodic process were contingent on the type of thiophene incorporated into the copolymer. The stability of the electrochemically produced radical cation and the azomethine’s resistance toward hydrolysis were also investigated with a model thiophene−fluorene−thiophene bisazomethine.



INTRODUCTION Conjugated materials have drawn much attention because of their opto-electronic properties that are suitable for use in plastic electronics including organic photovoltaics, organic field effect transistors, and organic light emitting diodes (OLEDs), to name but a few.1−10 Polyfluorene derivatives are particularity interesting. This is in part owing to their intrinsic high degree of fluorescence making them ideal for OLED applications.11,12 As a result, much effort has been dedicated to preparing new fluorene materials with enhanced properties including high brightness13,14 and pure color of emission.15−17 The preparation of polyfluorenes is possible by well established protocols, including Gilch,18,19 Horner-Emmons,20 Suzuki,21 Yamamoto,22 and Stille protocols.23 While high conversions are possible with these methods, they require stringent reaction conditions. Additionally, pristine polymers with high color purity are obtained only after substantial product purification. Polyazomethines are interesting alternatives to conventional coupling protocols. This is in part owing to their straightforward preparation that does not require stringent reaction conditions or coreagents. Water is also the unique reaction byproduct. Therefore, extensive purification for removing side products and coreagents is not required. In fact, pristine polymers are obtained upon solvent evaporation or by simple precipitation. © 2012 American Chemical Society

While azomethines are isoelectronic to their vinyl analogues and their preparation is synthetically advantageous relative to their carbon counterparts, they have not been pursued as functional materials for use in plastic electronics. This is partly because they are understood to be both hydrolytically and oxidatively unstable. More importantly, all previously investigated conjugated polyazomethines were nonfluorescent, despite incorporating intrinsically fluorescent monomers.24 Accurate opto-electronic property evaluation was further problematic because of the limited solubility of previously reported polyazomethines in common organic solvents.25−30 We recently addressed the optical property limitations and solubility issues encountered with previously investigated polyazomethines.31−34 This was in part owing to the use of C10 groups. This led to highly soluble polyazomethines whose molecular weight could accurately be measured using standard methods. While hydrolytically and oxidatively robust polyazomethines were possible,31,33,35,36 they were not fluorescent. The first example of a fluorescent homopolyazomethine (PFlco-Fl) was however possible by using the 9,9-dihexyl-2,7diaminofluorene monomer 1.37 In fact, the absolute Φfl (0.4 in thin films) of the all-fluorene polyazomethine was comparable to polyfluorene prepared by common aryl coupling protocols.38 Received: November 2, 2011 Revised: January 6, 2012 Published: January 24, 2012 1165

dx.doi.org/10.1021/ma2024304 | Macromolecules 2012, 45, 1165−1173

Macromolecules

Article

Chart 1. Polyazomethines and Representative Model Azomethines Prepared and Investigated

its maximum absorption in spectroscopic grade 2-methylcyclohexane solutions at room temperature relative to itself at 77 K under the same conditions. The slit widths and detector parameters were optimized for the fluorescence at 77 K. Absolute quantum yields were done at room temperature using an integrating sphere that was calibrated with fluorescein in ethanol with a calibrated excitation source. Quantum yields of the polymers in thin film were done at room temperature with the integrating sphere using a front face excitation/emission geometry. Thin films were prepared by spin coating THF solutions of the given polymer onto thin glass microscope slides. Polymer doping was done in dichloromethane (10−5 M) with stock solutions of FeCl3 and hydrazine that were sequentially added (5 μL). The resulting absorbance spectra after the addition of the dopant and reductant were subsequently recorded. Similarly, stock solutions of trifluoroacetic acid (TFA) and triethylamine (TEA) were sequentially added (5 μL) to the given polymer solution and the resulting absorbance spectra were recorded. The oxidation stoichiometry with FeCl3 was determined via the Job method using stock solutions of known mole fractions (1.0 mM) of the polymer and ferric chloride. Electrochemical Measurements. Cyclic voltammetry measurements were performed on a Bio Analytical Systems EC Epsilon potentiostat. Compounds were dissolved in anhydrous and deaerated dichloromethane at 10−4 M with 0.3 M TBAF. A platinum electrode and a saturated Ag/AgCl electrode were used as auxiliary and reference electrodes, respectively. Molecular Weight Measurements. GPC analyses were done with a Breeze system from Waters equipped with a 717 plus autosampler, a 1525 Binary HPLC pump and a 2410 refractive index detector. Three Styragel columns HR3, HR4, and HR6 (7.8 × 300 mm) in series were used for resolving the different samples. The flow rate of the THF eluent was 1 mL/min. The temperature of the columns was 33 °C. Molecular weights were calculated relative to polystyrene standards with a SM-105 kit of 10 molecular weight standards from Shodex. The average number degree of polymerization was also determined by NMR from the ratio of the imines to the terminal aldehyde protons. Thermal and Calorimetric Analyses. Thermogravimetric analyses (TGA) studies were carried out on a Hi-Res TGA 2950 thermogravimetric analyzer (TA Instruments) with a 10 °C/minute ramp to 700 °C (Tdec was defined as the onset of the decomposition

Given the importance of tuning the emission color for a given emitting application while preserving high Φfl, copolyazomethines consisting of thiophene and fluorene derivatives were prepared. It was expected that the HOMO−HOMO energygap could be tailored by incorporating different electron rich and deficient thiophenes, resulting in emissions covering the RGB primary colors. Similar to emission color tailoring, the oxidation potential was also expected to be modulated contingent on the electron richness of the thiophene units in addition to the placement of the azomethine nitrogen. The copolymers reported in Chart 1 would further provide insight into the structural elements required for reducing undesired fluorescence quenching processes and for making polyazomethines fluorescent. Knowing these key structural elements is important for designing and preparing polyazomethines that are consistently fluorescent, which to date, have been elusive. The preparation and opto-electronic characterization of the new conjugated polymers illustrated in Chart 1 are herein presented. The halochromism, chemical oxidation, and electrochromism of these new copolyazomethines are also presented for demonstrating their resistance toward acid hydrolysis and oxidation.



