Next Article in Journal
A Highly Hydrophilic and Biodegradable Novel Poly(amide-imide) for Biomedical Applications
Next Article in Special Issue
Rapid and Effective Removal of Cu2+ from Aqueous Solution Using Novel Chitosan and Laponite-Based Nanocomposite as Adsorbent
Previous Article in Journal
Applications of Solid-State NMR Spectroscopy for the Study of Lipid Membranes with Polyphilic Guest (Macro)Molecules
Previous Article in Special Issue
Preparation of Magnetic Iron Oxide Nanoparticles (MIONs) with Improved Saturation Magnetization Using Multifunctional Polymer Ligand
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Polythiophene–Fullerene Hybrid Additives as Potential Compatibilizers of BHJ Active Layers

1
Department of Chemistry, University of Patras, University Campus, GR26504 Patras, Greece
2
Foundation for Research and Technology Hellas/Institute of Chemical Engineering Sciences (FORTH/ICE-HT), GR26504 Patras, Greece
*
Author to whom correspondence should be addressed.
Polymers 2016, 8(12), 440; https://doi.org/10.3390/polym8120440
Submission received: 11 November 2016 / Revised: 9 December 2016 / Accepted: 13 December 2016 / Published: 18 December 2016
(This article belongs to the Special Issue Hybrid Polymeric Materials)

Abstract

:
Perfluorophenyl functionalities have been introduced as side chain substituents onto regioregular poly(3-hexyl thiophene) (rr-P3HT), under various percentages. These functional groups were then converted to azides which were used to create polymeric hybrid materials with fullerene species, either C60 or C70. The P3HT–fullerene hybrids thus formed were thereafter evaluated as potential compatibilizers of BHJ active layers comprising P3HT and fullerene based acceptors. Therefore, a systematic investigation of the optical and morphological properties of the purified polymer–fullerene hybrid materials was performed, via different complementary techniques. Additionally, P3HT:PC70BM blends containing various percentages of the herein synthesized hybrid material comprising rr-P3HT and C70 were investigated via Transmission Electron Microscopy (TEM) in an effort to understand the effect of the hybrids as additives on the morphology and nanophase separation of this typically used active layer blend for OPVs.

Graphical Abstract

1. Introduction

Bulk Heterojunction (BHJ) solar cell devices are widely investigated due to their potential to reach high efficiencies along with their potential of large area coverage and unique flexibility [1,2,3]. Current state of the art materials that form the bicontinuous interpenetrating network of the active layer are poly(3-hexylthiophene) (P3HT) and [6,6]-phenyl-C61-butyric acid methyl ester (PCBM), providing power conversion efficiencies (PCE) of above 6%, because of their thermal stability [4], high electron mobility [5,6] and especially because it is the current system of choice for printable over large areas organic photovoltaics (OPVs) [7,8,9,10]. Among others, decisive parameters that affect the device performance are the materials comprising the active layer, the morphology of the blend as well as the device architecture. In the direction of materials development, new polymeric electron donating materials with complex architectures and lower energy levels [11,12,13,14,15] compared to P3HT, and new electron acceptors based on fullerene [16,17,18] or even fully organic and polymeric ones [19,20,21], have been reported. On the other hand, efforts to control the morphology [22,23] of the active blend have been made using thermal annealing or solvent treatment of the blend resulting in significantly higher PCEs [1,4,24,25,26]. Besides the above-mentioned manipulation of the active layer, another approach that has been followed is the modification of the semiconducting polymer with functional groups that may lead to more complex architectures either co-polymeric or hybrid ones. More specifically, co-polymeric materials based on polythiophenes or other conjugated polymers that can interact with or bear fullerene species can provide better miscibility and tune the nanophase separation of the two plain blend components [27,28,29,30,31,32,33]. In some cases, the introduction of functional groups like hydroxyl or azide units, as side chain substituents or as end chain groups, not only controls but also stabilizes the morphology of the blended materials [34,35,36,37,38]. Another efficient way to control the morphology is the introduction into the active layer of small quantities of hybrid materials comprising the semiconducting polymer and a carbon nanostructure [39,40,41,42,43,44,45,46,47,48]. However, the synthesis of most semiconducting polymeric or copolymeric hybrids involves complicated and multistep synthetic procedures that affect the yield and the cost of the final materials. Moreover, in most such cases, a non-conjugated linkage between the semiconducting polymer and the carbon nanostructure has been employed preventing electronic interactions among the polymeric and fullerene parts of the hybrid materials. To overcome the above bottlenecks, new synthetic strategies involving the “direct” attachment of the two components, organic and fullerene ones, comprising the hybrid materials have been developed [49,50,51,52,53,54]. Working in this direction, we have recently reported a straightforward methodology for the development of hybrid semiconducting polymeric materials using the electron deficient perfluorophenyl functionality inserted either onto copolymeric materials or as end polymeric chain group [49,54]. Benefits of this methodology are the direct electronic interactions between the two components [50], the absence of extra aliphatic or other non-conjugated parts used as connection bridges which increase the volume of insulating material within the active layer as well as the use of the intermediate perfluorophenyl aromatic azides that are stable over time and ambient storage conditions.
In this present work, we focused on the introduction of perfluorophenyl functionalities as side chain substituents of regioregular-poly(3-hexyl thiophene) (rr-P3HT). Several ratios were attempted in order to investigate the influence of the side functionalization onto the properties of rr-P3HT. Thereafter, using our previously reported methodology [49,54], the perfluorophenyl functionalities were transformed into azides that performed a [3+2] cycloaddition reaction with C60 or C70 fullerene species to provide the desired hybrid materials. Extensive purification has been performed in all cases and the purified materials were characterized in respect to their optical and morphological properties. These particular hybrid materials are ideal candidates for the compatibilization and stabilization of P3HT:fullerene active layers used in BHJ polymer photovoltaics, since the introduction of the fullerene species directly along the polythiophene’s backbone ensures the hybrids’ incorporation in the nanophase created by the neat blend components potentially holding them in close proximity over time and temperature variations during device operation. In order to support the hybrids potentiality as compatibilizers, mixtures of typical P3HT:PC70BM active BHJ blends with various percentages of the C70 based hybrid material were investigated in respect to the effect of the additive on the morphology of the blend.

2. Materials and Methods

2.1. Materials

rr-P3HT was synthesized according to literature procedure [55]. Fullerene C60 (99.5%) was purchased from SES research. Carbon C70 (99.5%) was kindly provided by Prof. Kyriakos Porfyrakis of the Oxford Carbon Nanomaterials Group, Department of Materials, University of Oxford, UK. Tetrahydrofuran was purchased from Aldrich (Sigma-Aldrich Chemie GmbH, Taufkirchen, Germany) and was distilled with benzophenone and metallic sodium (THFdry). All the other solvents and reagents were purchased from Aldrich and used without further purification unless otherwise stated.

