RESEARCH ARTICLE

Side chains and backbone structures influence on 4,7-dithien-2-yl-2,1,3-benzothiadiazole (DTBT)-based low-bandgap conjugated copolymers for organic photovoltaics

  • Debin NI 1 ,
  • Dong YANG 2 ,
  • Shuying MA 1 ,
  • Guoli TU , 1 ,
  • Jian ZHANG , 2
Expand
  • 1. Wuhan National Laboratory for Optoelectronics, School of Optical and Electronic Information, Huazhong University of Science and Technology, Wuhan 430074, China
  • 2. State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian National Laboratory of Clean Energy, Dalian 116023, China

Received date: 21 May 2013

Accepted date: 28 Aug 2013

Published date: 05 Dec 2013

Copyright

2014 Higher Education Press and Springer-Verlag Berlin Heidelberg

Abstract

Five 4,7-dithien-2-yl-2,1,3-benzothiadiazole (DTBT)-based conjugated copolymers with controlled molecular weight were synthesized to explore their optical, energy level and photovoltaic properties. By tuning the positions of hexyl side chains on DTBT unit, the DTBT-fluorene copolymers exhibited very different aggregation properties, leading to 60 nm bathochromic shift in their absorptions and the corresponding power conversion efficiencies (PCEs) value of photovoltaic cells varied from 0.38%, 0.69% to 2.47%. Different copolymerization units, fluorene, carbazole and phenothiazine were also investigated. The polymer based on phenothiazine exhibited lower PCE value due to much lower molecular weight owing to its poor solubility, although phenothiazine units were expected to be a better electron donor. Compared with the fluorene-based polymer, the carbazole-DTBT copolymer showed higher short circuit current density (Jsc) and PCE value due to its better intermolecular stacking.

Cite this article

Debin NI , Dong YANG , Shuying MA , Guoli TU , Jian ZHANG . Side chains and backbone structures influence on 4,7-dithien-2-yl-2,1,3-benzothiadiazole (DTBT)-based low-bandgap conjugated copolymers for organic photovoltaics[J]. Frontiers of Optoelectronics, 2013 , 6(4) : 418 -428 . DOI: 10.1007/s12200-013-0343-9

Introduction

Organic photovoltaics (OPVs) are a promising sustainable energy source with unique advantages, such as low-cost, light weight, and flexibility [1-4]. Bulk heterojunction active layer was usually a blend of an electron-donating conjugated polymer and an electron-accepting fullerene derivative, such as [6,6]-phenyl-C61-bulyric acid methyl ester (PC60BM) [5,6] or [6,6]-phenyl-C71-bulyric acid methyl ester (PC70BM) [7-10]. Poly (3-hexylthiophene) (P3HT) has been intensively investigated as the electron donor in organic/polymer photovoltaic cells with the power conversion efficiency (PCE) up to 6.5% [11-15]. In the past few years, a lot of works were focused on the developing the low-bandgap donor polymers to further increase the OPVs performance [16-28]. The variable structures of low-bandgap conjugated polymers provided a possibility toward increasing the short circuit current density (Jsc) and the open circuit voltage (Voc) [29-34] by modifying their highest occupied molecular orbital (HOMO) levels. 4,7-dithien-2-yl-2,1,3-benzothiadiazole (DTBT) has been comprehensively studied as building blocks in low-bandgap copolymers [35-41]. The thiophenes in the DTBT units could reduce the bandgap for the extending of the donor-acceptor conjugation and increasing of the planarity, as well as improve hole-transporting properties of the resulting copolymers [42]. poly{ [(2,7-(9-(2’-ethylhexyl)-9-hexyl-fluorene)-alt-[5,5-(-(4,7-di-2-thienyl-2,1,3-benzothiadiazole))}(PFDTBT) was introduced into OPVs as the electron donor polymer and exhibited a relatively high PCE of 2.2% when it was blended with PC60BM. The solubility of PFDTBT was very poor with the number average molecular weight (Mn) of 4800, which limited the PCE of the corresponding devices [43]. Then, docecyl side chains were introduced onto the 9-positions of fluorene units to enhance the solubility of the copolymer poly{[(2,7-(9,9-didodecyl-fluorene)-alt-[5,5-(-(4,7-di-2-thienyl-2,1,3-benzothiadiazole))} (DiD-PFDTBT) and a higher Mn (31,000) was achieved. However, the distinctly steric hindrance of the side chains decreased the π-π stacking and subsequently reduced the charge mobility of DiD-PFDTBT, leading to a relatively lower PCE (1.4%) than that of PFDTBT [44]. Unbulky side chains with methyl groups located on the third and seventh carbon of octyl were applied to the fluorene units to prepare poly{[2,7-(9,9-bis-(3,7-dimethyl-octyl)-fluorene)]-alt-[5,5-(4,7-di-2-thienyl-2,1,3-benzothiadiazole)]} Bis(DMO-PFDTBT), which not only increased the solubility of the polymer but also decreased the steric effect between polymer chains. The resulting polymer exhibited a PCE up to 4.5% when it was blended with PC70BM [45].
To overcome the poor solubility of DTBT-based polymers, side chains were also introduced onto thiophenes of DTBT units. It was found that the side chains on the DTBT units not only affected their molecular weight, solubility and geometry of the corresponding polymers, but also determined their absorptions, energy levels, and film morphologies. Octyl groups were introduced into thiophenes at 3- or 4-positions of the DTBT units respectively to investigate their influence. The substituted on 4-positions of thiophenes exhibited a less steric hindrance than that on 3-positions. The corresponding polymer of 4-positions showed a triple PCE value of that with octyl substituted on 3-positions [8]. Charge mobility of the DTBT based low-bandgap polymers were also depend on the alkyl chains positions on thiophene and the polymers with substitutes on 4-positions provided higher mobility for its stronger intermolecular stacking than that on 3-positions [46].
To improve the charge mobility of the DTBT based polymer, nitrogen atoms were introduced into the 9-positions of fluorenes which provided carbazole-based polymer. Poly[N-9′-heptadecanyl-2,7-carbazole-alt-5,5-(4′,7′-di-2-thienyl-2′,1′,3′-ben-zothiadiazole)] (PCDTBT) was investigated as the donor material for its fully aromatic, which provided a better chemical and environmental stability. PCDTBT showed a relatively high PCE value of 3.6% when blended with PC60BM [47], and the PCE could be further improved to 6.1% by using the titanium oxide (TiOx) layer as an optical spacer and hole blocker when the PC70BM was chosen as the electron acceptor [48].
According with the results of P3HT, molecular weight and polydispersity index (PDI) of the donor polymers are one of the key factors for the corresponding OPVs performances [49-53]. All the above studies of the DTBT-based polymers included both chemical structure and molecular effects. To obtain a clear outline how the chemical structures determined the device performances, it is urgent to exclude the influence of molecular weight when tuning the side chains and copolymerization units in the DTBT-based polymers, which will help to provide a general procedure for materials optimization. In this paper, five DTBT-based copolymers with comparable repeated units were prepared with varying copolymerization segments and hexyl side chains on thiophenes of DTBT units (Scheme 1). 3,7-dimethyl-octyls were chosen as the side chains on fluorenes or carbazoles for the copolymers without side chains substituted on DTBT to improve their solubility. All the absorptions, energy levels and the morphologies of resulting polymers, as well as their photovoltaic properties were studied with similar repeating units in the polymer backbone. Copolymer based on phenothiazine and DTBT were also employed. However, its molecular weight is much lower than the other four polymers for its poor solubility.
Fig.1 Scheme 1 Molecular structures of five DTBT based polymer