MATERIALS AND METHODS

Reagents and solvents from commercial sources were used as received unless otherwise stated. Anhydrous and deoxygenated solvents were obtained with an activated alumina column system. 1H and 13C NMR spectra were recorded at room temperature on a 400 MHz spectrometer. All the samples were dissolved in deuterated solvents and the spectra were referenced to the solvent line (CDCl3: 1H = 7.26 and 13C = 77.0 ppm) relative to TMS. CDCl3 for NMR polymer analyses was purified by repeatedly passing it over a plug of activated basic alumina. Spectroscopic Measurements. Absorption measurements were done on a Cary-500 spectrometer and fluorescence studies were carried out on an Edinburgh Instruments FLS-920 fluorimeter after deaerating the samples thoroughly with nitrogen for 20 min in Precision Glass quartz cuvettes. Relative fluorescence quantum yields were measured at 10−5 M by exciting the corresponding compounds at 1166

dx.doi.org/10.1021/ma2024304 | Macromolecules 2012, 45, 1165−1173

Macromolecules

Article

2.02 (m, −CH2), 1.33 (m, −CH2−CH2−), 0.80 (m, −CH2−), 0.59 (m, −CH3). Poly(2,3-dihydrothieno[3,4-b][1,4]dioxine)-co-9,9-dihexylfluorenylazomethine (PFl-co-EDOT). The copolymerization was similar to that above with 2,3-dihydrothieno[3,4-b][1,4]dioxine-5,7-dicarbaldehyde (163 mg, 0.882 mmol) and 1 (300 mg, 0.882 mmol) in CHCl3 (6 mL) in a pressure tube at 90 °C for 48 h. The reaction was then cooled to room temperature and it was poured into a methanol/water mixture. The resulting red precipitate was filtered. It was then washed consecutively with methanol, water and acetone. The title polymer was isolated as a red solid (60%, 140 mg) after drying under vacuum overnight. 1H NMR (400 MHz, CDCl3) δ ppm: 10.01 (m, CHO), 8.75 (m, −NCH−), 7.31−7.71 (m, Fl−H), 6.7 (b, NH2), 4.44−4.47 (m, O−CH2, EDOT), 1.99 (m, −CH2), 1.05 (m, −CH2−CH2−), 0.78 (m, −CH2−), 0.64 (m, −CH3). 9,9-Dihexyl-N,N′-bisthiophen-2-ylmethylenefluorene-2,7-diamine (Th-Fl-Th1). In ethanol (20 mL) were dissolved 1 (40.0 mg, 0.1 mmol) and thiophene-2-carbaxyldehyde (36 mg, 0.32 mmol) to which was then added a 1 M solution of TFA (20 μL). The solution was stirred at room temperature for 12 h, after which the solvent was evaporated and the product was extracted into ethyl acetate. The organic layer was washed with a brine solution and dried over Na2SO4. After filtering, the solvent was removed under reduced pressure and the crude product was purified by silica gel column chromatography with hexanes/ethyl acetate (80/20% v/v). The product was obtained as an orange solid (32 mg, 58% yield). 1H NMR (CDCl3), δ, ppm: 8.7 (s, 2H), 7.68 (d, 2H), 7.53 (d, 4H), 7.23−7.29 (m, 4H), 7.01 (s, 2H), 1.99 (t, 4H), 1.25 (p, 4H), 1.11 (m, 4H), 0.78 (t, 6H), 0.66 (t, 4H). 13 C NMR (acetone), δ, ppm: 164.9, 164.3, 161.5, 147.0, 144.1, 140.4, 133.4, 131.0, 102.2, 61.1, 60.1, 32.2, 32.0, 26.7, 22.9, 14.4, 14.1, 13.9. HR−MS (+): calculated for [C35H40N2S2 + H]+, 553.27089; found, 553.27089. 9,9-Dihexyl-N,N′-bis(5-butylthiophen-2-ylmethylene)fluorene2,7-diamine (Th−Fl−Th2). In ethanol (40 mL) were dissolved 1 (110.0 mg, 0.41 mmol) and 5-butylthiophene-2-carbaxyldehyde (222 mg, 1.32 mmol) to which was then added a 1 M solution of TFA (60 μL). The solution was stirred at room temperature for 12 h. The solvent was then removed under reduced pressure and the crude product was extracted into ethyl acetate. The organic layer was washed with a brine solution and dried over Na2SO4. After filtering and then removing the solvent under reduced pressure, the crude product was purified by silica gel column chromatography with hexanes/ethyl acetate (70/30% v/v). The product was obtained as an orange solid (212 mg, 78% yield). 1H NMR (acetone-d6), δ, ppm: 8.78 (s, 2H), 7.78 (d, 2H), 7.46 (d, 2H), 7.37 (d, 2H), 7.25 (d, 2H), 6.94 (d, 2H), 2.88 (t, 4H), 1.73 (p, 4H), 1.30 (t, 4H), 1.08 (p, 8H), 0.95 (t, 6H), 0.74 (t, 12H). 13C NMR (acetone), δ, ppm: HR−MS (+): calculated for [C43H56N2S2 + H]+, 665.39840; found, 665.39751.