2.2. Instrumentation

1H, 19F, 15N Nuclear Magnetic Resonance (NMR) spectra were recorded on a Bruker Advance (Bruker BioSpin GmbH, Magnet Division, Karlsruhe, Germany) DPX 400.13, 376.5 and 40.55 MHz spectrometer, respectively, in CDCl3 containing TMS as internal standard.
Gel permeation chromatography (GPC) measurements were carried out using a Polymer Lab chromatographer (Agilent Technologies, Santa Clara, CA, USA) equipped with two PLgel 5 μm mixed columns and a UV detector, using CHCl3 as eluent with a flow rate of 1 mL/min at 25 °C, calibrated versus polystyrene standards.
Attenuated Total Reflectance (ATR) spectra were recorded on a “Bruker Optics’ Alpha-P Diamond ATR Spectrometer of Bruker Optics GmbH” (Ettlingen, Germany).
Thermogravimetric analysis (TGA) was carried out on ~8 mg samples contained in alumina crucibles in a Labsys TM TG apparatus of Setaram (Caluire, France) under nitrogen and at a heating rate of 10 °C/min.
UV-Vis spectra were recorded using a Hitachi U-1800 spectrophotometer (Hitachi High-Technologies Europe GmbH, Mannheim, Germany) Continuous wave photoluminescence was measured on a Perkin Elmer LS50B spectrofluorometer (Waltham, MA, USA). All UV-Vis and PL measurements were performed in air using quartz cuvettes and flat substrates for the examination of solutions and films, respectively.
The Blend Preparation involved the mixing of stock solutions of P3HT (10 mg·mL−1) and PC70BM (10 mg·mL−1) to prepare the P3HT:PC70BM 1:1 w/w. Afterwards, 0.125 and 0.625 mL of the [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] (8 mg·mL−1) were added to prepare the 20% and the 50% blends with the hybrid polymer, respectively.
Transmission electron microscopy (TEM) measurements were performed on a JEOL JEM2100 (Peabody, MA, USA) operating at 200 kV. Sample preparation for TEM examination of the [P3HT0.6-(P3HT-5F)0.4], [P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N-C60)0.2], [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] involved the preparation of dilute solutions of the samples in o-DCB and filtration through a 0.45 μm filter. For the hybrids a drop of the solution was placed on 3 mm carbon coated copper grids (Electron Microscopy Sciences) and the samples were dried in air for 2 days. For the blends examination via TEM, a drop of the blends solution was placed on 3 mm carbon coated copper grids (Electron Microscopy Sciences) and after 30 s the solvent was removed with filter paper.

2.3. Synthetic Procedures

2.3.1. Side [P3HT0.6-(P3HT-Br)0.4] [56]

A 50 mL round bottom flask, equipped with a reflux condenser and a magnetic stirrer, was flamed under vacuum degassed and filled with argon. P3HT (500 mg, 3 mmol) was added and dissolved in 30 mL CHCl3 and the system was degassed and flushed with argon again. The flask was covered with aluminum foil and NBS (214 mg, 1.2 mmol) was added in three portions. The mixture was stirred for 22 h at room temperature. Then, ~15 mL Na2SO3 5% aqueous solution were added to the system and the mixture was extracted with CHCl3 and deionized water. The organic phase was dried with anhydrous MgSO4, filtrated and rotary evaporated. The obtained solid was dried under vacuum at 40 °C. Bromination ratio calculated from 1H NMR: 40%, yield: 92% 1H NMR (δH; CDCl3; Me4Si): 7.00(broad, 1H), 2.80(broad, 2H), 2.64(broad, 2H–Br substituted thiophene), 1.65(broad, 4H), 1.53–1.22(broad, 12H), 0.88(broad, 6H).
The same procedure was used for all other ratios.

2.3.2. Side [P3HT0.6-(P3HT-5F)0.4]

A 100 mL round bottom flask, equipped with a reflux condenser and a magnetic stirrer, was degassed (flamed under vacuum) and filled with argon. Side [P3HT0.6-(P3HT-Br)0.4] (400 mg, 0.81 mmol), pentafluorophenyl boronic acid (205 mg, 1 mmol) and Pd(PPh3)4 (30 mg, 0.02 mmol) were dissolved in 20 mL toluene and then Na2CO3 (1.20 mL, 2.42 mmol) in 2 M aqueous were added. The mixture was degassed and flushed with argon again and refluxed for 72 h. Then, the mixture was filtered in order to remove the catalyst and the solvent was rotary evaporated. The obtained solid was washed several times with MeOH and deionized water. The final solid was dried under vacuum at 50 °C. Yield: 99% 1H NMR (δH; CDCl3; Me4Si): 7.00(broad, 1H), 2.80(broad, 2H), 2.64(broad, 2H), 1.65(broad, 4H), 1.53–1.22(broad, 12H), 0.88(broad, 6H), 19F NMR (δF; CDCl3; Me4Si): −139.10, −150.91, −159.94.

2.3.3. Side [P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N3)0.2]

A 50 mL round bottom flask with a reflux condenser and a magnetic stirrer was degassed and filled with argon. Side [P3HT0.6-(P3HT-5F)0.4] (400 mg, 1.72 mmol) was dissolved in 35 mL THF (dry) and NaN3 (22 mg, 0.34 mmol) was added. The mixture was degassed and flushed with argon again and heated at 40 °C for 48 h. The reaction mixture was then precipitated into a mixture of deionized water:methanol (1:1). The desired azide was obtained after filtration and the solid dried under vacuum at 30 °C for 24 h. Yield: 96% 1H NMR (δH; CDCl3; Me4Si): 7.00(broad, 1H), 2.80(broad, 2H), 2.64(broad, 2H), 1.65(broad, 4H), 1.53–1.22(broad, 12H), 0.88(broad, 6H), 19F NMR (δF; CDCl3; Me4Si): −135.76, −136.36, 15N NMR (δN; CDCl3; Me4Si): 233.84, 51.73.
The same procedure was used for the side [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N3)0.1].