Full size|PPT slide

Experimental section

Materials

All reagents and solvents, unless otherwise specified, were obtained from Sigma-Aldrich or TCI chemical Co., and were used as received. Anhydrous tetrahydrofuran (THF) was distilled over sodium/diphenylketone under N2 prior to use. All manipulations involving air-sensitive reagents were performed under an atmosphere of dry nitrogen. The 2,2'-(9,9-dioctyl-9H-fluorene-2,7-diyl)bis(4,4,5,5- tetramethyl-1,-3,2-dioxaborolane)(1) [54], 4,7-bis(5-bromothiophen-2-yl)benzo-2,1,3-thiadiazole (2) [54], 4,7-bis(5-bromo-3-hexylthiophen-2-yl)benzo-2,1,3-thiadiazole (4) [55], 2,6,12,16-tetramethylheptadecan-9-ol [56], P2 [56], and P3 [45] were synthesized according to literature procedures.

Synthesis of the monomers

2,6,12,16-tetramethylheptadecan-9-yl 4-methylbenzenesulfonate (8)

P-toluenesulfonyl chloride (11.20 g, 58.75 mmol) in CH2Cl2 (50 mL) was added to the stirred solution of 2,6,12,16-tetramethylheptadecan-9-ol (12.18 g, 39.16 mmol), Et3N (97.9 mmol, 13.75 mL) and Me3N•HCl (3.74 g, 39.16 mmol) in CH2Cl2 (250 mL) in a 500 mL flask at 0-5°C, the mixture was stirred overnight. The organic phase was washed with water and brine, dried over magnesium sulfate and concentrated under reduced pressure. The crude product was purified by silica-gel column chromatography (hexanes/ethyl acetate= 10/1) to obtain colorless oil (15.6 g, yield: 81%). 1H NMR (400 MHz, CDCl3, ppm) δ: 7.79 (d, J = 8.1 Hz, 2H), 7.31 (d, J = 8.0 Hz, 2H), 4.51 (qt, J = 5.8 Hz, 1H), 2.43 (s, 3H), 1.55-1.49 (m, 4H), 1.12-1.09 (br, 32H), 0.86 (t, J = 6.6 Hz, 6H); 13C NMR (400 MHz, CDCl3, ppm): 144.24, 134.92, 129.61, 127.73, 85.30, 39.27, 36.97, 32.52, 31.74, 31.55, 27.94, 24.65, 22.64, 21.57, 19.42; m/z[m+] 483.6 ; Calcd: 484.0.

10-(2,6,12,16-tetramethylheptadecan-9-yl)-10H-phenothiazine (10)

The NaH (1.15 g, 48 mmol) was dispersed in the dry THF (100 mL) and was added into the solution of 10H-phenothiazine (4.42 g, 21.2 mmol) in the dry THF (150 mL). After the mixture was refluxed for 1 h with stirring, 2,6,12,16-tetramethylheptadecan-9-yl-4-methylbenzenesulfonate (15.6 g, 32 mmol) in 100 mL dry THF was dropped to the solution within 1 h, and the reaction was stirred for overnight under refluxing. The mixture was quenched with water (200 mL), and extracted three times with hexane (300 mL). The combined organic fraction was dried over magnesium sulfate, and the solvent was removed under reduced pressure. The crude product was purified by silica-gel column chromatography (hexane as eluent) resulting a yellow oil (6.3 g, yield: 61%). 1H NMR (400 MHz, CDCl3, ppm) δ: 7.09-7.05 (m, 4H), 6.91-6.81 (m, 4H), 3.58-3.54 (m, J = 5.8 Hz, 1H), 2.07 (m, 2H), 1.79 (m, 2H), 1.49-1.08 (m, 20H), 0.87-0.81 (m, J = 5.2 Hz, 18H); 13C NMR (400 MHz, CDCl3, ppm): 145.49, 127.32, 126.87, 126.71, 122.44, 117.64, 64.89, 39.33, 37.36, 36.80, 34.66, 32.76, 32.01, 28.02, 24.77, 22.73,19.58; m/z[m+] 494.8; Calcd: 494.3.

3,7-dibromo-10-(2,6,12,16-tetramethylheptadecan-9-yl)-10H-phenothiazine (11)

The solution of N-bromosuccinimide (17.00 g, 95.5 mmol) in dry dimethylformamide (DMF) (50 mL) was added dropwise into the mixture of 10-(2,6,12,16-tetramethylheptadecan-9-yl)-10H-phenothiazine (19.66 g, 39.8 mmol) in dry DMF (50 mL), the reaction was stirred overnight at room temperature, quenched with water, extracted with CH2Cl2 three times (450 mL), washed with brine, and dried over magnesium sulfate. The solvent was removed under reduced pressure and the crude residue was purified by silica-gel column chromatography (hexane as eluent) to give yellow color oil (17.09 g, yield: 66%). 1H NMR (400 MHz, CDCl3, ppm) δ: 7.20-7.16 (m, 4H), 6.72 (d, J = 5.8 Hz, 2H), 3.47 (m, J = 5.5 Hz, 1H), 1.98 (m, 2H), 1.81 (m, 2H), 1.78-1.03 (br, 20H), 0.83 (br, J = 6.6 Hz, 18H); 13C NMR (400 MHz, CDCl3, ppm): 144.30, 129.88, 129.57, 128.20, 118.76, 114.91, 65.32, 39.26, 37.26, 36.73, 34.59, 32.63, 31.91, 31.73, 27.98, 24.67,22.68, 19.97; m/z[m+] 651.2; Calcd: 651.2.