temperature). Differential scanning calorimetry (DSC) studies were carried out on a TA Instruments with a 5 °C/minute ramp from −50 to +220 °C under both N2 and ambient atmosphere. Syntheses. The monomers 9,9-dihexyl-fluorene-2,7-diamine (1), 9,9-dioctyl-fluorene-2,7-dicarbaldehyde, diethyl 2,5-diaminothiophene3,4-dicarboxylate, and 2,3-dihydrothieno[3,4-b][1,4]dioxine-5,7-dicarbaldehyde were synthesized according to previously reported procedures.39−42 Similarly, PFl-co-Fl was prepared according to known methods.37 Experimental Procedures. 3-(2-Ethylhexyl) thiophene. In anhydrous THF (50 mL) was dissolved magnesium (2.65 g, 109 mmol) and 2-ethylhexyl bromide (15 g, 77 mmol). The reaction mixture was stirred for 30 min at 0 °C. The temperature was then gradually warmed to room temperature until most of the magnesium reacted, followed by refluxing for 2 h. In another flask was dissolved 3bromothiophene (7.0 g, 43 mmol) to which was then added [1, 3-bis (diphenylphosphino) propane] nickel(II) chloride (200 mg, 0.4 mmol) in anhydrous THF (100 mL). Three cycles of freeze− pump−thaw were performed to ensure complete removal of oxygen. The prepared 2-ethylhexyl magnesium bromide Grignard reagent was added by cannula to the red colored solution of nickel catalyst and the resulting brown solution was then refluxed for 18 h. The reaction mixture was washed with aqueous HCl (10% w/w) after cooling to room temperature. The organic phase was extracted with ethyl acetate, dried with MgSO4, and the solvent was evaporated after removing the MgSO4. The crude product was chromatographed on silica with 100% hexanes to afford a colorless oil (4.9 g, 58%). 1H NMR (CDCl3): δ = 7.23 (d, 1H), 6.94 (d, 2H), 2.53 (t, 2H), 1.44 (m, 1H), 1.30 (bs, 8H), 0.91 (t, 6H). 3-(2-Ethylhexyl)thiophene-2,5-dicarbaldehyde (5). To a solution of 3-(2-ethylhexyl) thiophene (7.5 g, 38 mmol) and freshly distilled TMEDA (9.8 g, 84 mmol) diluted in anhydrous hexanes (50 mL) under nitrogen, was added dropwise a solution of 2 M n- BuLi in hexane (42 mL, 84 mmol). After refluxing for 1.5 h, THF (40 mL) was added and the solution was cooled to −50 °C. Anhydrous DMF (14 mL, 190 mmol) was then added dropwise, after which the reaction mixture was allowed to warm to room temperature. After 2.5 h at room temperature, the reaction mixture was hydrolyzed with water (60 mL) and the mixture was extracted with ether. The organic layers were dried over MgSO4 and concentrated. The crude product was purified by column chromatography with hexanes/ethyl acetate (90/10 v/v) to give the product as a colorless oil (5.8 g, 61%). 1H NMR (CDCl3), δ, ppm: 10.09 (s, 1H), 9.56 (s, 1H), 7.16 (s, 1H), 2.61 (t, 2H), 1.59 (septet, 1H), 1.31 (qt, 2H), 1.04 (bs, 6H), 0.92 (t, 6H). HR−MS(+): calculated for [C14H20O2S + H]+, 253.12840; found, 253.12842. Poly(3-(2-ethylhexyl)thiophene)-co-9,9-dihexylfluorenylazomethine (PFl-co-Th). The copolymerization was done by mixing 1 (100 mg, 0.27 mmol) and 3-(2-ethyl-hexyl)-thiophene-2,5-dicarbaldehyde (68 mg, 0.27 mmol) in CHCl3 (3.0 mL) in a pressure tube. A catalytic amount of diluted TFA (30 μL) was then added. The pressure tube was sealed and heated to 90 °C for 48 h. The reaction mixture was cooled to room temperature and the polymer was precipitated from a methanol/water mixture. The resulting red solid (60%, 80 mg) was filtered and washed with methanol, water and acetone and then dried under vacuum overnight. 1H NMR (400 MHz, CDCl3), δ, ppm: 10.41 (m, CHO), 8.95 (m, −-NCH−), 8.40 (s, −Th−H) 7.34−7.71 (m, Fl−H), 6.71 (b, NH2), 2.82 (m, −CH2), 1.31 (m, −CH−CH2−), 1.93 (m, −CH2−), 0.77 (m, −CH3). Poly(thiophene-3,4-dicarboxylic acid diethyl ester)-co-9,9-dioctylfluorenylazomethine (PFl-co-DAT). The copolymerization was done similar to that above with 1 (485 mg, 1.88 mmol) and 9,9dioctyl-fluorene-2,7-dicarbaldehyde (840 mg, 1.88 mmol) in CHCl3 (10 mL) in a pressure tube at 90 °C for 48 h. The reaction mixture was cooled to room temperature and it was poured into a methanol/ water mixture. The resulting precipitate was filtered and it was washed with methanol, water and then acetone. The title polymer was isolated as a red solid (70%, 120 mg) after drying under vacuum overnight. 1H NMR (400 MHz, CDCl3) δ ppm: 10.1 (m, CHO), 8.52 (m, −N CH−), 7.53−7.91 (m, Fl-H), 6.35 (b, NH2), 4.24−4.49 (m, O−CH2),



DISCUSSION AND RESULTS The polymers presented in Chart 1 were targeted for examining the effect of incorporating different thiophenes on the polymer’s opto-electronic properties. The particular properties of interest are the emission colors and fluorescence quantum yields, given that previously investigated polyazomethines were all nonfluorescent. The exception is PFl-co-Fl that fluoresces in appreciable amounts (Table 2).37 The additional benefit of incorporating thiophene into the polymers is that significant spectral changes for both the absorbance and emission are expected relative to the all-fluorene PFl-co-Fl polymer. Significant spectral changes are additionally expected between PFl-co-Th vs PFl-co-EDOT, owing to the electronic differences of the EDOT and thiophene moieties. Meanwhile, the effect of the azomethine placement on the opto-electronic properties is expected by comparing the measured properties of PFl-co-DAT relative to both PFl-co-Th and PFl-co-EDOT. The monomers required for synthesizing the targeted copolymers were prepared by known means as outlined in 1167