2.3.4. Side [P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N-C60)0.2]

A 50 mL round bottom flask, equipped with a reflux condenser and a magnetic stirrer, was degassed (flamed under vacuum) and filled with argon. Side [P3HT0.6-(P3HT-5F-N3)0.4] (150 mg), C60 (87 mg, 0.12 mmol) and 60 mL o-DCB were added and the system was degassed and flushed with argon again. The reaction mixture was stirred at 140 °C for 72 h. After evaporation of the solvent, the solid was stirred in toluene for 72 h in order to remove the unreacted C60. The mixture was filtrated and the solid was evaluated via Thin Layer Chromatography (TLC), using toluene as the mobile phase on silica gel plates, and ATR for traces of unreacted C60 and if necessary the solid was again stirred in toluene. The final purified hybrid material was dried under vacuum at 50 °C overnight affording 100.00 mg of the hybrid C60 based polymer.

2.3.5. Side [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1]

A 50 mL round bottom flask, equipped with a reflux condenser and a magnetic stirrer, was degassed (flamed under vacuum) and filled with argon. Side [P3HT0.6-(P3HT-5F-N3)0.4] (150 mg), C70 (61 mg, 0.093 mmol) and 50 mL o-DCB were added and the system was degassed and flushed with argon again. The reaction mixture was stirred at 140 °C for 72 h. After evaporation of the solvent, the solid was stirred in toluene for 72 h in order to remove the unreacted C70. The mixture was filtrated and the solid was evaluated via TLC, using toluene as the mobile phase on silica gel plates, and ATR for traces of unreacted C70 in which case the solid was again stirred in toluene. The final purified material was dried under vacuum at 50 °C overnight affording 115.00 mg of the hybrid C70 based polymer.

3. Results and Discussion

3.1. Synthesis of Side-Br P3HTs

Typical GRIM polymerization conditions of 2,5-dibromo-3-hexyl-thiophene in freshly distilled THF using MeMgCl as the Grignard reagent and NiCl2(dppp) as catalyst were used to prepare the initial rr-P3HT [55]. The crude polymer was obtained after precipitation in methanol and was further purified via Soxhlet extractions using methanol, n-hexane and chloroform as fractionation solvents. For this work, we used only the chloroform fraction of the rr-P3HT with a corresponding Mn = 28,500, PDI = 1.2. For the side functionalization of rr-P3HT, bromination at the 4-position of the thiophene ring was performed [41,56,57,58,59] (Scheme 1). Various equivalents of the brominating reagent, N-Bromosuccinimide (NBS), in respect to the thiophene repeating units were used to achieve different bromination degrees of the polymeric backbone. The rate of bromine-functionalization is calculated from the 1H NMR of the side-[P3HT-(P3HT-Br)] (Figure 1). The appearance of a broadened bimodal peak at ~7–7.2 ppm and of two peaks at ~2.8–2.6 ppm is attributed to the successful bromination of the thiophene ring at 4-position and also the partial disruption of the regioregularity of the P3HT’s backbone. More specifically, the insertion of the bromine atom at the 4-position of the thiophene ring causes the stepwise reduction of the aromatic hydrogens of P3HT initially located at ~7 ppm and a stepwise shift of the α-methylene hydrogen peak of the hexyl group from 2.8 to ~2.6 ppm. The initial rr-P3HT shows a peak at 2.8 ppm corresponding to the α-methylene protons of the hexyl aliphatic chains at the 3-position of the thiophene ring connected in a head-to-tail mode. After bromination at the 4-position of the thiophene ring, via electrophilic substitution affording the side-[P3HT-(P3HT-Br)] polymers (Scheme 1), a new broad proton signal at about 2.6 ppm appears attributed to the α-methylene protons of those hexyl chains attached to the partially brominated thienyl ring. This chemical shift is down-field shifted compared to that of their initial counterparts, indicating the influence of the bromine as an electron-attracting group on the α-methylene protons. From the integral ratio of the peaks at 2.8–2.6 ppm to the one at ~7 ppm, the bromine substitution ratio in the side-[P3HT-(P3HT-Br)] polymers is calculated. In the ATR spectra (Figure 2), it is observed that as the ratio of the bromine-functionalization increases the characteristic peaks of the rr-P3HT are broadened, compared to the unfunctionalized rr-P3HT. The appearance of two new peaks at ~1600 and ~1720 cm−1, respectively, is attributed to the successful bromination of the thiophene ring at the 4-position [41].
As it is known from previous literature findings on the side functionalization of rr-P3HT [41,56,57,58,59], high levels of functionalization at the 4-position of the thiophene ring affect the planarity along the polymer’s backbone and also disrupt the effective conjugation length of P3HT. Additionally, high functionalization degrees with fullerene species along the rr-P3HT backbone, which is our herein ultimate target, were inevitably going to provide hybrid materials with poor solubility in common organic solvents. Thus, and in order to prepare processable hybrid materials, but with adequate functionalization levels and with efficient number of side-fullerene units to act as compatibilizers of the neat components’ blends, we chose the 40% side bromine-functionalized P3HT for further studies. Therefore, the side rr-[P3HT0.6-(P3HT-Br)0.4] reacted with an equivalent amount of perfluorophenyl boronic acid under Suzuki cross coupling conditions to form the respective side [P3HT0.6-(P3HT-5F)0.4] polymer (Scheme 2). The successful introduction of the perfluorophenyl ring at the side position of the P3HT’s polymeric backbone was proven through 19F NMR spectroscopy (Figure 3). The 19F NMR spectrum presents three peaks attributed to the three fluorine species of the perfluorinated ring. Afterwards, the perfluorophenyl groups of the P3HT0.6-(P3HT-5F)0.4 polymer were transformed into azides using sodium azide in THF [38]. Various azidation degrees were employed affording materials with 40%, 20% or 10% overall theoretical azidation degrees (Scheme 2, Table 1) in order to be able thereafter to incorporate different amounts of fullerenes and eventually produce processable hybrid polymers with C60 or C70. The successful azidation was confirmed using ATR and 15N NMR spectroscopies. The ATR spectrum of the 20% side perfluorophenyl azide showed a low intensity peak at 2064 cm−1 which demonstrates the azide formation (Figure 4), while for the 10% azide functionalized polymer the azide peak was not clearly observed due to the low ratio of the azide functionality over the total polymer. On the other hand, 15N NMR spectroscopy, as shown in Figure 5, revealed two clear peaks owing to the two nitrogen species of the azide functionality even for the 10% azide functionalized polymer.