10-(henicosan-11-yl)-3,7-bis(4,4,5,5-tetramethyl-1,3,2-diocaborolan-2-yl)-10-H-Phenothiazine (3)

N-butyllithium (16.0 mL, 30 mmol, 2.5 mol/L in hexane) was added dropwise to the solution of compound 11 (6.52 g, 10 mmol) in dried THF (150 mL) in a 250 mL flask at -78°C. The mixture was stirred at -78°C for 1 h and 2-iso-propoxy-4,4,5,5-tetramethyl-1,3,2-dioxaborolane (8.1 mL, 40 mmol) was added rapidly to the solution. After additional one hour at -78°C, the resulting mixture was warmed to room temperature and stirred overnight. The mixture was poured into water, extracted with diethyl ether four times and dried over magnesium sulfate. The solvent was removed under reduced pressure and the residue was purified by silica-gel column chromatography (hexane as eluent, the silica–gel pretreatment with Et3N) to obtain the title product as a yellow powder solid. (2.1 g, yield: 30%). 1HNMR (400 MHz, C6D6, ppm) δ: 7.98 (s, 2H), 7.92 (d, J = 8.4 Hz, 2H), 6.95 (d, J = 8.2 Hz, 2H), 3.67 (m, J = 6.5 Hz, 1H), 1.99 (br, 2H), 1.83 (br, 2H), 1.48-1.07 (br, 44H), 0.79-0.72 (br, J = 6.0 Hz, 18H); 13C NMR (400 MHz, C6D6, ppm): 147.66, 133.74, 133.68, 125.74, 116.67, 83.63, 64.78, 39.25, 37.35, 36.76, 34.61, 32.63, 31.75, 27.94, 24.79, 22.68, 19.99, 19.48; m/z[m+] 745.7; Calcd: 745.2.

2,7-dibromo-9-(2,6,12,16-tetramethylheptadecan-9-yl)-9H-carbazole (9)

A three neck flask fitted with an addition funnel and a magnetic stirring bar was charged with 2,7-dibromo-carbazole (5.0 g, 10 mmol) and dried THF (150 mL). When the carbazole was completely dissolved, NaH (1.47 g, 61.5 mmol) was added to the solution, and refluxed for 1 h with stirring. Then the solution of compound 8 (15.0 g, 30.8 mmol) in dried THF (50 mL) was dropped into the mixture within 1 h, and refluxed overnight with stirring. The mixture was poured into distilled water (400 mL), and the aqueous layer was extracted with hexane three times, the combined organic fraction was dried with over magnesium sulfate and the solvent was removed under reduced pressure. The crude product was purified by column chromatography (silica-gel, hexane as eluent) resulting a colorless oil (7.5 g, yield: 78%). 1H NMR (400 MHz, CDCl3, ppm) δ: 7.88 (t, J = 8.6 Hz, 2H), 7.69 (s, 1H), 7.52 (s, 1H), 7.27 (br, 2H), 4.34 (m, J = 5.5 Hz, 1H), 2.16 (br, 2H), 1.91 (br, 2H), 1.13 (br, 20H), 0.81 (br, 18H); 13C NMR (400 MHz, CDCl3, ppm): 142.98, 139.44, 122.36, 121.47, 120.89, 119.89, 119.20, 114.51, 112.21, 57.48, 39.15, 36.73, 33.60, 32.54, 31.25, 30.66, 27.89, 19.64, 14.15; m/z[m+] 619.5; Calcd: 620.2.

2,7-bis(4,4,5,5-tetramethyl-1,3,2-dioxaborolan-2-yl)-9-(2,6,12,16-tetramethylheptadecan-9-yl)-9H-carbazole (2)

N-butyllithium (7.5 mL, 12 mmol, 2.5 mol/L in hexane) was added dropwise to a solution of compound 9 (3.0 g, 4.8 mmol) in dry THF (150 mL) in a 250 mL flask at -78°C .The mixture was stirred at -78°C for 1 h and 2-iso-propoxy-4,4,5,5-tetramethyl-1,3,2-dioxaborolane (2.96 mL, 14.5 mmol) was added rapidly to the solution. After additional one hour at -78°C, the resulting was warmed to room temperature and stirred overnight. The mixture was poured into water, extracted with diethyl ether four times and dried over magnesium sulfate, the organic phase was removed under reduced pressure and the residue was purified by silica-gel column chromatography (CH2Cl2/hexane= 10:1 as eluent) to obtain a white solid (0.85 g, yield: 25%). 1H NMR (400 MHz, CDCl3, ppm) δ: 8.11 (br, 2H), 8.03 (s, 1H), 7.86 (s, 1H), 7.65 (d, J = 7.4 Hz, 2H), 4.63 (m, 1H), 2.31 (br, 2H), 1.96 (br, 2H), 1.41 (br, 24H), 1.14 (br, 20H), 0.79 (br, 18H); 13C NMR (400 MHz, CDCl3, ppm): 141.99, 138.68, 126.02, 124.76, 124.72, 124.58, 119.97, 119.70, 118.20, 115.44, 56.83, 39.18, 36.72, 33.62, 31.40, 31.12, 27.86, 24.74, 22.55, 19.61, 14.22; m/z[m+] 713.6; Calcd: 714.2.

Synthesis of the polymer

Poly[(9,9-dioctyl-fluoren)-alt-4,7-bis(3-hexyl-thiophene-2-yl)benzo-2,1,3-thiadia-zole] (PFDTBT-in, P1) is presented in detail as a representative example

In a 50 mL flask, compound (1) (0.3213 g, 0.5 mmol), compound (4) (0.3132 g, 0.5 mmol), tris(dibenzylideneacetone)dipalladium (20 mg), and postassium carbonate (1.3891 g, 10 mmol) were dissolved in degassed toluene (5 mL) and degassed deionized water (5 mL). After 4 days at 80°C, the polymer was obtained and precipitated in methanol/water (20:1). The polymer was filtered and washed with ethyl acetate, hexane, dichloromethane, chloroform on Soxhlet apparatus. The chloroform fraction was dried and characterized (370 mg, yield: 84%). 1H NMR (400 MHz, CDCl3, ppm) δ: 7.75 (br, 6H), 7.65 (br, 2H), 7.43 (br, 2H), 2.77 (br, 4H), 2.08 (br, 4H), 1.75 (br, 4H), 1.36-1.11 (br, 36H), 0.87-0.73 (br, 12H).