dx.doi.org/10.1021/ma2024304 | Macromolecules 2012, 45, 1165−1173

Macromolecules

Article

the Supporting Information. Notably, the diaminothiophene was prepared in one-step from elemental sulfur via the Gewald reaction.43−45 Meanwhile, the diaminofluorene (1) was prepared in three steps in 72% overall yield starting from fluorene.37,40 It is noteworthy that the resulting 1 is not stable under ambient conditions for extended periods of time. It was therefore immediately polymerized once it was obtained in high purity. The complementary fluorene-2,7-dicarboxaldehyde was prepared from 2,7-dibromofluorene in two steps in 58% overall yield. The 3-(2-ethylhexyl)thiophene derivative was prepared from 3-bromothiophene via the Kumada reaction.34 This monomer was chosen over 3,4-dialkylated thiophenes because the resulting polymer was expected to be soluble. Moreover, its preparation gives rise to fewer side products, in contrast when starting with 3,4-dibromothiophene. The targeted polymers were prepared by combining stoichiometric amounts of the given monomers in chloroform in addition to a catalytic amount of trifluoroacetic acid. The homogeneous reaction mixture was subsequently heated to 90 °C in a sealed pressure tube for 48 h. In all cases, the polymers were red in color, confirming their increased degree of conjugation relative to their colorless monomers. While highly pure polymers could be obtained by simple removal of the solvent, the polymers were precipitated from a methanol/water mixture. This was to ensure that undesired low molecular weight oligomers and monomers, that would otherwise influence the opto-electronics properties, were removed. The targeted polymers were obtained in both good yields and high purity after precipitation. Both 1H NMR and GPC were used for confirming polymer formation. While formation of the desired azomethine is inferred by the change in color during the polymerization owing to an increase degree of conjugation, absolute confirmation was possible by 1H NMR. The azomethine proton has a distinctive chemical shift at ca. 8.5 ppm that does not overlap with other protons. This provides a clean window for accurate integration relative to the chain-end aldehyde, which also has a distinct resonance at ca. 10 ppm. The average degree of polymerization (DPn) was calculated by integrating imine protons relative to the terminal aldehydes (see Supporting Information). The average Mn calculated by NMR was consistent with that determined by relative GPC calibrated against polystyrene standards. This is somewhat surprising since the hydrodynamic radii of the polyazomethines and polystyrene standards are assumed to be different. This should lead to artificially higher molecular weights determined by GPC relative to those calculated by NMR. The same correlation between GPC and NMR molecular weight was also previously found for other polyazomethines.33,37 The results suggest that GPC provides relatively accurate molecular weight for polyazomethines. This aside, the measured molecular weights and thermal properties are summarized in Table 1. There is an appreciable difference in the degree of polymerization of PFl-co-DAT relative to PFl-co-Th and PFlco-EDOT (Table 1). However, this was not expected to impact the opto-electronic properties. This is based on our previous work with polythiophene azomethines that showed no change in opto-electronic properties with DPn > 10.33,37 While, no attempt was made to optimize the polymerization conditions, higher molecular weights are expected with extended polymerization times and at higher concentrations. The added advantage of polyazomethines is that the terminal functional groups remain active postpolymerization. The polymers can

Table 1. Molecular Weights and Thermal Properties of the Polyazomethines Studied polymer PFl-coTh PFl-coDAT PFl-coEDOT

Mn Mw Mn (g/mol)a (g/mol)b (g/mol)b

PDIb

DPn

Td Tg (°C)c (°C)d

7 200

7 700

9 300

1.20

13

330

115

15 100

16 400

22 100

1.35

23

320

146

6 600

6 900

11 900

1.72

13

380

148

a

Determined by 1H NMR. bDetermined by GPC relative to polystyrene standards. cDetermined by thermal analysis measured at 10 °C/min under N2. dDetermined by DSC measured at a 5 °C/min under N2.

therefore be either repolymerized resulting in increased molecular weights or reacted with other polyazomethines leading to extended block copolymers.33 The thermal properties of the polyfluorenylazomethines were investigated by both thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC). Interestingly, no weight loss was seen at 100 °C, confirming that the polymers are not hydroscopic and that the azomethine is not as a hydrate. This is contrast to the all-thiophene polyazomethine that was hydroscope.34 All the polymers investigated similarly showed resolvable glass transition temperatures (Tg) between 115° and 148 °C (Figure S1, Supporting Information). PFl-co-Th has the lowest Tg as result of the branched 2-ethylhexyl side chain. Meanwhile, both PFl-co-DAT and PFl-co-EDOT were found to have similar Tg, despite having different DPn. Photophysical Properties. Both the absorbance and fluorescence properties of the polyazomethines were investigated. As seen in Figure 1, there are appreciable differences in the spectra that are dependent on the type of thiophene incorporated into the polymer. Taking PFl-co-Fl as the benchmark (λabs = 450 and λfl = 560 nm), both the absorbance and fluorescence of the thiophene copolymers are bathochromically shifted. This is a result of the electron richness of the heterocycle. Also, it is known that heterocycles adopt coplanar arrangements with the azomethine bond to which they are attached.46 This further increases the degree of conjugation leading to the observed spectral changes. This is in contrast to homoaryls that are twisted by up to 65° from the azomethine bond to which they are attached.47−51 The absorbance of PFl-co-DAT is red-shifted by 40 nm from PFl-co-EDOT, which in turn is red-shifted by 60 nm from PFL-co-Th. The bathochromic shifts are a result of electronic effects. For example, the electron donating effect of the EDOT moiety works in concert with the electron withdrawing azomethine. The net result is an electronic push−pull effect. The bathochromic shift is more pronounced with PFl-co-DAT owing to the electronic pull−pull effect between the electron withdrawn esters and azomethines. The trend of bathochromic shifts is also observed with the fluorescence spectra shown in Figure 1B. The effect of the different thiophenes on the optical properties is further evident in the photographs in Figure 2. While subtle variations in color are visibly seen in the absorbance, the effect is more pronounced in the fluorescence. As seen in panels D−F of Figure 2, the three primary red, blue, and green emission colors are possible contingent on the thiophene incorporated into the copolymer. Bathochromic shifts were also seen with spin coated samples of the polyazomethines deposited as thin films. The thin film 1168

dx.doi.org/10.1021/ma2024304 | Macromolecules 2012, 45, 1165−1173

Macromolecules

Article

Figure 1. Normalized absorbance (A) and fluorescence (B) spectra in solution (solid line) and thin film (dashed line) of PFl-co-Th (black line), PFlco-DAT (red line), and PFl-co-EDOT (blue line).

Figure 2. Photographs of PFl-co-Th (A), PFl-co-DAT (B), and PFl-co-EDOT (C) in dichloromethane under ambient light and PFl-co-Th (D), PFlco-DAT (E), and PFl-co-EDOT (F) irradiated with a UV lamp at 350 nm.