3.2. Synthesis of Side-P3HT-(P3HT-5F) Hybrids

The side P3HT-perfluorophenyl azides have been used to prepare side hybrid semiconducting polymers with either C60 or C70 fullerenes. When the side [P3HT0.6-(P3HT-5F-N3)0.4] reacted with an equivalent amount of C60 to the azide content, the final hybrid polymer was insoluble in common organic solvents. In order to avoid such insolubility problems, the side [P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N3)0.2] reacted with C60 fullerene while the side [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N3)0.1] reacted with C70 providing the desired hybrid semiconducting polymers (Scheme 3) and leading to processable hybrid materials. After the [3+2] cycloaddition reaction, the side [P3HT-(P3HT-5F)]-hybrids with fullerenes were thoroughly washed several times with toluene to remove any unreacted fullerene species. A thorough ATR examination was performed for both hybrid materials after each subsequent washing step. When no further reduction of the fullerene peaks was noticed, we concluded that all unreacted traces of C60 or C70 were successfully removed (Figure 6).
Thermogravimetric analysis (TGA) of the side-functionalized [P3HT0.6-(P3HT-5F)0.4] and the hybrid with fullerene C70 (Figure 7) revealed at 800 °C a 44% residue for the [P3HT0.6-(P3HT-5F)0.4] and a 62% residue for the [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] hybrid, respectively. As expected, the hybrid polymer presented a higher residue due to the incorporation of the fullerene species.

3.3. Optical Properties

The optical properties of the synthesized hybrid materials compared to the initial side [P3HT0.6-(P3HT-5F)0.4] were investigated both in o-DCB solutions and in film form. The UV-Vis spectra of the initial side-[P3HT0.6-(P3HT-5F)0.4] both in o-DCB solution (Figure 8a) and in film form (Figure 8a) presented a blue shift compared to the unfunctionalized rr-P3HT. This phenomenon is attributed to the perturbation of the regioregularity of the P3HT backbone after the functionalization at the 4-position of the thiophene unit. The hybrid material [P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N-C60)0.2] presented an even greater blue-shift compared to its functionalized P3HT precursor, while the characteristic absorbance bands of the fullerene C60 were observed in solution as a shoulder at ~330 nm whereas, in film form, the absorption peak up to ~300 nm peak is attributed to the fullerenic part of the hybrid. On the other hand, the hybrid material [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] clearly presented the characteristic absorbance bands of C70 fullerene both in solutions and in film form in the region 300–400 nm.
The PL spectra of side-[P3HT0.6-(P3HT-5F)0.4] and its respective hybrids with C60 and C70 both in o-DCB solutions and in film form are shown in Figure 9. The o-DCB solutions of the hybrid materials, upon excitation at 470 nm, presented blue shifted spectra and broader photoluminescence curves in the region 500–650 nm, compared to the un-functionalized P3HT (Figure 9a). In film form, after excitation at the UV-Vis films maxima, the PL spectra followed the trend of the respective solutions also showing a blue shift compared to the neat rr-P3HT. A small quenching of the hybrids’ photoluminescence intensity in film form was observed, [P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N-C60)0.2] green-colored line and [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] blue-colored line of Figure 9b, that is attributed to the electron accepting fullerene parts interacting electronically with the electron donating P3HT polymeric backbone and the perturbation of the P3HT’s backbone due to the side functionalization.

3.4. Morphology Characterization

Thin film morphology of side perfluorophenyl functionalized [P3HT0.6-(P3HT-5F)0.4] and its respective hybrids with C60 and C70 was investigated using transmission electron spectroscopy (TEM) (Figure 10). The initial side [P3HT0.6-(P3HT-5F)0.4] formed a uniform thin film in the nanometer scale. The corresponding C60 based hybrid, [P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N-C60)0.2], formed films with large domains compared to the initial perfluorophenyl modified P3HT, with darker and brighter regions. In contrast, the C70 based hybrid polymer, [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1], presented a uniform nanophase separated thin film formation with well resolved small domains in analogy to its non hybrid initial polymer, [P3HT0.6-(P3HT-5F)0.4].
Moreover, a typical P3HT:PC70BM blend of 1:1 w/w ratio was employed to prepare two new blends with the [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] hybrid polymer. The hybrid material was added at a 20 wt % and at a 50 wt % percent, respectively, over the total weight of the P3HT:PC70BM initial blend. The effect of the additive on the morphology of the initial blend was investigated via TEM microscopy. Figure 11 presents the respective TEM images in 100 nm and in 20 nm scale, of the initial blend and of the blends containing the hybrid polymer without any thermal or other solvent vapor treatment. The P3HT:PC70BM blend presented uniform thin film with nanophase separation. The morphology of the blend containing 20% of the side C70 based hybrid polymer did not present any differentiations, even at the 20 nm scale, showing the probable stabilization of the initial blends morphology. On the other hand, the blend with the 50% of the side C70 hybrid polymer formed more heterogeneous films in which darker domains are present.

4. Conclusions

In this work, modification in various ratios of the rr-P3HT backbone in order to introduce the perfluorophenyl functionality at the 4-position of the thiophene unit has been efficiently performed. Thereafter, and based on our previous reported methodology, the perfluorophenyl moieties transformed into azides, from which hybrid materials with fullerenes, C60 or C70, have been synthesized aiming to develop efficient compatibilizers and stabilizers of P3HT:fullerene active layers typically employed in large area flexible OPVs. The side-functionalized P3HTs and their respective hybrid polymers have been characterized in respect to their optical and morphological properties. In solutions and in film form, all the side-functionalized materials presented blue shifted absorbance curves due to the perturbation of regioregularity of polythiophene. Regarding their photoluminescence in film form, the hybrid materials presented moderate photoluminescence quenching. The C70 based hybrid polymer formed uniform nanophase separated thin films while the C60 based hybrid formed more heterogeneous thin films exhibiting larger domains. In order to investigate their applicability as potential compatibilizers of polythiophene:fullerene BHJ active layers, selected hybrids were evaluated for their influence on the morphology of a typical P3HT:PC70BM active layer blend. More specifically, when the percentage of the hybrid was low compared to the total blend weight ratio, the morphology was not affected. However, when the hybrid polymer was added at a higher percentage, the final blend presented irregularities and large domains were created throughout the blend.
Based on these findings it can be concluded that the particular “perfluorophenyl”-“azide”-“carbon nanostructure hybrid” methodology is conveniently applicable for the development of side hybrid semiconducting polymer electron donors allowing the insertion of fullerene species in the desired percentage. Such hybrid polymer–fullerene materials can be employed as additives of typical polymer-donor:fullerene-acceptor active layer blends for OPVs.

Acknowledgments

The authors thank Mr. Dimitris Vachliotis of the Instrumental Analysis Laboratory, University of Patras for his efforts during the NMR experiments and Maria Kollia from the Laboratory of Electron Microscopy and Microanalysis at the University of Patras for the TEM images. This research has been financed by the “ARISTEIA” Action of the “Operational Programme Education and Lifelong Learning” co-funded by the European Social Fund (ESF) and National Resources) through the project “Design and Development of New Electron Acceptor Polymeric and Hybrid Materials and their Application in Organic Photovoltaics”—DENEA 2780.