Poly[(N-9'-2,6,12,-16-Tetramethyl-heptadecane-2,7-carbazole)-alt-5,5-(4',7'-di-2-thienyl-2',1',3′-benzothiadizaole)](TMHD-PCzDTBT, P4)

In a 50 mL flask, compound (2) (0.7137 g, 1 mmol), compound (5) (0.4582 g, 1 mmol), tris(dibenzylideneacetone)dipalladium (30 mg) and postassium carbonate (1.3891 g, 10 mmol) were dissolved in the mixture of degassed toluene (10 mL) and degassed deionized water (5 mL). After 50 h reaction at 80°C, the polymer was obtained and precipitated in methanol /water (20:1), filtered and washed on Soxhlet apparatus with ethyl acetate, hexane, dichloromethane, chloroform and o-dichlorobenzene. The o-dichlorobenzene fraction was dried and characterized (423 mg, yield: 49%). 1H NMR (400 MHz, CDCl3, ppm) δ: 8.21 (br, 1H), 8.12 (br, 2H), 7.93 (br, 2H), 7.81 (br, 2H), 7.79 (br, 2H), 7.73 (br, 1H), 7.61 (br, 2H), 7.51 (br, 2H), 7.17 (br, 2H), 4.65 (br, 1H), 2.38 (br, 2H), 2.07 (br, 2H), 1.43-1.03 (br, 10H), 0.84-0.78 (br, 18H).

Poly[(N-10'-2,6,12,16-tetramethylheptadecane-3,7-phenothiazine)-alt-5,5-(4',7'-di-2-thienyl-2',1',3′-benzothiadizaole)] (TMHD-PPDTBT, P5)

In a 50 mL flask, compound (3) (0.3729 g, 0.5 mmol), compound (5) (0.2291 g, 0.5 mmol), tris(dibenzylideneacetone)dipalladium (20 mg), postassium carbonate (1.3891 g, 10 mmol) were dissolved in degassed toluene (10 mL) and degassed deionized water (5 mL). After 5 days at 80°C, the polymer was obtained and precipitated in methanol /water (20:1), The polymer was filtered and washed with ethyl acetate, hexane, dichloromethane, chloroform on Soxhlet apparatus. The chloroform fraction was dried and characterized (270 mg, yield: 70%). 1H NMR (400 MHz, CDCl3, ppm) δ: 8.10 (br, 2H), 7.83 (br, 2H), 7.53 (br, 2H), 7.42 (br, 4H), 6.95 (br, 2H), 3.65 (br, 1H), 2.13 (br, 2H), 1.86 (br, 2H) 1.47-1.26 (br, 20H), 0.89-0.84 (br, 18H).

Measurements

1H-NMR and 13CNMR spectra were recorded by Bruker-AF301 at 400 MHz. Mass spectra were carried out on an Agilent (1100 LC/MSD Trap). UV-visible absorption spectra were measured using a Shimadzu UV-VIS-NIR Spectrophotometer (UV-3600). Solid films of UV-vis spectra measurement were spin-coated on quatrz plate from o-dichlorobenzene solution. Thermogravimetric analysis (TGA) were undertaken using a PerkinElmer Instruments (Pyris1 TGA) under nitrogen atmosphere at a heating rate of 10°C/min. Cyclic voltammetry measurements were carried out in a conventional three electrode cell using a glassy carbon as working electrode in acetonitrile solution of 0.1 mol/L of tetrabutylammoniumhexafluorophosphate (Bu4NPF6) as the supporting electrolyte at scan rate of 50 mV/s. In each case, a glassy carbon working electrode were coated with a thin layer of polymer, a platinum wire as the counter electrode, and a sliver wire as the quasi-reference electrode were used, and Ag/AgNO3 (0.1 mol/L) electrode were served as a reference electrode for all potentials quoted herein. The redox couple of ferrocene/ferrocenium ion (Fc/Fc+) was used as external standard. The corresponding HOMO levels were calculated using Eox/onset for experiments in solid films. Polymer films on indium tin oxide (ITO) anode electrode were formed by spin-coating method. Atomic force microscopy (AFM) was obtained in tapping-mode using Veeco (DIMENSION 3100). Size chromatography was performed on Waters Alliance GPCV2000 with a refractive index detector. Columns: Water Styvagel GE*1, Water Styvagel HMW GE*2. The eluent was 1,2,4-trichlorobenzene. The working temperature was 135°C, and the resolution time was 2 h. The concentration of the sample was 0.5 mg/mL, which was filtered (filter: 0.2 μm) prior to the analysis. The molecular weight was calculated according to calibration with polystyrene standards. Devices characterization were carried out under AM 1.5 G irradiation with the intensity of 100 mW/cm2 (Oriel 91160, 150 W) calibrated by a NREL certified standard silicon cell. Current versus potential (I-V) curves were recorded with a Keithley 2400 digital source meter. External quantum efficiencies (EQEs) were detected under monochromatic illumination (Oriel Cornerstone 260 1/4 m monochromatic equipped with Oriel 70613 NS QTH lamp), and the calibration of the incident light were performed with a monocrystalline silicon diode. The thickness of films were recorded by a profilometer (Alpha-Step 200, Tencor Instruments).

Fabrication of polymer photovoltaic cells

The structure of device was glass/ITO/PEDOT:PSS/Polymer:PCBM/Ca/Al. ITO-coated glass substrates were ultrasonically cleaned in detergent, deionized water, acetone, and isopropyl alcohol before the device fabrication. Then the substrates were treated with O2 plasma for 5 min, and PEDOT:PSS (Baytron PVP Al 4083) was spin-coated onto ITO-coated glass substrate, followed by annealing at 150°C for 30 min. The thickness of the PEDOT:PSS layer was about 50 nm, as determined by a Dektak 6 M surface profilometer. Then, the active layer was prepared by spin coating from the solution of polymer and PC60BM. The substrate was transfer into an evaporation chamber. A Ca layer (10 nm) was evaporated on the substrate under high vacuum (2×10-4 Pa), and then Al electrode (100 nm) was evaporated on the Ca layer under high vacuum as well. Post device annealing was carried out inside a nitrogen-filled glove box. The current-voltage characteristics of the photovoltaic cells were measured using a Keithley 2400 I-V measurement system. The measurement were conducted under the irradiation of AM 1.5 G simulated solar light (100 mW/cm2), light intensity was adjusted by using a standard silicon photovoltaic cells with an optical filter.