Table 2. Photophysical and Electrochemical Properties of the Fluorenylazomethines λabs with added TFA (nm)

λabs with added FeCl3 (nm)

Egopt (eV)b

Epa (V)c

Epc (V)d

HOMO (eV)e

LUMO (eV)e

Egel (eV)f

Φfl (thinfilm)i

compound

λabs (nm)

Φfl rt (H )

Φfl 77 Kh

PFl-coDAT PFl-co-Th PFl-coEDOT PFl-co-Flj Th-Fl-Th1 Th-Fl-Th2

514, 553 (516,558)

576 (622)

565

600/652

2.11

0.93

−0.98

5.16

3.47

1.7

0.13 (0.66)



0.11

415 (438) 475 (493)

473 (546) 495 (565)

475 505

440/595 540

1.58 2.04

0.84 0.71

−1.05 −1.11

4.88 4.98

3.67 3.63

1.2 1.3

0.15 0.26 (0.85)

4× 3×

0.09 0.03

425 (452) 388 395

551 (563) 468 475

510 440 460

531 450 475

2.20 2.53 2.41

1.54 0.92 0.90

−0.87 -

5.94 5. 2 5.16

3.53 2.67 2.7

2.4 2.5 2.4

0.19 (0.75) 0.27 (0.86) 0.12 (0.42)

≈5× 3× 2×

0.4 -

a

λfl (nm)

a

+ g

a

In dichloromethane. Values in parentheses are for thin film samples. bSpectroscopically determined energy gap taken from the absorption onset (Eg = 1240/ λonset). cOxidation potential relative to saturated Ag/Ag+ electrode. dReduction potential relative to Ag/Ag+. eRelative to the vacuum level. f Electrochemically derived energy gap. gAbsolute fluorescence quantum yield at room temperature. Values in parentheses are the room temperature absolute fluorescence quantum yield upon protonation with TFA. hFluorescence enhancement at 77 K relative to room temperature and uncorrected for the change in refractive index at different temperatures. iSpin-coated samples. jFrom ref 37.

As seen in Table 2, the polymers fluoresce in considerable amounts. In fact, the Φfl are comparable to polyfluorene vinylenes that are used emitting devices.29 This is in contrast to previously investigated thiophene azomethine derivatives whose fluorescence quantum yields were ca. 0.01. 46,49−51,55,56 Analogous copolyazomethines derived from 9,9-unsubstituted 2,7-diaminofluorene also do not fluoresce.25,26 As seen in the photographs D−F in Figure 2, the polymer fluorescence can visually be detected by the naked eye. The less intense emission seen in Figure 2F,G is from the poor photoresponse of the camera’s PMT and not from a low fluorescence yield. This is supported by the absolute Φfl of the polymers that are all on the order of 10%. High Φfl were even observed in thin films, suggesting that there is little self-quenching. The measured

absorbance spectra are bathochromically shifted by ca. 20 nm while the fluorescence spectra are shifted by about 70 nm from the corresponding polyazomethine in solution. The property of particular interest for the polyazomethines is their fluorescence quantum yield (Φfl). This is because the azomethine linkage is known to quench the fluorescence of even the most intrinsically fluorescent fluorophores.46−54 The desired fluorescence quantum yields were measured using an integrating sphere. This method is preferred over relative methods because the absolute quantum yields can be measured independent of references whose fluorescence yields are wavelength dependent and typically contain errors up to 20%. The Φfl can therefore be accurately determined. 1169

dx.doi.org/10.1021/ma2024304 | Macromolecules 2012, 45, 1165−1173

Macromolecules

Article

absolute fluorescence quantum yields when compared to previously reported studies, confirm that 9,9-dialkylation of the diaminofluorene is a critical structural element for preserving the inherent fluorescence of the fluorene. While the polymers’ Φfl are ca. 10%, their fluorescence could be restored to near unity by two means; low temperature and acid doping. Reducing the temperature to 77 K resulted in fluorescence yield enhancement. It should be noted that the calculated values at 77 K are uncorrected for the change in solvent refractive index with temperature. Therefore, the measured values are lower limits and the true amount of fluorescence enhancement is expected to be 20% higher. Nonetheless, the increased fluorescence with temperature implies that reduced fluorescence at room temperature most likely involves bond rotation deactivation processes. Similar to the temperature induced fluorescence changes, fluorescence enhancement was also possible by protonating the azomethine with trifluoroacetic acid. The exact mechanism responsible for fluorescence increase with protonation has not yet been confirmed. Despite this, the collective fluorescence measurements confirm that polyazomethines can me made fluorescent. Meanwhile, their fluorescence enhancement with temperature and protonation illustrate the sensor-like properties of polyazomethines. The thiophene-fluorene triads (Th-Fl-Th1 and Th-Fl-Th2) were prepared to investigate whether high Φfl were also possible with triads derived from 1 and having reduced degree of conjugation relative to their polymer counterparts. Equally high Φfl were observed with the model compounds. They similarly exhibited the same fluorescence enhancement at low temperature and with acid protonation. The key component for achieving fluorescent fluorene azomethine derivatives is therefore the 9,9-dialkylation derivative of 2,7-diaminofluorene. This is evident from the unalkylated derivatives of Th-Fl-Th that are nonfluorescent (Φfl< 0.03).49,55 While the exact structural composition required for making highly fluorescent azomethines is currently being investigated, the collective spectroscopic data confirm that both the color and emission can be tailored as a function of the thiophene. Also, fluorescent alternating copolyazomethines that emit with the primary RGB colors are possible by adjusting the HOMO−LUMO energy levels. Electrochemistry. The electrochemical properties of the polymers were investigated by cyclic voltammetry to further gain insight into structural effects on the oxidation potential (Epa). As seen in Figure 3, the polymers all undergo an oxidation process. Both the Epa and the reversibility of the anodic process are contingent on the thiophene. For example, the Epa of PFl-co-EDOT is less positive than PFl-co-Th by 130 mV, as expected owing to the strong electron donating character of EDOT. Meanwhile, the Epa of PFl-co-DAT is 70 mV higher than PFl-co-Th. The increased Epa is expected given the strong electron withdrawing character of the thiophene esters. The measured Epa of the polyazomethines confirm that the polymers are stable under ambient conditions. Meanwhile, the reversible oxidation for both PFl-co-Th and PFl-co-EDOT confirm the robustness of the heteroatomic bond under oxidation conditions. The polymers also underwent a cathodic processes corresponding to a two-electron transfer. Interestingly, the reduction potential (Epc) was less affected by the thiophene incorporated into the copolymer than the corresponding Epa.