Author Contributions

Sofia Kakogianni performed the experiments, analyzed data and wrote parts of the manuscript. Aikaterini K. Andreopoulou and Joannis K. Kallitsis supervised the work and wrote parts of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yu, J.; Zheng, Y.; Huang, J. Towards high performance organic photovoltaic cells: A review of recent development in organic photovoltaics. Polymers 2014, 6, 2473–2509. [Google Scholar] [CrossRef]
  2. Mazzio, K.A.; Luscombe, C.K. The future of organic photovoltaics. Chem. Soc. Rev. 2015, 44, 78–90. [Google Scholar] [CrossRef] [PubMed]
  3. Lu, L.; Zheng, T.; Wu, Q.; Schneider, A.M.; Zhao, D.; Yu, L. Recent advances in bulk heterojunction polymer solar cells. Chem. Rev. 2015, 115, 12666–12731. [Google Scholar] [CrossRef] [PubMed]
  4. Ma, W.; Yang, C.; Gong, X.; Lee, K.; Heeger, A.J. Thermally stable, efficient polymer solar cells with nanoscale control of the interpenetrating network morphology. Adv. Funct. Mater. 2005, 15, 1617–1622. [Google Scholar] [CrossRef]
  5. Choulis, S.A.; Kim, Y.; Nelson, J.; Bradley, D.D.C.; Giles, M.; Shkunov, M.; McCulloch, I. High ambipolar and balanced carrier mobility in regioregular poly(3-hexylthiophene). Appl. Phys. Lett. 2004, 85, 3890–3892. [Google Scholar] [CrossRef] [Green Version]
  6. Schafferhans, J.; Baumann, A.; Wagenpfahl, A.; Deibel, C.; Dyakonov, V. Oxygen doping of P3HT:PCBM blends: Influence on trap states, charge carrier mobility and solar cell performance. Org. Electron. 2010, 11, 1693–1700. [Google Scholar] [CrossRef]
  7. Dang, M.T.; Wantz, G.; Bejbouji, H.; Urien, M.; Dautel, O.J.; Vignau, L.; Hirsch, L. Polymeric solar cells based on P3HT:PCBM: Role of the casting solvent. Sol. Energy Mater. Sol. Cells 2011, 95, 3408–3418. [Google Scholar] [CrossRef]
  8. Baek, W.-H.; Yang, H.; Yoon, T.-S.; Kang, C.J.; Lee, H.H.; Kim, Y.-S. Effect of P3HT:PCBM concentration in solvent on performances of organic solar cells. Sol. Energy Mater. Sol. Cells 2009, 93, 1263–1267. [Google Scholar] [CrossRef]
  9. Carlé, J.E.; Krebs, F.C. Technological status of organic photovoltaics (OPV). Sol. Energy Mater. Sol. Cells 2013, 119, 309–310. [Google Scholar] [CrossRef] [Green Version]
  10. Kapnopoulos, C.; Mekeridis, E.D.; Tzounis, L.; Polyzoidis, C.; Zachariadis, A.; Tsimikli, S.; Gravalidis, C.; Laskarakis, A.; Vouroutzis, N.; Logothetidis, S. Fully gravure printed organic photovoltaic modules: A straightforward process with a high potential for large scale production. Sol. Energy Mater. Sol. Cells 2016, 144, 724–731. [Google Scholar] [CrossRef]
  11. Kallitsis, J.K.; Kourkouli, S.; Andreopoulou, A.K. Organic Polymeric Semiconductor Materials for Applications in Photovoltaic Cells Part 1: Properties and Materials, Handbook on Flexible Organic Electronics: Materials, Manufacturing and Applications; Logothetidis, S., Ed.; Woodhead Publishing Limited: Cambridge, UK, 2015; pp. 85–114. [Google Scholar]
  12. Scharber, M.C.; Mühlbacher, D.; Koppe, M.; Denk, P.; Waldauf, C.; Heeger, A.J.; Brabec, C.J. design rules for donors in bulk-heterojunction solar cells—Towards 10% energy-conversion efficiency. Adv. Mater. 2006, 18, 789–794. [Google Scholar] [CrossRef]
  13. Liu, X.; Chen, H.; Tan, S. Overview of high-efficiency organic photovoltaic materials and devices. Renew. Sustain. Energy Rev. 2015, 52, 1527–1538. [Google Scholar] [CrossRef]
  14. Gao, M.; Subbiah, J.; Geraghty, P.B.; Chen, M.; Purushothaman, B.; Chen, X.; Qin, T.; Vak, D.; Scholes, F.H.; Watkins, S.E.; et al. Development of a high-performance donor–acceptor conjugated polymer: Synergy in materials and device optimization. Chem. Mater. 2016, 28, 3481–3487. [Google Scholar] [CrossRef]
  15. Chen, J.; Cao, Y. Development of novel conjugated donor polymers for high efficiency bulkheterojunction photovoltaic devices. Acc. Chem. Res. 2009, 42, 1709–1718. [Google Scholar] [CrossRef] [PubMed]
  16. He, Y.; Li, Y. Fullerene derivative acceptors for high performance polymer solar cells. Phys. Chem. Chem. Phys. 2011, 13, 1970–1983. [Google Scholar] [CrossRef] [PubMed]
  17. He, Y.; Chen, H.-Y.; Hou, J.; Li, Y. Indene-C60 bisadduct: A new acceptor for high-performance polymer solar cells. J. Am. Chem. Soc. 2010, 132, 1377–1382. [Google Scholar] [CrossRef] [PubMed]
  18. He, Y.; Zhao, G.; Peng, B.; Li, Y. High-yield synthesis and electrochemical and photovoltaic properties of indene-C70 bisadduct. Adv. Funct. Mater. 2010, 20, 3383–3389. [Google Scholar] [CrossRef]
  19. Facchetti, A. Polymer donor–polymer acceptor (all-polymer) solar cells. Mater. Today 2013, 16, 123–132. [Google Scholar] [CrossRef]
  20. Bin, H.; Zhang, Z.-G.; Gao, L.; Chen, S.; Zhong, L.; Xue, L.; Yang, C.; Li, Y. Non-fullerene polymer solar cells based on alkylthio and fluorine substituted 2D-conjugated polymers reach 9.5% efficiency. J. Am. Chem. Soc. 2016, 138, 4657–4664. [Google Scholar] [CrossRef] [PubMed]
  21. Kim, Y.; Lim, E. Development of polymer acceptors for organic photovoltaic cells. Polymers 2014, 6, 382–407. [Google Scholar] [CrossRef]
  22. Brady, M.A.; Su, G.M.; Chabinyc, M.L. Recent progress in the morphology of bulk heterojunction photovoltaics. Soft Matter 2011, 7, 11065–11077. [Google Scholar] [CrossRef]
  23. Ruderer, M.A.; Muller-Buschbaum, P. Morphology of polymer-based bulk heterojunction films for organic photovoltaics. Soft Matter 2011, 7, 5482–5493. [Google Scholar] [CrossRef]