Results and discussion

Synthesis and characterization of the polymers

The synthesis of all the monomers and polymers are outlined in Scheme 2. Polymerizations were carried out by a Suzuki polymerization using the Pd(PPh3)4 as catalyst at 80°C for varying days. The Mn and PDI of copolymers were determined by high temperature gel permeation chromatography (GPC) with dichlorobenzene as eluent and polystyrene as a standard, and shown in Table 1. The Mn of P5 is 8600 that is much lower than other four polymers, which can be attributed to its poor solubility in toluene and precipitated during polymerization. The decomposition temperatures of these copolymers were investigated by thermogravimetric analysis (TGA) and shown in Fig. 1. P1, P2 and P3 exhibited 5% decomposition temperature over 400°C. P4 and P5 showed relatively low decomposition temperature (Td) about 330°C. Neither obvious glass transition nor melting transition was observed for all the copolymers, indicating that the amorphous nature of the obtain polymers.
Fig.2 Scheme 2 Synthetic procedures of monomers and polymers

Full size|PPT slide

Fig.3 TGA curves of five DBTB based copolymer at a scan rate of 10°C/min under a nitrogen atmosphere

Full size|PPT slide

Optical and electrochemical properties of the polymers

The UV-vis absorption spectra of polymer films were investigated, as shown in Fig. 2 and Table 1. All the copolymers exhibited two absorption peaks. The long wavelength absorption bands could be attributed to the intramolecular charge transfer (ICT). Compared with the ICT absorption of P1 at 486 nm, the absorption of P2 showed 44 nm red shift, which could be attributed to strong intermolecular stacking, indicating that alkyl side chains at the 4-positions of thiophenes displayed much less steric hindrance than at the 3-positions and leading to a red-shift absorption of P2. For the polymer without hexyl side chains on DTBT units (P3), its long wavelength absorption located at 550 nm and exhibited 18 nm red-shift than P2, indicating stronger intermolecular π-stacking in P3 film. According to the above absorption spectra comparison of P1, P2, and P3, it clearly showed that alkyl substituted and its positions on DTBT units displayed strong steric effect and determined the intermolecular stacking. For P4, in which fluorene units were replaced by carbazoles, showed maxima absorptions at 400 and 562 nm respectively. There was 12 nm red shift of long wavelength absorption from P3 to P4, which may be attributed a stronger ICT for the atom replacing or better intermolecular π-stacking. P5 exhibited two absorption peaks situated at 390 and 564 nm. The long wavelength absorption was almost the same as P4 although the sulfur atom in phenothiazine was expected to show a strong electron donating ability. The reason may be originated from less flatness in phenothiazine than carbazole, and resulted in decreased intermolecular interaction.
The electrochemical properties of the polymers were characterized by cyclic voltammetry (CV) at room temperature. The CV curves were recorded referenced to an Ag/Ag+ (0.01 mol/L of AgNO3 in acetonitrile) electrode, which was calibrated against the ferrocene/ferrocenium (Fc/Fc+) redox couple (4.74 eV) below the vacuum level. The HOMO levels of the copolymers were obtained from the onset of oxidation potentials (Eox,onset), as shown in Fig. 3. The LUMO levels of these five copolymers were about at -3.4~-3.5 eV, which was mainly dominated by the BT units (Table 1).
Tab.1 Yield, molecular weight, PDI, and thermal properties, opticals and electrochemical properties of the polymers
Mna)
/(kg∙mol-1)
PDIrepeat
unit
Tdb)
/°C
lmax
/nm
Egopt c)
/eV
HOMO d)
/eV
LUMO e)
/eV
P126.11.6930438.9374/4862.10-5.53-3.43
P228.81.7133443.6372/5301.98-5.47-3.49
P318.81.4025444.7390/5501.93-5.44-3.51
P416.62.4922327.2400/5621.85-5.44-3.59
P58.61.9311339.3390/5641.81-5.18-3.37

Notes: a) Determined by the GPC in dichlorobenzene using polystyrene as standards. Mn: number-average molecular weight; PDI: polydispersity index (Mw/Mn);

b) Temperature at 5% weight loss determined by TGA at a heating rate of 10°C/min under nitrogen;

c) Calculated from the absorption band edge of the copolymer film, Egopt = 1240/ledge;

d) Thin films in CH3CN/n-Bu4NPF6, versus ferrocenium/ferrocene at 50 mV/s . HOMO value was estimated from the onset oxidation potential, and e) LUMO value was estimated from LUMO= HOMO-Egopt

Fig.4 UV-vis absorption spectra of the thin film of the polymers

Full size|PPT slide

Calculated from CV, the energy gap of P2 was raised 0.12 eV than that of P1. The reason was that the alkyl chains were introduced at the 3-positions of the thienyl group, apparently reduced the electron delocalization and rendered a lower HOMO and higher LUMO energy levels. The HOMO level of P3 is -5.44 eV and its LUMO level was calculated from the bandgap results based on its absorption spectrum. P4 exhibited a similar HOMO level to P3, which indicated the carbazole units did not exhibit much stronger electron donating ability than fluorenes and the ICT absorption in the two copolymers were mainly contributed by the two thienyl groups in DTBT. So its 12 nm red-shift ICT absorption from P3 to P4 mainly originated from much stronger intermolecular aggregation in P4, which might result better π-π stacking and higher charge mobility. The HOMO and LUMO levels of P5 was about 0.2 eV higher than those of P4 for a stronger ICT from phenothiazine to 2,1,3-benzothiadiazole due to the existence of sulfur atoms. All the polymers expect for P5 had a relative low HOMO levels, indicating that the devices would have a relative large Voc, which was proved by the results of the photovoltaic device performances.

Photovoltaic characteristics of the polymer thin films

OPV devices were fabricated with the five polymers as the electron donors and the PC60BM as the electron acceptor to explore the effect of the side chains and backbone on the properties of DTBT based copolymer photovoltaic cells. The structure of devices were ITO/PEDOT:PSS (50 nm)/polymer:PC60BM/ Ca (10 nm) /Al (100 nm).
Fig.5 Cyclic voltammetry of polymers film coated on a glass carbon electrode in Bu4NPF4/CH3CN solution

Full size|PPT slide

The solutions of polymer/PC60BM were spin-coated on PEDOT:PSS modified ITO glass with of 1,2-dichlorobenzene as the solvent. No additive was applied to simplify the comparison. Ratios of the different polymer/PC60BM had been optimized from 1:1, 1:2, 1:3, 1:4, and the best ratio was chosen as shown in Table 2. The thickness of the active layers, around 60-70 nm, was also optimized for all polymer films. The resulting films were dried under vacuum at room temperature and then annealed for 15 min. The optimized annealing temperatures for the five polymers were listed in Table 2. Then cathodes were deposited onto active layers at 2 × 10-4 Pa with a metal mask. The J-V characteristics of photovoltaic cells were shown in Fig. 4 and the Jsc, fill factor (FF), and PCE of the devices were summarized in Table 2.
Compared with devices of P1, the devices of P2 showed doubled Jsc and PCE value, which is consistent with the result that P2 had stronger intermolecular aggregation than P1 for the different position of hexyl. And due to the same reason, P3 showed much higher Jsc and PCE than P2. When the fluorene units were replaced by carbazoles, P4 exhibited a little higher PCE than that of P3, which originated from the increased Jsc although its Voc and FF were lower than those of P3. The increased Jsc in devices of P4 was contributed by its strong intermolecular stacking in films. P5 showed much lower PCE than P4 for the reasons of poor solubility and Jsc. All the devices had shown a relatively higher Voc value from 0.76 to 0.87 V except for P5. According to the above results, the PCE of the five DTBT-based polymers were mainly determined by the Jsc of the devices, as shown in Table 2.
Fig.6 J-V characteristics of photovoltaic cells of five polymers

Full size|PPT slide

To study the contribution of polymer absorption to Jsc, the external quantum efficiency (EQE) of device based on the copolymers were measured under the illumination of monochromatic light as showed in Fig. 5. The maximum EQEs of P3 and P4 were much higher than those of P1 and P2, agreeing with the devices results.
Fig.7 External quantum efficiency (EQE) of the PSCs based on five polymers

Full size|PPT slide

Morphology

The nanoscale morphologies for polymer/PCBM films were studied using tapping-mode atomic force microscopy (AFM). Phase images (top) and surface topography (under) were taken for each film and shown in Fig. 6. Distinctly different morphologies of polymer:PCBM blend films were observed in their phase images, indicating that the active layer morphology was strongly affected by the structures of the polymers. The root mean square roughnesses (Rrms) of these five polymers were 0.26, 0.32, 0.49, 0.45 and 0.26 nm respectively, which indicated that the stronger intermolecular stacking would exhibit large size of phase separations and Rrms. This agreed with corresponding devices performances.
Tab.2 Optimized device characteristics of five DTBT based copolymers
polymer: PC60BM
(weight ratio)
annealing temperature
/°C
thickness
/nm
Jsc
/(mA∙cm-2)
Voc
/V
FFPCE
/%
P11:460741.650.820.280.38
P21:370713.510.870.290.69
P31:3130736.670.860.432.47
P41:280668.650.760.402.60
P51:2100562.980.570.320.54
Fig.8 Fig. 6 AFM (5 μm × 5 μm) topography and phase images of five polymers

Full size|PPT slide

Conclusions

Five DTBT based copolymer with controlled Mn were synthesized and characterized to explore the influence of side chains and backbone structure on the properties of absorption spectrums, energy levels, and photovoltaic devices. Strong steric hindrance would be produced when side chains were introduced onto the thiophenes in DTBT units, which also exhibited position dependence when modifying the side chains from 3- to 4-positions. For the polymer without alkyl substituted on DTBT, the polymer showed stronger intermolecular stacking and lower bandgap. When the copolymerization units were changed from fluorene to carbazole, the red-shift ICT absorption in carbazole copolymer was mainly attributed to its better π-π stacking. And for all the five DTBT-based copolymers, the PCE of their corresponding devices mainly depended on their Jsc, which were determined by the intermolecular stacking of the donor polymers and had also been proved by their morphology results.

Acknowledgements

Financial support by the National Natural Science Foundation of China (Grant Nos. 51073063 and 20904057) and Open Project of State Key Laboratory for Supramolecular Structure and Materials (No. SKLSSM201129) of Jilin university are gratefully acknowledged. J. Zhang thanks the support by 100 Talents Programme of the Chinese Academy of Science.
1
Yu G, Gao J, Hummelen J C, Wudl F, Heeger A J. Polymer photovoltaic cells: enhanced efficiencies via a network of internal donor-acceptor heterojunctions. Science, 1995, 270(5243): 1789–1791

DOI

2
Brabec C J, Sariciftci N S, Hummelen J C. Plastic solar cells. Advanced Functional Materials, 2001, 11(1): 15–26

DOI

3
Coakley K M, McGehee M D. Conjugated polymer photovoltaic cells. Chemistry of Materials, 2004, 16(23): 4533–4542

DOI

4
Günes S, Neugebauer H, Sariciftci N S. Conjugated polymer-based organic solar cells. Chemical Reviews, 2007, 107(4): 1324–1338

DOI PMID

5
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. Advanced Functional Materials, 2005, 15(10): 1617–1622

DOI

6
Reyes-Reyes M, Kim K, Carroll D L. High-efficiency photovoltaic devices based on annealed poly(3-hexylthiophene) and 1-(3-methoxycarbonyl)-propyl-1-phenyl-(6,6)C61blends. Applied Physics Letters, 2005, 87(8): 083506–083508

DOI

7
Qin R P, Li W W, Li C H, Du C, Veit C, Schleiermacher H F, Andersson M, Bo Z, Liu Z P, Inganäs O, Wuerfel U, Zhang F L. A planar copolymer for high efficiency polymer solar cells. Journal of the American Chemical Society, 2009, 131(41): 14612–14613

DOI PMID

8
Peet J, Kim J Y, Coates N E, Ma W L, Moses D, Heeger A J, Bazan G C. Efficiency enhancement in low-bandgap polymer solar cells by processing with alkane dithiols. Nature Materials, 2007, 6(7): 497–500

DOI PMID

9
Thompson B C, Fréchet J M J. Polymer-fullerene composite solar cells. Angewandte Chemie International Edition, 2007, 47(1): 58–77

DOI PMID

10
Kim Y, Cook S, Tuladhar S M, Choulis S A, Nelson J, Durrant J R, Bradley D D C, Giles M, McCulloch I, Ha C S, Ree M. A strong regioregularity effect in self-organizing conjugated polymer films and high-efficiency polythiophene:fullerene solar cells. Nature Materials, 2006, 5(3): 197–203

DOI

11
Li G, Shrotriya V, Huang J S, Yao Y, Moriarty T, Emery K, Yang Y. High-efficiency solution processable polymer photovoltaic cells by self-organization of polymer blends. Nature Materials, 2005, 4(11): 864–868

DOI

12
Shi C J, Yao Y, Yang Y, Pei Q B. Regioregular copolymers of 3-alkoxythiophene and their photovoltaic application. Journal of the American Chemical Society, 2006, 128(27): 8980–8986

DOI PMID

13
Campoy-Quiles M, Ferenczi T, Agostinelli T, Etchegoin P G, Kim Y, Anthopoulos T D, Stavrinou P N, Bradley D D C, Nelson J. Morphology evolution via self-organization and lateral and vertical diffusion in polymer:fullerene solar cell blends. Nature Materials, 2008, 7(2): 158–164

DOI PMID

14
Zhao G J, He Y J, Li Y F. 6.5% efficiency of polymer solar cells based on poly(3-hexylthiophene) and indene-C60 bisadduct by device optimization. Advanced Materials, 2010, 22(39): 4355–4358

DOI PMID

15
Chang C Y, Wu C E, Chen S Y, Cui C H, Cheng Y J, Hsu C S, Wang Y L, Li Y F. Enhanced performance and stability of a polymer solar cell by incorporation of vertically aligned, cross-linked fullerene nanorods. Angewandte Chemie International Edition, 2011, 50(40): 9386–9390

16
He F, Wang W, Chen W, Xu T, Darling S B, Strzalka J, Liu Y, Yu L P. Tetrathienoanthracene-based copolymers for efficient solar cells. Journal of the American Chemical Society, 2011, 133(10): 3284–3287

DOI PMID

17
Piliego C, Holcombe T W, Douglas J D, Woo C H, Beaujuge P M, Fréchet J M J. Synthetic control of structural order in N-alkylthieno[3,4-c]pyrrole-4,6-dione-based polymers for efficient solar cells. Journal of the American Chemical Society, 2010, 132(22): 7595–7597

DOI PMID

18
Chen H Y, Hou J H, Zhang S Q, Liang Y Y, Yang G W, Yang Y, Yu L P, Wu Y, Li G. Polymer solar cells with enhanced open-circuit voltage and efficiency. Nature Photonics, 2009, 3(11): 649–653

DOI

19
Price S C, Stuart A C, Yang L Q, Zhou H X, You W. Fluorine substituted conjugated polymer of medium band gap yields 7% efficiency in polymer-fullerene solar cells. Journal of the American Chemical Society, 2011, 133(12): 4625–4631

DOI PMID

20
Kim J Y, Lee K, Coates N E, Moses D, Nguyen T Q, Dante M, Heeger A J. Efficient tandem polymer solar cells fabricated by all-solution processing. Science, 2007, 317(5835): 222–225

DOI PMID

21
Wang E, Hou L T, Wang Z Q, Hellström S, Zhang F L, Inganäs O, Andersson M R. An easily synthesized blue polymer for high-performance polymer solar cells. Advanced Materials, 2010, 22(46): 5240–5244

DOI PMID

22
Amb C M, Chen S, Graham K R, Subbiah J, Small C E, So F, Reynolds J R. Dithienogermole as a fused electron donor in bulk heterojunction solar cells. Journal of the American Chemical Society, 2011, 133(26): 10062–10065

DOI PMID

23
Chang C Y, Wu C E, Chen S Y, Cui C H, Cheng Y J, Hsu C S, Wang Y L, Li Y F. Enhanced performance and stability of a polymer solar cell by incorporation of vertically aligned, cross-linked fullerene nanorods. Angewandte Chemie International Edition, 2011, 50(40): 9386–9390

DOI PMID

24
Jin J K, Choi J K, Kim B J, Kang H B, Yoon S C, You H, Jung H T. Synthesis and photovoltaic performance of low-bandgap polymers on the basis of 9,9-dialkyl-3,6-dialkyloxysilafluorene. Macromolecules, 2011, 44(3): 502–511

DOI

25
Peng Q, Liu X J, Su D, Fu G W, Xu J, Dai L M. Novel benzo[1,2-b:4,5-b’]dithiophene-benzothiadiazole derivatives with variable side chains for high-performance solar cells. Advanced Materials, 2011, 23(39): 4554–4558

DOI PMID

26
Huo L J, Guo X, Zhang S Q, Li Y F, Hou J H. PBDTTTZ: a broad band gap conjugated polymer with high photovoltaic performance in polymer solar cells. Macromolecules, 2011, 44(11): 4035–4037

DOI

27
Dou L T, You J B, Yang J, Chen C C, He Y J, Murase S, Moriarty T, Emery K, Li G, Yang Y. Tandem polymer solar cells featuring a spectrally matched low-bandgap polymer. Nature Photonics, 2012, 6(3): 180–185

DOI

28
Li G, Zhu R, Yang Y. Polymer solar cells. Nature Photonics, 2012, 6(3): 153–161

DOI

29
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. Advanced Materials, 2006, 18(6): 789–794

DOI

30
Blouin N, Michaud A, Gendron D, Wakim S, Blair E, Neagu-Plesu R, Belletête M, Durocher G, Tao Y, Leclerc M. Toward a rational design of poly(2,7-carbazole) derivatives for solar cells. Journal of the American Chemical Society, 2008, 130(2): 732–742

DOI PMID

31
Huo L J, Hou J H, Chen H Y, Zhang S Q, Jiang Y, Chen T L, Yang Y. Bandgap and molecular level control of the low-bandgap polymers based on 3,6-dithiophen-2-yl-2,5-dihydropyrrolo[3,4-c]pyrrole-1,4-dione toward highly efficient polymer solar cells. Macromolecules, 2009, 42(17): 6564–6571

DOI

32
Liang Y Y, Feng D Q, Wu Y, Tsai S T, Li G, Ray C, Yu L P. Highly efficient solar cell polymers developed via fine-tuning of structural and electronic properties. Journal of the American Chemical Society, 2009, 131(22): 7792–7799

DOI PMID

33
Zoombelt A P, Fonrodona M, Wienk M M, Sieval A B, Hummelen J C, Janssen R A J. Photovoltaic performance of an ultrasmall band gap polymer. Organic Letters, 2009, 11(4): 903–906

DOI PMID

34
Mondal R, Ko S, Norton J E, Miyaki N, Becerril H A, Verploegen E, Toney M F, Bredas J L, McGehee M D, Bao Z N. Molecular design for improved photovoltaic efficiency: band gap and absorption coefficient engineering. Journal of Materials Chemistry, 2009, 19(39): 7195–7197

DOI

35
Dhanabalan A, Van Duren J K J, Van Hal P A, Van Dongen J L J, Janssen R A J. Synthesis and characterization of a low bandgap conjugated polymer for bulk heterojunction photovoltaic cells. Advanced Functional Materials, 2001, 11(4): 255–262

DOI

36
Boudreault P L T, Michaud A, Leclerc M. A new poly(2,7-Dibenzosilole) derivative in polymer solar cells. Macromolecular Rapid Communications, 2007, 28(22): 2176–2179

DOI

37
Song S, Jin Y, Kim S H, Moon J, Kim K, Kim J Y, Park S H, Lee K, Suh H. Stabilized polymers with novel indenoindene backbone against photodegradation for LEDs and solar cells. Macromolecules, 2008, 41(20): 7296–7305

DOI

38
Moulé A J, Tsami A, Bünnagel T W, Forster M, Kronenberg N M, Scharber M, Koppe M, Morana M, Brabec C J, Meerholz K, Scherf U. Two novel cyclopentadithiophene-based alternating copolymers as potential donor components for high-efficiency bulk-heterojunction-type solar cells. Chemistry of Materials, 2008, 20(12): 4045–4050

DOI

39
Liao L, Dai L M, Smith A, Durstock M, Lu J P, Ding J F, Tao Y. Photovoltaic-active dithienosilole-containing polymers. Macromolecules, 2007, 40(26): 9406–9412

DOI

40
Zhou E, Nakamura M, Nishizawa T, Zhang Y, Wei Q S, Tajima K, Yang C H, Hashimoto K. Synthesis and photovoltaic properties of a novel low band gap polymer based on N-substituted dithieno[3,2-b:2′,3′-d]pyrrole. Macromolecules, 2008, 41(22): 8302–8305

DOI

41
Wang M, Hu X W, Liu P, Li W, Gong X, Huang F, Cao Y. Donor-acceptor conjugated polymer based on naphtho[1,2-c:5,6-c]bis[1,2,5]thiadiazole for high-performance polymer solar cells. Journal of the American Chemical Society, 2011, 133(25): 9638–9641

DOI PMID

42
Zhou H X, Yang L Q, Xiao S Q, Liu S B, You W. Donor-acceptor polymers incorporating alkylated dithienylbenzothiadiazole for bulk heterojunction solar cells: pronounced effect of positioning alkyl chains. Macromolecules, 2009, 43(2): 811–820

DOI

43
Svensson M, Zhang F, Veenstra S C, Verhees W J H, Hummelen J C, Kroon J M, Inganäs O, Andersson M R. High-performance polymer solar cells of an alternating polyfluorene copolymer and a fullerene derivative. Advanced Materials, 2003, 15(12): 988–991

DOI

44
Inganäs O, Svensson M, Zhang F, Gadisa A, Persson N K, Wang X, Andersson M R. Low bandgap alternating polyfluorene copolymers in plastic photodiodes and solar cells. Applied Physics A, 2004, 79(1): 31–35

DOI

45
Chen M H, Hou J H, Hong Z, Yang G W, Sista S, Chen L M, Yang Y. Efficient polymer solar cells with thin active layers based on alternating polyfluorene copolymer/fullerene bulk heterojunctions. Advanced Materials, 2009, 21(42): 4238–4242

DOI

46
Lee S K, Cho S, Tong M, Seo J H, Heeger A J. Effects of substituted side-chain position on donor–acceptor conjugated copolymers. Journal of Polymer Science Part A: Polymer Chemistry, 2011, 49(8): 1821–1829

DOI

47
Blouin N, Michaud A, Leclerc M. A low-bandgap poly(2,7-carbazole) derivative for use in high-performance solar cells. Advanced Materials, 2007, 19(17): 2295–2300

DOI

48
Park S H, Roy A, Beaupre S, Cho S, Coates N, Moon J S, Moses D, Leclerc M, Lee K, Heeger A J. Bulk heterojunction solar cells with internal quantum efficiency approaching 100%. Nature Photonics, 2009, 3(5): 297–302

DOI

49
Kline R J, McGehee M D, Kadnikova E N, Liu J S, Fréchet J M J, Toney M F. Dependence of regioregular poly(3-hexylthiophene) film morphology and field-effect mobility on molecular weight. Macromolecules, 2005, 38(8): 3312–3319

DOI

50
Schilinsky P, Asawapirom U, Scherf U, Biele M, Brabec C J. Influence of the molecular weight of poly(3-hexylthiophene) on the performance of bulk heterojunction solar cells. Chemistry of Materials, 2005, 17(8): 2175–2180

DOI

51
Koppe M, Brabec C J, Heiml S, Schausberger A, Duffy W, Heeney M, McCulloch I. Influence of molecular weight distribution on the gelation of P3HT and its impact on the photovoltaic performance. Macromolecules, 2009, 42(13): 4661–4666

DOI

52
Osaka I, Saito M, Mori H, Koganezawa T, Takimiya K. Drastic change of molecular orientation in a thiazolothiazole copolymer by molecular-weight control and blending with PC61BM leads to high efficiencies in solar cells. Advanced Materials, 2012, 24(3): 425–430

DOI

53
Müller C, Wang E, Andersson L M, Tvingstedt K, Zhou Y, Andersson M R, Inganäs O. Influence of molecular weight on the performance of organic solar cells based on a fluorene derivative. Advanced Functional Materials, 2010, 20(13): 2124–2131

DOI

54
Chu T Y, Alem S, Tsang S W, Tse S C, Wakim S, Lu J P, Dennler G, Waller D, Gaudiana R, Tao Y. Morphology control in polycarbazole based bulk heterojunction solar cells and its impact on device performance. Applied Physics Letters, 2011, 98(25): 253301–253303

DOI

55
Admassie S, Inganäs O, Mammo W, Perzon E, Andersson M R. Electrochemical and optical studies of the band gaps of alternating polyfluorene copolymers. Synthetic Metals, 2006, 156(7–8): 614–623

DOI

56
Koeckelberghs G, Cremer L D, Persoons A, Verbiest T. Influence of the substituent and polymerization methodology on the properties of chiral poly(dithieno[3,2-b:2′,3′-d]pyrrole)s. Macromolecules, 2007, 40(12): 4173–4181

Outlines

/