Figure 3. Cyclic voltammograms of PFl-co-Th (black line), PFl-coDAT (red line), PFl-co-EDOT (blue line), Th−Fl−Th1 (green line), and Th−Fl−Th2 (maroon line) in dichloromethane at a scan rate of 50 mV/s under N2. Inset: Photographs of original (left) and electrochemically oxidized (right) PFl-co-DAT by applying a potential of 1.0 V for 1 min.

Only in the case of PFl-co-EDOT (Figure S14, Supporting Information) was the cathodic process reversible. The electrochemistry of the Th−Fl−Th triads was also investigated. Comparing their Epa to that of PFl-co-Th would provide insight into the effect of the degree of conjugation on the oxidation potential. Furthermore, comparing the anodic behavior of Th−Fl−Th2 relative to Th−Fl−Th1 would provide information about both the stability of the radical cation and the azomethine bond. The Epa of both Th−Fl−Th were 80 mV more positive than the corresponding polymer. While the electron withdrawing character of the azomethine should increase the oxidation potential, the high degree of conjugation of PFl-co-Th offsets this effect, leading to the observed less positive Epa. As expected, there is a noticeable difference between the reversibility of the oxidation process of Th−Fl− Th1 and Th−Fl−Th2. The oxidation process of Th−Fl−Th2 was reversible while that of Th−Fl−Th1 wass pseudoreversible. The difference in reversibility is uniquely a result of the substitution in the α,α′ position. The α,α′ alkylation of Th−Fl− Th2 prevents the electrochemically generated radical cation from coupling according to known means.57−60 The observed reversible oxidation confirms that the azomethine is robust. It further confirms that heteroatomic bond does not decompose when oxidized. Meanwhile, the pseudoreversible oxidation of Th−Fl−Th1 suggests that the radical cation is not highly reactive. This is most likely a result of the high degree of conjugation of the compounds allowing the radical cation to be delocalized across the entire molecule. The spectroscopic properties of the electrochemically generated radical cation were additionally investigated by spectroelectrochemistry in solution. From an application point of view, spectroelectrochemical studies of thin films of the polymers would be more advantageous. However, their solubility in standard electrochemical solutions precludes such studies. Spectroelectrochemical studies in solution were done by applying a potential slightly more positive than the Epa of the given polymer and observing the resulting absorbance changes. In all cases, the absorbance of the generated radical cation was bathochromically shifted from the corresponding neutral form (Table 2). The most stark visible color change was with PFl-co1170

dx.doi.org/10.1021/ma2024304 | Macromolecules 2012, 45, 1165−1173

Macromolecules

Article

the absorbance upon chemical doping. The oxidized species were subsequently neutralized to the original state after equilibrating for a few minutes and by adding excess hydrazine hydrate. The reversible color changes between the neutral and oxidized states are seen in the pictures of Figure 5. Both the original intensity and absorbance of the neutral azomethines were observed when reducing the oxidized states with hydrazine hydrate. It should be noted that no special precautions such as anhydrous or degassed solvents were used for the oxidation/neutralization studies. Hypsochromic shifts, decrease in intensity, and even complete disappearance of the absorbance would be observed if the polyazomethines decomposed or hydrolyzed during the oxidation/neutralization cycle. The absence of detectable change in the absorbance spectra between the original and the oxidized/neutralized azomethines (Figures 4 and 5) confirm the azomethines are both hydrolytically and oxidatively robust. The exception is PFl-co-DAT whose original color cannot be restored once neutralized with hydrazine hydrate (Figure S10, Supporting Information). The added advantage of oxidation studies in solution is that the stoichiometry can be spectroscopically determined via the Job method. By varying the mole fraction of the oxidant, the stoichiometry can be derived from the absorbance maximum, as seen in the inset of Figure 4. A 1:1 stoichiometry was found for the all the compounds in Chart 1. Given that ferric chloride is a one-electron oxidant taken together with the determined stoichiometry, the chemically generated intermediate is a radical cation. While this intermediate is consistent with the electrochemically produced species measured by cyclic voltammetry, there are spectroscopic differences between the chemically and electrochemically generated species (Figures S4−S6, Supporting Information). The spectroelectrochemically observed intermediates in solution for the all the polymers were hypsochromically shifted by 40 nm relative to their corresponding intermediates produced by FeCl3. The observed difference may be a counteranion effect, which is currently being examined. Azomethines are known to be halochromic with significant spectral changes occurring when protonated.31,36,37,46,61 The advantage of the studied polyazomethines is that they absorb strongly in the visible as a result of their extended degree of conjugation. Therefore, their protonation should lead to significant visible color changes. This provides the means to spectroscopically follow the stability of the protonated form and to investigate their resistance toward acid hydrolysis. The halochromic behavior (color change with protonation) of the azomethines in Chart 1 was subsequently investigated with

DAT, whose color changed from red to blue upon oxidation. This is seen in the inset of Figure 3. The new absorbance occurred at 620 nm while the absorbance of the electrochemically generated radical cation for PFl-co-Th and PFl-co-EDOT was 442 and 450 nm, respectively. The original color of the polymers could be quantitatively restored by applying a potential of −100 mV. This once again illustrates the robustness of the polyazomethines. Halochromism and Chemical Oxidation. Given that electrochemically induced color changes were possible, the chemical doping of the polymers was additionally investigated. Oxidizing the polymers in solution has the advantage that the stability of the doped state can be assessed by monitoring the reversible color changes. This is in contrast to the spectroelectrochemical studies where diffusion does not occur and the intermediates are artificially stable. Dichloromethane solutions of the copolymers were oxidized with FeCl3. This is a choice oxidant because it is spectroscopically transparent in the spectral window investigated. Its reduction potential is also compatible with the oxidation potentials of the polymers investigated. Significant color changes occurred upon adding the oxidant and the resulting absorbance shifts summarized in Table 2. A representative absorbance spectral change of PFl-co-EDOT is shown in Figure 4, illustrating a 50 nm bathochromic shift in

Figure 4. Normalized absorbance spectra of PFl-co-EDOT in dichloromethane (▲) with the addition of 1 equiv of TFA (red ◆), followed by the addition of 1 equiv of triethylamine (blue ●), a pristine sample of PFl-co-EDOT (▲) with the addition of 1 equiv of FeCl3 (green ■), followed by the addition of 1 equiv of hydrazine hydrate (□). Inset: Absorbance of PFl-co-Th monitored at 595 nm as a function of FeCl3 mole fraction.