  24. Zhao, Y.; Xie, Z.; Qu, Y.; Geng, Y.; Wang, L. Solvent-vapor treatment induced performance enhancement of poly(3-hexylthiophene): Methanofullerene bulk-heterojunction photovoltaic cells. Appl. Phys. Lett. 2007, 90, 043504. [Google Scholar] [CrossRef]
  25. Moulé, A.J.; Meerholz, K. Controlling morphology in polymer–fullerene mixtures. Adv. Mater. 2008, 20, 240–245. [Google Scholar] [CrossRef]
  26. Yao, Y.; Hou, J.; Xu, Z.; Li, G.; Yang, Y. Effects of solvent mixtures on the nanoscale phase separation in polymer solar cells. Adv. Funct. Mater. 2008, 18, 1783–1789. [Google Scholar] [CrossRef]
  27. Angiolini, L.; Cocchi, V.; Lanzi, M.; Salatelli, E.; Tonelli, D.; Vlamidis, Y. Novel regioregular polythiophenes containing side-chain porphyrin groups for polymeric photovoltaic cells. Mat. Chem. Phys. 2014, 146, 464–471. [Google Scholar] [CrossRef]
  28. Bertho, S.; Campo, B.; Piersimoni, F.; Spoltore, D.; D’Haen, J.; Lutsen, L.; Maes, W.; Vanderzande, D.; Manca, J. Improved thermal stability of bulk heterojunctions based on side-chain functionalized poly(3-alkylthiophene) copolymers and PCBM. Sol. Energy Mater. Sol. Cells 2013, 110, 69–76. [Google Scholar] [CrossRef]
  29. Jaymand, M.; Hatamzadeh, M.; Omidi, Y. Modification of polythiophene by the incorporation of processable polymeric chains: Recent progress in synthesis and applications. Prog. Polym. Sci. 2015, 47, 26–69. [Google Scholar] [CrossRef]
  30. Lobez, J.M.; Andrew, T.L.; Bulović, V.; Swager, T.M. Improving the performance of P3HT–fullerene solar cells with side chain functionalized poly(thiophene) additives: A new paradigm for polymer design. ACS Nano 2012, 6, 3044–3056. [Google Scholar] [CrossRef] [PubMed]
  31. Wang, H.-J.; Chen, C.-P.; Jeng, R.-J. Polythiophenes comprising conjugated pendants for polymer solar cells: A review. Materials 2014, 7, 2411–2439. [Google Scholar] [CrossRef]
  32. Valderrama-García, B.X.; Rodríguez-Alba, E.; Morales-Espinoza, E.G.; Moineau Chane-Ching, K.; Rivera, E. Synthesis and characterization of novel polythiophenes containing pyrene chromophores: Thermal, optical and electrochemical properties. Molecules 2016, 21, 172. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, J.; Zhu, X.; Li, J.; Shen, J.; Tu, G. Enhancing the thermal stability of the bulk-heterojunction photovoltaics based on P3HT/PCBM by incorporating diblock amphipathic P3HT–PEO at D/A interface. RSC Adv. 2016, 6, 61934–61943. [Google Scholar] [CrossRef]
  34. Rumer, J.W.; McCulloch, I. Organic photovoltaics: Crosslinking for optimal morphology and stability. Mater. Today 2015, 18, 425–435. [Google Scholar] [CrossRef]
  35. Wantz, G.; Derue, L.; Dautel, O.; Rivaton, A.; Hudhomme, P.; Dagron-Lartigau, C. Stabilizing polymer-based bulk heterojunction solar cells via crosslinking. Polym. Int. 2014, 63, 1346–1361. [Google Scholar] [CrossRef] [Green Version]
  36. Li, F.; Yager, K.G.; Dawson, N.M.; Jiang, Y.-B.; Malloy, K.J.; Qin, Y. Nano-structuring polymer/fullerene composites through the interplay of conjugated polymer crystallization, block copolymer self-assembly and complementary hydrogen bonding interactions. Polym. Chem. 2015, 6, 721–731. [Google Scholar] [CrossRef]
  37. Huang, P.-T.; Chen, Y.-H.; Lin, B.-Y.; Chuang, W.-P. Homogenized poly(3-hexylthiophene)/methanofullerene film by addition of end-functionalized compatibilizer and its application to polymer solar cells. Int. J. Photoenergy 2015, 2015, 762532. [Google Scholar] [CrossRef]
  38. Kim, H.J.; Han, A.-R.; Cho, C.-H.; Kang, H.; Cho, H.-H.; Lee, M.Y.; Fréchet, J.M.J.; Oh, J.H.; Kim, B.J. Solvent-resistant organic transistors and thermally stable organic photovoltaics based on cross-linkable conjugated polymers. Chem. Mater. 2012, 24, 215–221. [Google Scholar] [CrossRef]
  39. Yuan, K.; Chen, L.; Chen, Y. Nanostructuring compatibilizers of block copolymers for organic photovoltaics. Polym. Int. 2014, 63, 593–606. [Google Scholar] [CrossRef]
  40. Kallitsis, J.K.; Anastasopoulos, C.; Andreopoulou, A.K. Functional semiconductors targeting copolymer architectures and hybrid nanostructures. MRS Commun. 2015, 5, 365–382. [Google Scholar] [CrossRef]
  41. Li, M.; Xu, P.; Yang, J.; Yang, S. Donor-p-acceptor double-cable polythiophenes bearing fullerene pendant with tunable donor/acceptor ratio: A facile postpolymerization. J. Mater. Chem. 2010, 20, 3953–3960. [Google Scholar] [CrossRef]
  42. Gholamkhass, B.; Peckham, T.J.; Holdcroft, S. Poly(3-hexylthiophene) bearing pendant fullerenes: Aggregation vs. self-organization. Polym. Chem. 2010, 1, 708–719. [Google Scholar] [CrossRef]
  43. Lanzi, M.; Salatelli, E.; Benelli, T.; Caretti, D.; Giorgini, L.; Di-Nicola, F.P. A regioregular polythiophene–fullerene for polymeric solar cells. J. Appl. Polym. Sci. 2015, 132. [Google Scholar] [CrossRef]
  44. Hufnagel, M.; Fischer, M.; Thurn-Albrecht, T.; Thelakkat, M. Influence of fullerene grafting density on structure, dynamics, and charge transport in P3HT-b-PPC61BM block copolymers. Macromolecules 2016, 49, 1637–1647. [Google Scholar] [CrossRef]
  45. Raïssi, M.; Erothu, H.; Ibarboure, E.; Cramail, H.; Vignau, L.; Cloutet, E.; Hiorns, R.C. Fullerene-capped copolymers for bulk heterojunctions: Device stability and efficiency improvements. J. Mater. Chem. A 2015, 3, 18207–18221. [Google Scholar] [CrossRef]
  46. Kourkouli, S.N.; Siokou, A.; Stefopoulos, A.A.; Ravani, F.; Plocke, T.; Müller, M.; Maultzsch, J.; Thomsen,, C.; Papagelis, K.; Kallitsis, J.K. Electronic properties of semiconducting polymer-functionalized single wall carbon nanotubes. Macromolecules 2013, 46, 2590–2598. [Google Scholar] [CrossRef]
  47. Stefopoulos, A.A.; Chochos, C.L.; Prato, M.; Pistolis, G.; Papagelis, K.; Petraki, F.; Kennou, S.; Kallitsis, J.K. Novel hybrid materials consisting of regioregular poly(3-octylthiophene)s covalently attached to single wall carbon nanotubes. Chem. Eur. J. 2008, 14, 8715–8724. [Google Scholar] [CrossRef] [PubMed]
  48. Stefopoulos, A.A.; Kourkouli, S.N.; Economopoulos, S.; Ravani, F.; Andreopoulou, A.; Papagelis, K.; Siokou, A.; Kallitsis, J.K. Polymer and hybrid electron accepting materials based on a semiconducting perfluorophenylquinoline. Macromolecules 2010, 43, 4827–4828. [Google Scholar] [CrossRef]
  49. Kakogianni, S.; Kourkouli, S.N.; Andreopoulou, A.K.; Kallitsis, J.K. A versatile approach for creating hybrid semiconducting polymer–fullerene architectures for organic electronics. J. Mater. Chem. A 2014, 2, 8110–8117. [Google Scholar] [CrossRef]
  50. Sygellou, L.; Kakogianni, S.; Andreopoulou, A.K.; Theodosiou, K.; Leftheriotis, G.; Kallitsis, J.K.; Siokou, A. Evaluation of the electronic properties of perfluorophenyl functionalized quinolines and their hybrids with carbon nanostructures. Phys. Chem. Chem. Phys. 2016, 18, 4154–4165. [Google Scholar] [CrossRef] [PubMed]
  51. Raja, R.; Liu, W.-S.; Hsiow, C.-Y.; Rwei, S.-P.; Chiu, W.-Y.; Wang, L. Terthiophene-C60 dyad as donor/acceptor compatibilizer for developing highly stable P3HT/PCBM bulk heterojunction solar cells. J. Mater. Chem. A 2015, 3, 14401–14408. [Google Scholar] [CrossRef]
  52. Perrin, L.; Legros, M.; Mercier, R. Design of a series of polythiophenes containing C60 groups: Synthesis and optical and electrochemical properties. Macromolecules 2015, 48, 323–336. [Google Scholar] [CrossRef]
  53. Yameen, B.; Puerckhauer, T.; Ludwig, J.; Ahmed, I.; Altintas, O.; Fruk, L.; Colsmann, A.; Barner-Kowollik, C. π-conjugated polymer–fullerene covalent hybrids via ambient conditions diels–alder ligation. Small 2014, 10, 3091–3098. [Google Scholar] [CrossRef] [PubMed]
  54. Kakogianni, S.; Lebedeva, M.A.; Paloumbis, G.; Andreopoulou, A.K.; Porfyrakis, K.; Kallitsis, J.K. Semiconducting end-perfluorinated P3HT–fullerenic hybrids as potential additives for P3HT/IC70BA blends. RSC Adv. 2016, 6, 98306–98316. [Google Scholar] [CrossRef]
  55. Loewe, R.S.; Ewbank, P.C.; Liu, J.; Zhai, L.; McCullough, R.D. Regioregular, head-to-tail coupled poly(3-alkylthiophenes) made easy by the GRIM method: Investigation of the reaction and the origin of regioselectivity. Macromolecules 2001, 34, 4324–4333. [Google Scholar] [CrossRef]
  56. Chochos, C.L.; Economopoulos, S.P.; Deimede, V.; Gregoriou, V.G.; Lloyd, M.T.; Malliaras, G.G.; Kallitsis, J.K. Synthesis of a soluble n-type cyano substituted polythiophene derivative: A potential electron acceptor in polymeric solar cells. J. Phys. Chem. C 2007, 111, 10732–10740. [Google Scholar] [CrossRef]
  57. Economopoulos, S.P.; Chochos, C.L.; Gregoriou, V.G.; Kallitsis, J.K.; Barrau, S.; Hadziioannou, G. Novel brush-type copolymers bearing thiophene backbone and side chain quinoline blocks. Synthesis and their use as a compatibilizer in thiophene-quinoline polymer blends. Macromolecules 2007, 40, 921–927. [Google Scholar] [CrossRef]
  58. Li, Y.N.; Vamvounis, G.; Holdcroft, S. Facile functionalization of poly(3-alkylthiophene)s via electrophilic substitution. Macromolecules 2001, 34, 141–143. [Google Scholar] [CrossRef]
  59. Li, Y.N.; Vamvounis, G.; Yu, J.; Holdcroft, S. A novel and versatile methodology for functionalization of conjugated polymers. Transformation of poly(3-bromo-4-hexylthiophene) via palladium-catalyzed coupling chemistry. Macromolecules 2001, 34, 3130–3132. [Google Scholar] [CrossRef]
Scheme 1. Synthetic procedure of side functionalization of rr-P3HT.
Scheme 1. Synthetic procedure of side functionalization of rr-P3HT.
Polymers 08 00440 sch001
Figure 1. 1H NMR spectra of the various side bromine functionalized P3HTs, in CDCl3 (the asterisks indicate solvent traces).
Figure 1. 1H NMR spectra of the various side bromine functionalized P3HTs, in CDCl3 (the asterisks indicate solvent traces).
Polymers 08 00440 g001
Figure 2. ATR spectra of an unfunctionalized P3HT and various ratios of side bromine-functionalized P3HTs.
Figure 2. ATR spectra of an unfunctionalized P3HT and various ratios of side bromine-functionalized P3HTs.
Polymers 08 00440 g002
Scheme 2. Synthetic procedure of side perfluorophenyl P3HT and their azide transformation; a indicates the azide functionalization degree.
Scheme 2. Synthetic procedure of side perfluorophenyl P3HT and their azide transformation; a indicates the azide functionalization degree.
Polymers 08 00440 sch002
Figure 3. 19F NMR spectra (in CDCl3) of the P3HT0.6-(P3HT-5F)0.4.
Figure 3. 19F NMR spectra (in CDCl3) of the P3HT0.6-(P3HT-5F)0.4.
Polymers 08 00440 g003
Figure 4. ATR spectra of the [P3HT0.6-(P3HT-5F)0.4] and azide [P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N3)0.2] (the asterisk indicates the azide functionality).