Figure 5. Pictures of PFl-co-Th (A), PFl-co-EDOT (B), and PFl-co-DAT (C) in dichloromethane (left), after the addition of FeCl3 (right), followed by the addition of hydrazine hydrate (left). 1171

dx.doi.org/10.1021/ma2024304 | Macromolecules 2012, 45, 1165−1173

Macromolecules



ACKNOWLEDGMENTS The Natural Sciences and Engineering Research Council Canada is acknowledged for financial support in terms of Discovery, Strategic Research, and Research Tools and Instruments grants. The Canada Foundation for Innovation is also thanked for equipment funding. The Centre for SelfAssembled Chemical Structures is also acknowledged.

trifluoroacetic acid. Exposing the polyazomethines to concentrated acid resulted in absorbance bathochromic shifts. However, the acid induced spectral changes were slightly different than those observed with ferric chloride. This is owing to the resulting cation that has weaker charge transfer effect than the radical cation that is produced by oxidation with ferric chloride (vide supra). The protonated azomethines could be neutralized by adding triethylamine. Similar to the oxidation studies, the azomethines could be reversibly protonated and neutralized without any spectroscopically detectable decomposition. The exception being PFl-co-DAT that undergoes minor color degradation. The reversible protonation/neutralization is illustrated in Figure 4. The salient features of the absorbance spectrum are the reversible color change without color degradation after protonation and neutralization. Given that anhydrous solvents or special precautions were not used for the halochromic studies, residual water in the solvent or ambient moisture would readily hydrolyze the azomethine bond upon acid addition. However, no color degradation was observed after protonation/neutralization cycles. This demonstrates the resistance of the azomethines toward acid hydrolysis. The collective halochromic and oxidation spectroscopic studies confirm that the conjugated azomethine derivatives are robust and they resist both oxidation and acid hydrolysis.



CONCLUSION



ASSOCIATED CONTENT



REFERENCES

(1) Szarko, J. M.; Guo, J.; Rolczynski, B. S.; Chen, L. X. J. Mater. Chem. 2011, 21, 7849. (2) Koo, H.-J.; Chang, S. T.; Slocik, J. M.; Naik, R. R.; Velev, O. D. J. Mater. Chem. 2011, 21, 72. (3) Xue, J. Polym. Rev. 2010, 50, 411. (4) Wong, W.-Y.; Tang, B. Z. Macromol. Chem. Phys. 2010, 211, 2460. (5) Gather, M. C.; Köhnen, A.; Meerholz, K. Adv. Mater. 2011, 23, 233. (6) Kamtekar, K. T.; Monkman, A. P.; Bryce, M. R. Adv. Mater. 2010, 22, 572. (7) Arias, A. C.; MacKenzie, J. D.; McCulloch, I.; Rivnay, J.; Salleo, A. Chem. Rev. 2010, 110, 3. (8) Ma, H.; Yip, H.-L.; Huang, F.; Jen, A. K. Y. Adv. Funct. Mater. 2010, 20, 1371. (9) Facchetti, A. Chem. Mater. 2010, 23, 733. (10) Bassani, D. M.; Jonusauskaite, L.; Lavie-Cambot, A.; McClenaghan, N. D.; Pozzo, J.-L.; Ray, D.; Vives, G. Coord. Chem. Rev. 2010, 254, 2429. (11) Wang, Y.; Tan, H.; Liu, Y.; Jiang, C.; Hu, Z.; Zhu, M.; Wang, L.; Zhu, W.; Cao, Y. Tetrahedron 2010, 66, 1483. (12) Zhao, Q.; Liu, S.-J.; Huang, W. Macromol. Chem. Phys. 2009, 210, 1580. (13) Wang, P.-H.; Ho, M.-S.; Yang, S.-H.; Chen, K.-B.; Hsu, C.-S. J. Polym. Chem. A 2010, 48, 516. (14) Kucherak, O. A.; Didier, P.; Mély, Y.; Klymchenko, A. S. J. Phys. Chem. Lett. 2010, 1, 616. (15) Wu, H. Y.; Wang, K. L.; Jiang, J. C.; Liaw, D. J.; Lee, K. R.; Lai, J. Y.; Chen, C. L. J. Polym. Sci., Part A 2010, 48, 3913. (16) Scheler, E.; Strohriegl, P. Chem. Mater. 2010, 22, 1410. (17) Li, K.; Liu, B. Polym. Chem. 2010, 1, 252. (18) Wei, W.; Yu, N.; Zhu, R.; Peng, B.; Huang, W. J. Appl. Polym. Sci. 2008, 108, 2438. (19) Jin, S.-H.; Kang, S.-Y.; Kim, M.-Y.; Chan, Y. U.; Kim, J. Y.; Lee, K.; Gal, Y.-S. Macromolecules 2003, 36, 3841. (20) Phanichphant, S.; Anuragudom, P.; Newaz, S. S.; Lee, T. R. Macromolecules 2006, 39, 3494. (21) Leclerc, M.; Ranger, M.; Rondeau, D. Macromolecules 1997, 30, 7686. (22) Scherf, U.; Nothofer, H. G.; Meisel, A.; Miteva, T.; Neher, D.; Forster, M.; Oda, H.; Lieser, G.; Sainova, D.; Yasuda, A.; Lupo, D.; Knoll, W. Macromol. Symp. 2000, 154, 139. (23) Miller, R. D.; Kreyenschmidt, M.; Klaerner, G.; Fuhrer, T.; Ashenhurst, J.; Karg, S.; Chen, W. D.; Lee, V. Y.; Scott, J. C. Macromolecules 1998, 31, 1099. (24) Lee, T. S.; Kim, H. J.; Lee, J. H.; Lee, M. React. Funct. Polym 2008, 68, 1696. (25) Tsai, F.-C.; Chang, C.-C.; Liu, C.-L.; Chen, W.-C.; Jenekhe, S. A. Macromolecules 2005, 38, 1958. (26) Liu, C.-L.; Chen, W.-C. Macromol. Chem. Phys. 2005, 206, 2212. (27) Hyun-Chul Kim, J.-S. K.; Kim, K.-S.; Park, H.-K.; Baek, S.; Ree, M. J. Polym. Sci., Part A 2004, 42, 825. (28) Bruma, M.; Hamciuc, E.; Schulz, B.; Koepnick, T.; Kaminorz, Y.; Robinson, J. Macromol. Symp. 2003, 199, 511. (29) Marin, L.; Perju, E.; Damaceanu, M. D. Eur. Polym. J. 2011, 47, 1284. (30) Liu, C. L.; Chen, W. C. Macromol. Chem. Phys. 2005, 206, 2212. (31) Barik, S.; Bishop, S.; Skene, W. G. Mater. Chem. Phys. 2011, 129, 529.

Stable and conjugated copolyazomethines consisting of alternating fluorene and thiophenes were prepared. Both the absorbance and fluorescence of the polymers were contingent on the type of thiophene incorporated, with emission of the three primary colors being possible. The polymers were found to be consistently fluorescent with fluorescence quantum yield enhancement being possible by either protonation or with temperature decrease. The reported polymers are among the first examples of fluorescent polyazomethines. It was found that a key structural element for making polyazomethines fluorescent was the use of 9,9-dialkyl-2,7-diamino fluorene. The reasonable Φfl make the polyazomethines interesting candidates for use in emitting applications especially when considering their ease of preparation and purification. Meanwhile, the temperature and acid dependent fluorescence make the polyazomethines interesting sensor-like materials. It was further demonstrated that the polyazomethines resists both chemical and electrochemical oxidation in addition to acid hydrolysis, contrary to common belief. In fact, the stark color changes possible between their neutral and oxidized/protonated states concomitant with the chemical and hydrolytic robustness make these polyazomethines interesting materials for use in devices such as electrochromics and sensor applications.

S Supporting Information *

Synthetic schemes, absorbance, fluorescence, NMR, and doping spectra and GPC and electrochemical data. This material is available free of charge via the Internet at http://pubs.acs.org/.



Article

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. 1172

dx.doi.org/10.1021/ma2024304 | Macromolecules 2012, 45, 1165−1173

Macromolecules

Article

(32) Bolduc, A.; Lachapelle, V.; Skene, W. G. Macromol. Symp. 2010, 297, 87. (33) Barik, S.; Skene, W. G. Macromolecules 2010, 43, 10435. (34) Bourgeaux, M.; Skene, W. G. Macromolecules 2007, 40, 1792. (35) Pérez Guarìn, S. A.; Skene, W. G. Mater. Lett. 2007, 61, 5102. (36) Bolduc, A.; Dufresne, S.; Skene, W. G. J. Mater. Chem. 2010, 20, 4820. (37) Barik, S.; Skene, W. G. Polym. Chem. 2011, 2, 1091. (38) Bright, D. W.; Moss, K. C.; Kamtekar, K. T.; Bryce, M. R.; Monkman, A. P. Macromol. Rapid Commun. 2011, 32, 983. (39) Pérez Guarìn, S. A.; Bourgeaux, M.; Dufresne, S.; Skene, W. G. J. Org. Chem. 2007, 72, 2631. (40) Barik, S.; Friedland, S.; Skene, W. G. Can. J. Chem. 2010, 88, 945. (41) Pérez Guarìn, S. A.; Dufresne, S.; Tsang, D.; Sylla, A.; Skene, W. G. J. Mater. Chem. 2007, 17, 2801. (42) Tsang, D.; Bourgeaux, M.; Skene, W. G. J. Photochem. Photobiol. A 2007, 192, 122. (43) Gewald, K. Chem. Hetero. Comp. 1976, 12, 1077. (44) Gütschow, M.; Schröter, H.; Kuhnle, G.; Eger, K. Monats. Chem. 1996, 127, 297. (45) Bourgeaux, M.; Vomscheid, S.; Skene, W. G. Synth. Commun. 2007, 37, 3551. (46) Dufresne, S.; Skene, W. G. J. Phys. Org. Chem. 2011, No. DOI: 10.1002/poc.1894. (47) Dufresne, S.; Skalski, T.; Skene, W. G. Can. J. Chem. 2011, 89, 173. (48) Dufresne, S.; Roche, I. U.; Skalski, T.; Skene, W. G. J. Phys. Chem. 2010, 114, 13106. (49) Dufresne, S.; Guarìn, S. A. P.; Bolduc, A.; Bourque, A. N.; Skene, W. G. Photochem. Photobiol. Sci. 2009, 8, 796. (50) Dufresne, S.; Callaghan, L.; Skene, W. G. J. Phys. Chem. B 2009, 13, 15541. (51) Bourque, A. N.; Dufresne, S.; Skene, W. G. J. Phys. Chem. C 2009, 113, 19677. (52) Knipping, É.; Roche, I. U.; Dufresne, S.; McGregor, N.; Skene, W. G. Tetrahedron Lett. 2011, 52, 4385. (53) Dong, Y.; Bolduc, A. a.; McGregor, N.; Skene, W. G. Org. Lett. 2011, 13, 1844. (54) Bolduc, A.; Dufresne, S.; Hanan, G. S.; Skene, W. G. Can. J. Chem. 2010, 88, 236. (55) Skene, W. G.; Pérez Guarìn, S. A. J. Fluores. 2007, 17, 540. (56) Dufresne, S.; Hanan, G. S.; Skene, W. G. J. Phys. Chem. B 2007, 111, 11407. (57) Natera, J.; Otero, L.; Sereno, L.; Fungo, F.; Wang, N.-S.; Tsai, Y.-M.; Hwu, T.-Y.; Wong, K.-T. Macromolecules 2007, 40, 4456. (58) Zotti, G.; Zecchin, S.; Schiavon, G.; Vercelli, B.; Berlin, A. J. Electroanal. Chem. 2005, 575, 169. (59) Hapiot, P.; Lagrost, C.; LeFloch, F.; Raoult, E.; Rault-Berthelot, J. Chem. Mater. 2005, 17, 2003. (60) Roncali, J. Chem. Soc. Rev. 2005, 34, 483. (61) Dufresne, S.; Bolduc, A.; Skene, W. G. J. Mater. Chem. 2010, 20, 4861.

1173

dx.doi.org/10.1021/ma2024304 | Macromolecules 2012, 45, 1165−1173