Figure 4. ATR spectra of the [P3HT0.6-(P3HT-5F)0.4] and azide [P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N3)0.2] (the asterisk indicates the azide functionality).
Polymers 08 00440 g004
Figure 5. 15N NMR spectra of the side [P3HT0.6-(P3HT-5F)0.3-P3HT-5F-N3)0.1], in CDCl3.
Figure 5. 15N NMR spectra of the side [P3HT0.6-(P3HT-5F)0.3-P3HT-5F-N3)0.1], in CDCl3.
Polymers 08 00440 g005
Scheme 3. Synthetic procedure of side-[P3HT0.6-(P3HT-5F)0.4] hybrids with C60 or C70 fullerenes.
Scheme 3. Synthetic procedure of side-[P3HT0.6-(P3HT-5F)0.4] hybrids with C60 or C70 fullerenes.
Polymers 08 00440 sch003
Figure 6. ATR spectra of: (a) C60; and (b) C70 based side [(P3HT0.6-P3HT-5F)0.4] hybrids, in comparison with the initial P3HT-(P3HT-5F) and the plain fullerenes, respectively.
Figure 6. ATR spectra of: (a) C60; and (b) C70 based side [(P3HT0.6-P3HT-5F)0.4] hybrids, in comparison with the initial P3HT-(P3HT-5F) and the plain fullerenes, respectively.
Polymers 08 00440 g006
Figure 7. TGA thermograms of rr-P3HT, [P3HT0.6-(P3HT-Br)0.4], [P3HT0.6-(P3HT-5F)0.4], [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] and C70 fullerene.
Figure 7. TGA thermograms of rr-P3HT, [P3HT0.6-(P3HT-Br)0.4], [P3HT0.6-(P3HT-5F)0.4], [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] and C70 fullerene.
Polymers 08 00440 g007
Figure 8. Normalized Absorption spectra in: (a) o-DCB solutions; and (b) film form of un functionalized P3HT, side functionalized [P3HT0.6-(P3HT-Br)0.4], [P3HT0.6-(P3HT-5F)0.4], [P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N-C60)0.2] and [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1], respectively. The inset in (a) shows the Normalized Absorption spectra of C60 and C70 in o-DCB solutions.
Figure 8. Normalized Absorption spectra in: (a) o-DCB solutions; and (b) film form of un functionalized P3HT, side functionalized [P3HT0.6-(P3HT-Br)0.4], [P3HT0.6-(P3HT-5F)0.4], [P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N-C60)0.2] and [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1], respectively. The inset in (a) shows the Normalized Absorption spectra of C60 and C70 in o-DCB solutions.
Polymers 08 00440 g008
Figure 9. Photoluminescence spectra of: (a) o-DCB solutions; and (b) in film form of un functionalized rr-P3HT, side functionalized [P3HT0.6-(P3HT-5F)0.4], [P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N-C60)0.2] hybrid and [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] hybrid, respectively.
Figure 9. Photoluminescence spectra of: (a) o-DCB solutions; and (b) in film form of un functionalized rr-P3HT, side functionalized [P3HT0.6-(P3HT-5F)0.4], [P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N-C60)0.2] hybrid and [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] hybrid, respectively.
Polymers 08 00440 g009
Figure 10. TEM images of the initial [P3HT0.6-(P3HT-5F)0.4] and its respective hybrids [P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N-C60)0.2] and [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] without thermal annealing.
Figure 10. TEM images of the initial [P3HT0.6-(P3HT-5F)0.4] and its respective hybrids [P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N-C60)0.2] and [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] without thermal annealing.
Polymers 08 00440 g010
Figure 11. TEM images of a P3HT:PC70BM 1:1 w/w blend and of P3HT:PC70BM:[P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] blends with a 20 wt % and a 50 wt % of the [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] hybrid, respectively. For the top images, the scale bars correspond to 100 nm while, for the bottom images, they correspond to 20 nm.
Figure 11. TEM images of a P3HT:PC70BM 1:1 w/w blend and of P3HT:PC70BM:[P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] blends with a 20 wt % and a 50 wt % of the [P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N-C70)0.1] hybrid, respectively. For the top images, the scale bars correspond to 100 nm while, for the bottom images, they correspond to 20 nm.
Polymers 08 00440 g011
Table 1. Azidation ratios of [P3HT0.6-(P3HT-5F)0.4] and the respective abbreviations of the functionalized polymers obtained.
Table 1. Azidation ratios of [P3HT0.6-(P3HT-5F)0.4] and the respective abbreviations of the functionalized polymers obtained.
Azide Functionalization Degree of (P3HT-5F-N3)ax-(P3HT-5F)x-(P3HT)1−xSide azide
1[P3HT0.6-(P3HT-5F-N3)0.4]
0.5[P3HT0.6-(P3HT-5F)0.2-(P3HT-5F-N3)0.2]
0.25[P3HT0.6-(P3HT-5F)0.3-(P3HT-5F-N3)0.1]

Share and Cite

MDPI and ACS Style

Kakogianni, S.; Andreopoulou, A.K.; Kallitsis, J.K. Synthesis of Polythiophene–Fullerene Hybrid Additives as Potential Compatibilizers of BHJ Active Layers. Polymers 2016, 8, 440. https://doi.org/10.3390/polym8120440

AMA Style

Kakogianni S, Andreopoulou AK, Kallitsis JK. Synthesis of Polythiophene–Fullerene Hybrid Additives as Potential Compatibilizers of BHJ Active Layers. Polymers. 2016; 8(12):440. https://doi.org/10.3390/polym8120440

Chicago/Turabian Style

Kakogianni, Sofia, Aikaterini K. Andreopoulou, and Joannis K. Kallitsis. 2016. "Synthesis of Polythiophene–Fullerene Hybrid Additives as Potential Compatibilizers of BHJ Active Layers" Polymers 8, no. 12: 440. https://doi.org/10.3390/polym8120440

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop