Research article Topical Sections

A simple route to linear and hyperbranched polythiophenes containing diketopyrrolopyrrole linking groups with improved conversion efficiency

  • Two novel polythiophene derivatives with linear or hyperbranched architectures using diketopyrrolopyrrole (DPP) as linking groups as well as normal poly(3-hexylthiophene) (P3HT) were synthesized via the Universal Grignard metathesis polymerization. The molecular weights of the linear polythiophene containing DPP linking moieties are higher than those of P3HT, while molecular weights of the hyperbranched one are smaller. The main decomposition temperatures of polymers were measured at 470 °C from TGA experiments. The UV-vis absorption behaviors of the DPP-containing polymers are similar to that of P3HT; moreover, these DPP-containing polythiophenes show distinct PL decay both in solution and thin film states. The electrochemical experiments reveal that the incorporation of DPP groups resulted in lowering HOMO levels of polymers. All polymers were blended with PC61BM and used as active layers for the fabrication of inverted polymer solar cells. The power conversion efficiency of devices based on linear and hyperbranched polythiophenes reached 3.74% and 2.38%, respectively, revealing comparable or even higher efficiency than the one based on normal P3HT.

    Citation: Chia-Hao Hsieh, Wei-Chi Chen, Sheng-Hsiung Yang, Yu-Chiang Chao, Hsiao-Chin Lee, Chia-Ling Chiang, Ching-Yi Lin. A simple route to linear and hyperbranched polythiophenes containing diketopyrrolopyrrole linking groups with improved conversion efficiency[J]. AIMS Materials Science, 2017, 4(4): 878-893. doi: 10.3934/matersci.2017.4.878

    Related Papers:

    [1] Avner Neubauer, Shira Yochelis, Gur Mittelman, Ido Eisenberg, Yossi Paltiel . Simple down conversion nano-crystal coatings for enhancing Silicon-solar cells efficiency. AIMS Materials Science, 2016, 3(3): 1256-1265. doi: 10.3934/matersci.2016.3.1256
    [2] Aruna P. Wanninayake, Shengyi Li, Benjamin C. Church, Nidal Abu-Zahra . Electrical and optical properties of hybrid polymer solar cells incorporating Au and CuO nanoparticles. AIMS Materials Science, 2016, 3(1): 35-50. doi: 10.3934/matersci.2016.1.35
    [3] Wei-Chi Chen, Pin-Yao Chen, Sheng-Hsiung Yang . Solution-processed hybrid light emitting and photovoltaic devices comprising zinc oxide nanorod arrays and tungsten trioxide layers. AIMS Materials Science, 2017, 4(3): 551-560. doi: 10.3934/matersci.2017.3.551
    [4] Silvia Colodrero . Conjugated polymers as functional hole selective layers in efficient metal halide perovskite solar cells. AIMS Materials Science, 2017, 4(4): 956-969. doi: 10.3934/matersci.2017.4.956
    [5] Katrin Gossen, Marius Dotter, Bennet Brockhagen, Jan Lukas Storck, Andrea Ehrmann . Long-term investigation of unsealed DSSCs with glycerol-based electrolytes of different compositions. AIMS Materials Science, 2022, 9(2): 283-296. doi: 10.3934/matersci.2022017
    [6] Aruna P. Wanninayake, Benjamin C. Church, Nidal Abu-Zahra . Effect of ZnO nanoparticles on the power conversion efficiency of organic photovoltaic devices synthesized with CuO nanoparticles. AIMS Materials Science, 2016, 3(3): 927-937. doi: 10.3934/matersci.2016.3.927
    [7] Andrea Ehrmann, Tomasz Blachowicz . Recent coating materials for textile-based solar cells. AIMS Materials Science, 2019, 6(2): 234-251. doi: 10.3934/matersci.2019.2.234
    [8] Yong X. Gan . A review of electrohydrodynamic casting energy conversion polymer composites. AIMS Materials Science, 2018, 5(2): 206-225. doi: 10.3934/matersci.2018.2.206
    [9] Etsana Kiros Ashebir, Berhe Tadese Abay, Taame Abraha Berhe . Sustainable A2BBX6 based lead free perovskite solar cells: The challenges and research roadmap for power conversion efficiency improvement. AIMS Materials Science, 2024, 11(4): 712-759. doi: 10.3934/matersci.2024036
    [10] Ha Thanh Tung, Ho Kim Dan, Dang Huu Phuc . Effect of the calcination temperature of the FTO/PbS cathode on the performance of a quantum dot-sensitized solar cell. AIMS Materials Science, 2023, 10(3): 426-436. doi: 10.3934/matersci.2023023
  • Two novel polythiophene derivatives with linear or hyperbranched architectures using diketopyrrolopyrrole (DPP) as linking groups as well as normal poly(3-hexylthiophene) (P3HT) were synthesized via the Universal Grignard metathesis polymerization. The molecular weights of the linear polythiophene containing DPP linking moieties are higher than those of P3HT, while molecular weights of the hyperbranched one are smaller. The main decomposition temperatures of polymers were measured at 470 °C from TGA experiments. The UV-vis absorption behaviors of the DPP-containing polymers are similar to that of P3HT; moreover, these DPP-containing polythiophenes show distinct PL decay both in solution and thin film states. The electrochemical experiments reveal that the incorporation of DPP groups resulted in lowering HOMO levels of polymers. All polymers were blended with PC61BM and used as active layers for the fabrication of inverted polymer solar cells. The power conversion efficiency of devices based on linear and hyperbranched polythiophenes reached 3.74% and 2.38%, respectively, revealing comparable or even higher efficiency than the one based on normal P3HT.


    1. Introduction

    Polythiophene (PT) and its derivatives have been extensively utilized for the fabrication of polymer solar cells (PSCs) with moderate power conversion efficiency (PCE) in the past decade. The most famous PT derivative, namely poly(3-hexylthiophene-2, 5-diyl) (P3HT), was firstly synthesized by McCullough et al. with high regioregularity > 95% by metathesis polymerization [1,2]. P3HT has high tendency to form close packing of main chains due to strong π–π interaction, which favors transport and migration of carriers. By applying thermal annealing process, the PSC using the blend of P3HT and [6,6] phenyl-C61-butyric acid methyl ester (PC61BM) as the active layer reached a PCE value of 4.4% [3]. Apart from P3HT, miscellaneous main-chain and side-chain types of PTs have widely been developed and reported in the literature [4,5,6,7,8,9]. A novel donor–acceptor PT bearing the C60 fullerene pendant has also been synthesized as the photoactive layer [10,11,12]. Most of those PTs belong to alternating copolymers, requiring two monomers tethering with different functional groups to react for copolymerization, such as dibromothiophenes with diboronic esters via the Suzuki-coupling reaction, or dibromo-monomers with bis(trimethyltin)-monomers via the Stille-coupling reaction. These monomers require well-designed transformation of functional groups and careful handling of stoichiometric balance to achieve high molecular weights and good photovoltaic performance of final polymers.

    In addition to linear architecture, bridging PTs have also been synthesized by introducing bridging moieties between polymer main chains to study their thermal and optical properties as well as device performance. A terthiophene-bridged P3HT derivative, namely PT-VTThV, was proposed by Li et al. by incorporating divinyl-terthiophene moiety between polymer main chains [13]. The conjugated bridging structure provides an alternative pathway for charge carriers to transport between polymer chains, and the enhancement in hole mobility of the polymer was verified. A PCE value of 1.72% for the solar device using PT-VTThV:PC61BM as the active layer was obtained. Another P3HT derivative B-P3HT containing bridging 3, 3'-dithiophene moiety was reported by Tu et al. [14]. They found that the hole mobility of B-P3HT was decreased compared with linear P3HT, possibly due to chain distortion brought by the small and rigid 3, 3'-dithiophene moiety. The solar devices using B-P3HT:PC61BM blends as the active layer showed PCE values of 0.13–2%. Two other hyperbranched conjugated polymers, namely p-3T-TCM and p-4T-TCM, were synthesized by Mangold et al. [15], using 2, 3-dithienylthiophene and 2, 3, 5-trithienylthiophene as bridging cores, respectively. The optimized PSC using p-3T-TCM:PC61BM blend as the active layer revealed an open-circuit voltage (VOC) of 714 mV which was significantly higher than that of P3HT:PC61BM blend. The obtained efficiency of 0.58–0.61% and extra high VOC of hyperbranched polymers show potential application in PSCs. In 2014, our group reported novel hyperbranched PT derivatives containing chlorinated perylene bisimide (PBI) or soft alkyl spacer as bridging moieties [16]. The synthesized polymers showed increased molecular weights and improved thermal stabilities compared with normal P3HT. Photovoltaic devices based on those hyperbranched polymers blended with PC61BM showed PCE values of 0.45–0.84% and high VOC values up to 0.70–0.72 V. To be honest, the device performance of those hyperbranched PT derivatives is still low and far beyond practical application, which requires further improvement through chemical modification.

    Here we propose new linear and hyperbranched PT derivatives containing diketopyrrolopyrrole (DPP) as linking moieties. DPP possesses highly electron-withdrawing nature and are commonly used for the synthesis of alternating donor–acceptor polymers with low band-gaps [17,18,19]. To synthesize alternating PTs, dibromo-monomers accompanying with diboronic ester derivatives or bis(trimethyltin)-monomers must be prepared before carrying out the polymerization. In this study, new P3HT derivative were obtained by simply introducing DPP-containing dibromothiophene or two-headed monomers during polymerization of P3HT via the Universal Grignard metathesis polymerization. Normal P3HT was also synthesized according to the same polymerization condition to examine the effect of introducing DPP linking moieties. The thermal, optical, and electrochemical behaviors of the synthesized polymers were investigated via different characterizing techniques. Finally, inverted PSCs based on polymer:PC61BM blends were fabricated and characterized to evaluate the performance of those polymers.


    2. Materials and Method


    2.1. Characterization Methods

    1H-and 13C-NMR spectra were performed on a Bruker Avance 600 MHz NMR spectrometer. Mass spectra were recorded on a Micromass TRIO-2000 GC-MS instrument, using electron impact (EI) or fast atom bombardment (FAB) as ionization source. Gel permeation chromatography (GPC) assembled from a Viscotek VE3850 RI detector and three columns in series were used to measure molecular weights and polydispersity index (PDI) values of polymers relative to polystyrene standards at 32 ℃. Tetrahydrofuran (THF) was used as the eluent to carry out GPC experiments. Fourier-transform infrared (FTIR) spectra were measured on a Nicolet iS-10 spectrometer. Thermogravimetric analysis (TGA) was undertaken on a Seiko TG/DTA 7200 instrument at a heating rate of 20 ℃/min. Differential scanning calorimetry (DSC) was performed on a Seiko DSC 6200 instrument at a heating rate of 10 ℃/min. UV-vis absorption and photoluminescence (PL) spectra were obtained with a Princeton Instruments Acton 2150 spectrophotometer equipped with an 150 W Xenon lamp (USHIO UXL-150S). Cyclic voltammetric measurements of materials were performed on an AUTOLAB PGSTAT30 electrochemical instrument. Indium-tin oxide (ITO) electrodes were used as both the working and counter electrodes, and silver/silver ions (Ag in 0.1 M AgNO3 solution, from Bioanalytical Systems, Inc.) was used as the reference electrode. The polymers were deposited on ITO electrodes and placed in acetonitrile solution containing 0.1 M tetrabutylammonium tetrafluoroborate as the supporting electrolyte with a scan rate of 50 mV/s. Ferrocene was used as an internal standard, and the potential values were obtained and converted to vs Fc/Fc+ (ferrocene reference electrode). Atomic force microscopy (AFM) experiments were performed on a Bruker Innova AFM instrument to investigate surface morphologies of polymer:PC61BM blend films. The current density-voltage (J-V) characteristics of photovoltaic devices were taken using a Keithley 2400 source measurement unit under an AM 1.5G exposure from a Yamashito Denso YSS-150AA solar simulator in ambient environment. The external quantum efficiency (EQE) measurements were performed on an assembled apparatus in the laboratory, comprising a solar simulator (Oriel 9600,150W), the monochromator (CornerstoneTM 130) and the optical power meter (818-UV/DB).


    2.2. Synthesis of Monomers

    The synthetic routes to monomers M1 and M2 are shown in Scheme 1. The preparation of intermediates (1) and (2) as well as monomers M1 and 2, 5-dibromo-3-hexylthiophene (M3) was referred to the previous literature [20,21]. The detailed synthetic procedure of M2 is described as follows.

    Scheme 1. Synthesis of the monomers M1 and M2.

    1-(6'-Bromohexyloxy)-4-methoxybenzene (3). A mixture of 4-methoxyphenol (10.0 g, 85.58 mmol), 1, 6-dibromohexane (60.0 g, 245.9 mmol), potassium hydroxide (6.0 g, 106.95 mmol) and dimethyl sulfoxide (DMSO) (100 mL) was stirred at room temperature for 6 hr. The mixture was then extracted with dichloromethane (DCM) and water, and the organic phase was dried with anhydrous magnesium sulfate (MgSO4). The crude product was concentrated in vacuo and purified by column chromatography (silica gel, using DCM:hexane = 2/1 in volume ratio as the eluent) to give a white solid (15.0 g, 65%). 1H-NMR (CDCl3, ppm): 1.46–1.51 (m, 4H, –OCH2CH2(CH2)2–CH2CH2Br), 1.76–1.81 (m, 2H, –CH2CH2Br), 1.87–1.91 (m, 2H, –OCH2CH2–), 3.40–3.44 (t, J = 6.9 Hz, 2H, –CH2Br), 3.77 (s, 3H, –OCH3), 3.90–3.93 (t, J = 6.3 Hz, 2H, –OCH2–), 6.83 (s, 4H, aromatic protons). 13C-NMR (CDCl3, ppm): 25.32, 27.94, 29.21, 32.70, 33.78, 55.75, 68.41,114.65,115.46,153.22,153.75. MASS (EI): m/z 287.

    1-(6'-Bromohexyloxy)-2, 5-dibromo-4-methoxybenzene (4). A mixture of compound (3) (5.0 g, 17.41 mmol), a small amount of iron powder and chloroform (100 mL) was stirred at 0 ℃ in an ice-water bath. Bromine (8.0 g, 50.0 mmol) was then slowly added to the mixture and stirred at room temperature for 12 hr. The reaction mixture was then extracted with ethyl acetate (EA) and sodium thiosulfate aqueous solution for three times, and the organic phase was dried with anhydrous MgSO4. The crude product was concentrated in vacuo and purified by column chromatography (silica gel, using EA/hexane = 1/14 in volume ratio as the eluent) to yield a white solid (5.0 g, 65%). 1H-NMR (CDCl3, ppm): 1.52–1.54 (m, 4H, –OCH2–CH2(CH2)2CH2CH2Br), 1.81–1.83 (m, 2H, –CH2CH2Br), 1.88–1.91 (m, 2H, –OCH2CH2–), 3.41–3.44 (t, J = 6.9 Hz, 2H, –CH2Br), 3.84 (s, 3H, –OCH3), 3.95–3.97 (t, J = 6.3 Hz, 2H, –OCH2–), 7.09 (s, 2H, aromatic protons). 13C-NMR (CDCl3, ppm): 25.19, 27.81, 28.92, 32.64, 33.74, 56.99, 70.06,110.41,111.27,117.01,118.65,150.04,150.55. MASS (EI): m/z 445.

    2, 5-Bis[6-(2, 5-dibromo-4-methoxyphenyl)hexyloxy]-3, 6-dithiophen-2-yl-pyrrolo[3, 4-c]pyrrole-1, 4-dione (M2). A mixture of compounds (1) (1.0 g, 3.33 mmol), (4) (3.3 g, 7.47 mmol), K2CO3 (3.83 g, 27.71 mmol), tetrabutylammonium bromide (TBABr) (0.13 g, 0.4 mmol), 18-crown-6 (0.1 g, 0.38 mmol) and N, N-dimethylformamide (50 mL) was refluxed at 145 ℃ for 24 hr under nitrogen atmosphere. After cooling to room temperature, the mixture was poured into 400 mL of water and filtered. The filter solid was washed with water for several times and then purified by column chromatography (silica gel, using DCM as the eluent) to give a dark red solid (1.0 g, 29%). 1H-NMR (CDCl3, ppm): 1.5 (m, 4H, –NCH2CH2CH2CH2CH2CH2O–), 1.56 (m, 4H, –NCH2CH2CH2CH2–CH2CH2O–), 1.7–1.9 (m, 8H, –NCH2CH2CH2CH2CH2CH2O–), 3.83 (s, 6H, –OCH3), 3.94 (m, 4H, –NCH2CH2CH2CH2CH2CH2O–), 4.09 (m, 4H, –NCH2CH2CH2CH2–CH2CH2O–), 7.07 (s, 4H, aromatic protons), 7.28 (s, 2H, aromatic protons), 7.63 (s, 2H, aromatic protons), 8.91 (s, 2H, aromatic protons). 13C-NMR (CDCl3, ppm): 25.63, 26.52, 28.96, 29.85, 42.04, 56.98, 70.11,107.68,110.37,111.28,116.99,118.65,128.64,129.68,130.72,135.30,139.99,150.06,150.51,161.35. MASS (FAB): m/z 1029.


    2.3. Synthesis of Polymers

    The syntheses of linear polymer P1 and hyperbranched polymer P2 are depicted in Scheme 2, which were polymerized via the Universal Grignard metathesis polymerization [22]. The molar ratio of isopropylmagnesium chloride–lithium chloride (i-PrMgCl·LiCl) complex to monomer was controlled to 1.3:1. The molar ratio of the catalyst 1, 2-bis(diphenyl-phosphinoethane)nickel(Ⅱ) chloride (Ni(dppp)Cl2) to monomer was controlled to 0.005:1. The feed ratio of monomers M1 (or M2) to M3 was selected to 0.05:1. The detailed synthetic procedures of the polymer P1 were given as an example.

    Scheme 2. Synthesis of the polymers P1 and P2.

    P1: to a mixture of M1 (0.104 g, 0.15 mmol) and M3 (1.0 g, 3.07 mmol) in anhydrous THF (40 mL) was added 1 M i-PrMgCl·LiCl (4 mL, 4.0 mmol) using syringe under nitrogen atmosphere. The reaction mixture was heated to 80 ℃ and stirred for 2 hr. A dispersion of Ni(dppp)Cl2 (9.0 mg, 1.66 × 10−2 mmol) in anhydrous THF (20 mL) was injected into the reaction mixture and stirred at 80 ℃ for 3 hr. The resulting solution was then poured into methanol (250 mL) and stirred for 1 hr. The crude product was collected by filtration and re-precipitated in hexane for several times to give a purple-black solid (0.33 g, 63%). 1H-NMR (CDCl3, ppm): 0.9–2.0 (m, alkyl–H), 2.56–2.8 (t, thiophene–CH2–), 4.16 (t, imide–CH2–), 6.9–7.1 (br, thiophene–H). FTIR (cm−1): 3055,2955,2922,2853,1635,1509,1455,819.

    P2: by following the synthetic procedure of P1 and using M2 (0.158 g, 0.152 mmol), M3 (1.0 g, 3.07 mmol), i-PrMgCl·LiCl (4 mL, 4.0 mmol) and Ni(dppp)Cl2 (9.0 mg, 1.66 × 10−2 mmol) as starting materials, P2 was obtained as a purple-black solid (0.16 g, 31%). 1H-NMR (CDCl3, ppm): 0.9–2.0 (m, alkyl–H), 2.58–2.8 (t, thiophene–CH2–), 3.78 (s, –O–CH3), 3.82–3.84 (m, –O–CH2– and imide–CH2–), 6.9–7.1 (br, aromatic–H). FTIR (cm−1): 3053,2954,2922,2851,1638,1508,1451,820.


    2.4. Device Fabrication

    P3HT, P1, and P2 were blended with PC61BM (1:1 in weight ratio) individually in o-dichlorobenzene (o-DCB) at a total concentration of 40 mg/mL. 30 μL of 1, 8-diiodooctane (DIO) was added in the above solutions to improve morphology of resulting blend films. The solutions were treated with ultrasonication at 70 ℃ for 30 min for better dissolution. The solutions were then filtered with 0.22 μm PTFE filters before use. Inverted photovoltaic devices with the configuration of ITO/ZnO nanorod arrays/ionic PF/polymer:PC61BM/PEDOT:PSS/WO3/Au were fabricated. The ZnO nanorod arrays were prepared according to the previous report and used as the electron transporting layer (ETL) [23]. An ultra-thin ionic PF layer was deposited on top of ZnO nanorod arrays from its dilute solution (0.1 wt% in acetonitrile) before depositing active layers. The ionic PF was prepared according to the previous report [24]. The active layers of polymer:PC61BM (120 nm) were prepared by spin-coating from their solutions at 1000 rpm for 30 sec, followed by solvent annealing in a covered glass petri dish at 60 ℃ for 5 min and dried in a vacuum oven at 90 ℃ for 30 min. Poly(3, 4-ethylenedioxythiophene):poly(styrene sulfonate) (PEDOT:PSS) (CleviosTM P VP AI 4083) and tungsten trioxide (WO3) layers were sequentially deposited on top of active layers. The WO3 layer was obtained from spin-coating tungsten(Ⅴ) ethoxide solution (0.1 wt% in anhydrous ethanol) into thin film, followed by hydrolysis in air for 15 min. Finally, 100 nm of gold electrodes were deposited as the anode by thermal evaporation at a pressure of 8 × 10−6 torr. The area of active layers is 4 mm2.


    3. Results and Discussion


    3.1. Synthesis and Characterization of Polymers

    In this research, linear PT derivative P1 and hyperbranched P2 by were synthesized by incorporating DPP-containing monomers M1 and M2, respectively, during polymerization of P3HT via the Universal Grignard metathesis. Normal P3HT was also synthesized from its monomer M3 under the same reaction condition for comparison. For polymer P1, DPP moieties are randomly incorporated into P3HT main chains to improve electron transport and dissociation of carriers. For polymer P2, both heads in M2 serve as polymerizable monomers to extend polymer chains, forming hyperbranched architectures of final polymers. The feed ratio of monomers M1 (or M2) to M3 is controlled to be 0.05:1 to ensure successful incorporation of DPP moieties and sufficient solubility of final polymers in common organic solvents, such as THF, chlorobenzene, and o-DCB. The number-average molecular weight (Mn), weight-average molecular weight (Mw), and PDI values of all synthesized polymers were determined by GPC. The Mn and Mw values of P1 are 2.54 × 104 and 6.23 × 104 g/mol, respectively, with a PDI value of 2.45. Moreover, the Mn and Mw values of P2 are 2.14 × 104 and 3.92 × 104 g/mol, respectively, with a smaller PDI value of 1.83. Meanwhile, the Mn, Mw and PDI values of synthesized P3HT were also measured to be 2.68 × 104, 4.87 × 104 g/mol and 1.81, respectively. It is seen that the molecular weight and PDI value of P1 are somewhat larger than those of P3HT and P2. Due to the fact that linear polystyrenes are used as the standards to construct the calibration curve, we consider that the measured molecular weight of branched P2 could be overestimated. Besides, it is reported that hyperbranched polymers have more densely packed structures, resulting in smaller hydrodynamic volume compared to those linear ones [25]. Hence, we speculate that the difference in molecular weights of polymers is affected by the polymer architecture and experimental method that we have for GPC measurements. The spectroscopic characterizations including NMR and FT-IR techniques were carefully carried out for all synthesized polymers. Besides, these polymers possess sufficient solubility in o-ODCB and can be cast into free-standing thin films for further investigations.


    3.2. Thermal Properties

    The TGA thermograms of the synthesized polymers P1, P2 and P3HT are depicted in Figure 1(a). The main decomposition temperatures (Td) due to thermal degradation of polymer main chains are found around 470 ℃, which is consistent with the previous literature [16,26]. A slight first-stage weight loss around 350 ℃, possibly due to the break of alkyl side chains or spacers, is also observed and in good accordance with our previous work [16].Besides, we notice that this first-stage weight loss can be suppressed by introducing DPP moiety on P3HT, indicating that the incorporation of rigid DPP moieties helps to improve thermal stability of the synthesized PT derivatives.

    The glass transition temperature (Tg) and melting temperature (Tm) of the synthesized polymers were determined by DSC. Several research groups have studied the thermal transition behaviors of P3HT; different Tg values of P3HT ranging from −14 to 110 ℃ [27,28,29,30], and Tm values in the range of 200–230 ℃ were reported in the literature [31,32]. This is understandable since the measurement of Tg and Tm is strongly dependent on the qualities and sources of polymers as well as experimental parameters. In this research, polymerization conditions for all polymers were carefully controlled to examine the effect of DPP linking moieties on thermal transition behaviors of final polymers. The DSC thermograms of the polymers P1, P2, and P3HT are shown in Figure 1(b). The Tg of the synthesized P1, P2, and P3HT in this study are observed at 72, 67, and 55 ℃, respectively. It is seen that DPP-containing derivatives show lower Tg values compared with P3HT. This can be attributed to less close packing of PT main chains by inserting rigid and bulky DPP moieties. P2 possesses the lowest Tg value due to its smallest molecular weights among three polymers. Furthermore, the Tm of the synthesized P1, P2, and P3HT are found at 231,230, and 215 ℃, respectively. Here again, P2 shows the lowest Tm value because of its less packed polymer chains and smaller molecular weights. As for the polymer P1, although its Tm value is close to that of P3HT, the broader melting peak indicates more disordered main chains of P1 by introducing DPP moieties compared with the regular homopolymer P3HT.

    Figure 1. (a) TGA and (b) DSC thermograms of P1, P2 and P3HT.

    3.3. Optical Properties

    The UV-vis absorption spectra of all polymers in solution or thin film state are shown in Figure 2. o-DCB is selected as solvent for optical measurement since it is also used for the preparation of blend solutions to fabricate photovoltaic devices in this study. In Figure 2(a), the absorption band centered at 460 nm are observed for P1 and P3HT in o-DCB, corresponding to π–π* transition along conjugated main chains. Besides, P2 reveals a slightly blue-shifted absorption maximum at 454 nm. This can be explained by the bridging architecture of P2 that forms twisted polymer chains to reduce the effective conjugation length. Figure 2(b) shows the absorption bands of all three polymers ranging from 400 to 650 nm in thin film state, revealing three vibronic absorption bands located at 527,552 and 602 nm. The wavelengths at 527 and 552 nm come from π–π* transition along shorter and longer conjugated polymer chains, respectively [32].The shoulder band at 602 nm is related to aggregates of polymer chains in thin film state [33]. It is seen that the absorption behaviors of the synthesized polymers in this study are similar to the regioregular P3HT reported in the literatures [21,27,33]. Moreover, we notice some differences in absorption between hyperbranched P2 and linear polymers P3HT and P1. For P3HT and P1, both spectral shapes are quite similar and the intensity of longer conjugation at 552 nm is stronger than the shorter one at 527 nm, as shown in Figure 2(b). Apparently, the two polymers P3HT and P1 own close conjugation length in thin film state. On the contrary, P2 shows stronger absorption at shorter wavelength of 527 nm, and its absorption shoulder at 602 nm is weaker than other two polymers. This could be due to the insertion of bridging DPP moieties that twist polymer chains and reduce effective conjugation length. The weaker absorption shoulder band implies hindered interaction and less packing between polymer chains, which may affect the charge transport and conversion efficiency of solar devices. The optical bandgaps (Eg) of polymers are estimated to be 1.89–1.91 eV from the absorption edge of their UV-vis absorption spectra, as shown in Figure 2(b).

    Figure 2. UV-vis absorption spectra of P1, P2 and P3HT in (a) o-DCB and (b) thin film state.

    The original PL emission spectra of polymers were depicted in Figure 3, which were obtained by excitation of samples at 460 nm in solution and 552 nm in solid state. The main emission band of all polymers is observed at 588 nm in o-DCB that is not affected by introducing DPP moieties. The vibronic shoulder band arised from interchain interaction can also be found around 635 nm for all polymers in Figure 3(a). Turning to PL emission spectra in Figure 3(b), two clear emission maxima around 650 and 700 nm are observed for all polymers in thin film state, which are attributed to intra-and inter-chain emissive behaviors, respectively [33,34]. Besides, two DPP-containing polymers P1 and P2 show distinct PL decay both in solution and thin film states compared to normal P3HT. This phenomenon implies a more effective transport process for carriers instead of recombination process to emit light, which has been reported in the previous reports [9,16,35]. It is noted that all samples were prepared according to the same standard, i.e., the same concentration in o-DCB and thickness in film.

    Figure 3. PL spectra of P1, P2 and P3HT in (a) o-DCB and (b) thin film state.

    3.4. Electrochemical Properties

    Figure 4 shows the cyclic voltammograms of the synthesized polymers in the oxidation process. The oxidation onset (EOX) of P3HT is found at 0.81 V, corresponding to its highest-occupied molecular orbital (HOMO) of −5.21 eV calculated from the semi-empirical equations [16]. The HOMO of our synthesized P3HT is close to several previous reports, including values of −5.2 [4], −5.1 [16,36,37,38], and −4.9 eV [39]. Here again, the determination of the HOMO of a certain P3HT can be varied with different material sources and experimental parameters. In our CV experiments, we carefully controlled all parameters in the same condition, such as the supporting electrolyte, scan rate and working/counter electrodes, to verify the influence of DPP moieties on polymers. The DPP-containing polymers P1 and P2 show higher EOX at 0.89 and 0.84 V, indicating their HOMO levels of −5.29 and −5.24 eV, respectively. The lower HOMO levels of DPP-containing polymers compared with P3HT is resulted from the incorporation of electron-withdrawing DPP groups. The electrochemical properties of P1, P2 and P3HT are summarized in Table 1. It is known that lower HOMO levels of polymers bring benefits of increasing VOC values for solar devices, since VOC can be determined from the energy difference between LUMO of electron acceptor PC61BM (LUMO = −4.3 eV) and HOMO of electron-donor polymers [35,40,41]. Combining the results of PL decay and lower HOMO levels, the synthesized DPP-containing P1 and P2 in this study may show comparable or even higher photovoltaic performance compared with normal P3HT.

    Figure 4. Cyclic voltammograms of P1, P2 and P3HT in the oxidation scan.
    Table 1. Electrochemical properties of P1, P2 and P3HT.
    Polymer EOX (V) a HOMO (eV) b LUMO (eV) c Eg (eV) d
    P1 0.89 −5.29 −3.4 1.89
    P2 0.84 −5.24 −3.34 1.9
    P3HT 0.81 −5.21 −3.31 1.89
    a data from CV experiments in the oxidation scan;
    b HOMO = −|EOX + 4.4|;
    c LUMO = −|HOMO + Eg|;
    d data from the edge of the absorption spectra in film state.
     | Show Table
    DownLoad: CSV

    3.5. Photovoltaic Properties

    Inverted solar cells with the configuration of ITO/ZnO nanorod arrays/ionic PF/polymer:PC61BM/PEDOT:PSS/WO3/Au were fabricated and evaluated. The synthesized polymers P1, P2 and P3HT were used as the electron donor, while commercial PC61BM was used as the electron acceptor to form the active layer. ZnO nanorod arrays were adopted as the ETL for the fabrication of inverted solar cells [42]. To improve contact between inorganic ZnO nanorods and organic polymer:PC61BM blend layer, a thin layer of ionic PF containing trimethylammonium hexafluorophosphate (–N+(CH3)3PF6) groups were deposited on top of ZnO nanorods, which was also used in our previous study [23]. Moreover, a thin layer of WO3 was inserted between PEDOT:PSS and Au electrode to act as hole extraction layer to achieve higher conversion efficiency [43,44]. Apart from WO3, one may choose other metal oxide buffer layers for efficient and stable polymer solar cells, such as nickel oxide, molybdenum trioxide, vanadium oxide and rhenium oxide [45]. The device parameters, including VOC, short-circuit current density (JSC), fill factor (FF) and PCE of PSCs based on P1, P2 and P3HT are summarized in Table 2, and corresponding current density-voltage characteristics as well as EQE spectra versus wavelength are shown in Figure 5. The optimized inverted solar device based on P1:PC61BM blend showed VOC, JSC, FF, and PCE values of 0.57 V, 16.21 mA/cm2, 40%, and 3.74%, respectively, which were significantly higher than those of the reference device using P3HT:PC61BM in this study. In other words, the introduction of DPP moieties enhances the photovoltaic properties of P3HT. Besides, the device using hyperbranched P2:PC61BM as the active layer showed VOC, JSC, FF, and PCE values of 0.58 V, 11.29 mA/cm2, 37%, and 2.38%, respectively. It can be seen that the devices using P1 and P2 as the electron donor showed larger VOC than the one of normal P3HT. This is reasonable since the deeper HOMO levels are found for polymers P1 and P2, as discussed in the previous part. The device results also show an increase in VOC for hyperbranched PTs that is in accordance with previous literatures [15,16], although the increment is less significant. From EQE spectra of three solar devices in Figure 5(b), it is also seen that P1 possesses the highest EQE from 300 to 650 nm, indicating the best JSC value among three polymers. Comparing the EQE spectra of P2 and P3HT, the intensity of P2 around 600 nm is lower and the curve drops sharply, making its final JSC and PCE value smaller than normal P3HT. The comparison of photovoltaic properties of PSCs based on bridging polymers from the previous literature and our study are listed in Table 3. It is seen that the device based on our synthesized hyperbranched P2 shows much higher JSC and PCE values compared with other works. To the best of our knowledge, the synthesized DPP-containing P2 in this study demonstrates the best device result among bridging polymers so far.

    Table 2. Parameters of inverted PSCs based on P1, P2 and P3HT blended with PC61BM under AM 1.5G illumination at 100 mA/cm2.
    Polymer VOC (V) JSC (mA/cm2) FF (%) PCE (%)
    P1 0.57 16.21 40 3.74
    P2 0.58 11.29 37 2.38
    P3HT 0.55 12.67 36 2.55
     | Show Table
    DownLoad: CSV
    Figure 5. (a) Current density-voltage characteristics and (b) EQE spectra of inverted PSCs based on P1, P2 and P3HT blended with PC61BM under AM 1.5G illumination at 100 mW/cm2.
    Table 3. Comparison of photovoltaic properties of PSCs based on bridging polymers (donor) and PC61BM (acceptor).
    Polymer VOC (V) JSC (mA/cm2) FF (%) PCE (%) Ref
    P1 0.57 16.21 40 3.74 this work
    P2 0.58 11.29 37 2.38 this work
    PT-VTThV 0.67 6.82 38 1.72 [10]
    B-P3HT 0.55 N/A a 47 2.0 [11]
    p-3T-TCM 0.714 2.09 40 0.58 [12]
    hyperbranched PT containing PBI 0.72 3.17 37 0.84 [13]
    a the data was not reported in the literature.
     | Show Table
    DownLoad: CSV

    From Table 2 and Figure 5, it is concluded that P1 shows the best device performance among three polymers in this study, mainly due to larger JSC, FF and EQE behavior. Although the P2 device shows lower JSC value compared with P3HT, the higher VOC and FF values make it comparable with P3HT. To further understand the performance trends of different polymers, we applied the AFM technique for morphological investigations. It is believed that proper phase separation between donor polymer and acceptor PC61BM is the key issue to high-efficiency solar devices [46,47,48]. The AFM topographic images of different polymer:PC61BM blend films are shown in Figure 6. The P1:PC61BM blend film shows the most homogeneous dispersion of aggregations, which is beneficial for charge dissociation between electron donor and acceptor. The root-mean-square roughness (Ra) of P1:PC61BM blend film is measured to be 7.25 nm that is also better than P3HT and P2 blend films, revealing Ra values of 8.24 and 11.5 nm, respectively. Combining adequate aggregate distribution and Ra value, the P1:PC61BM blend film represents the best surface morphology and film-forming property among three polymer blends, and higher device performance based on this P1:PC61BM blend is expected.

    Figure 6. AFM topographic images of (a) P1, (b) P2 and (c) P3HT thin films blended with PC61BM.

    4. Conclusion

    In this research, two different ways for incorporating DPP linking groups to form linear or hyperbranched PT derivatives were proposed and characterized. The effects of introducing DPP groups onto P3HT were investigated by miscellaneous techniques, such as GPC, thermal analysis, optical and electrochemical measurements. The distinct PL decay of DPP-containing polymers in thin film state indicates a more efficient transport process for charge carriers. The lower HOMO levels of polymers resulted from incorporating electron-withdrawing DPP groups brings benefits in increasing VOC values for photovoltaic application. The inverted solar device based on P1:PC61BM blend shows high JSC and PCE values of 16.21 mA/cm2 and 3.74%, respectively. The solar device using P2:PC61BM blend as the active layer reveals a PCE value of 2.38% that demonstrates the highest record among bridging polymers in the literature.


    Acknowledgments

    The authors gratefully thank to the Ministry of Science and Technology of the Republic of China (NSC 102-2113-M-009-007) for financial support of this work.


    Conflict of Interest

    All authors declare no conflicts of interest in this paper.


    [1] Loewe RS, Khersonsky SM, McCullough RD (1999) A simple method to prepare head-to-tail coupled, regioregular poly(3-alkylthiophenes) using Grignard metathesis. Adv Mater 11: 250–253. doi: 10.1002/(SICI)1521-4095(199903)11:3<250::AID-ADMA250>3.0.CO;2-J
    [2] Loewe RS, Ewbank PC, Liu J, et al. (2001) Regioregular, head-to-tail coupled poly(3-alkylthiophenes) made easy by the GRIM method: investigation of the reaction and the origin of regioselectivity. Macromolecules 34: 4324–4333. doi: 10.1021/ma001677+
    [3] Kim Y, Cook S, Tuladhar SM, et al. (2006) A strong regioregularity effect in self-organizing conjugated polymer films and high-efficiency polythiophene:fullerene solar cells. Nat Mater 5: 197–203. doi: 10.1038/nmat1574
    [4] Cheng YJ, Yang SH, Hsu CS (2009) Synthesis of conjugated polymers for organic solar cell applications. Chem Rev 109: 5868–5923. doi: 10.1021/cr900182s
    [5] Chang YT, Hsu SL, Chen GY, et al. (2008) Intramolecular donor–acceptor regioregular poly(3-hexylthiophene)s presenting octylphenanthrenyl-imidazole moieties exhibit enhanced charge transfer for heterojunction solar cell applications. Adv Funct Mater 18: 2356–2365. doi: 10.1002/adfm.200701150
    [6] Chang YT, Hsu SL, Su MH, et al. (2009) Intramolecular donor–acceptor regioregular poly(hexylphenanthrenyl-imidazole thiophene) exhibits enhanced hole mobility for heterojunction solar cell applications. Adv Mater 21: 2093–2097. doi: 10.1002/adma.200802379
    [7] Hou J, Tan Z, Yan Y, et al. (2006) Synthesis and photovoltaic properties of two-dimensional conjugated polythiophenes with bi(thienylenevinylene) side chains. J Am Chem Soc 128: 4911–4916. doi: 10.1021/ja060141m
    [8] Li Y, Zou Y (2008) Conjugated polymer photovoltaic materials with broad absorption band and high charge carrier mobility. Adv Mater 20: 2952–2958. doi: 10.1002/adma.200800606
    [9] Zhang Q, Cirpan A, Russell TP, et al. (2009) Donor–acceptor poly(thiophene-block-perylene diimide) copolymers: synthesis and solar cell fabrication. Macromolecules 42: 1079–1082. doi: 10.1021/ma801504e
    [10] Lanzi M, Paganin L, Errani F (2012) Synthesis, characterization and photovoltaic properties of a new thiophene-based double-cable polymer with pendent fullerene group. Polymer 53: 2134–2145. doi: 10.1016/j.polymer.2012.03.040
    [11] Lanzi M, Salatelli E, Benelli T, et al. (2015) A regioregular polythiophene-fullerene for polymeric solar cells. J Appl Polym Sci 132: 42121.
    [12] Piereini F, Lanzi M, Nakielski P, et al. (2017) Single-Material Organic Solar Cells Based on Electrospun Fullerene-Grafted Polythiopene Nanofibers. Macromolecules 50: 4972–4981. doi: 10.1021/acs.macromol.7b00857
    [13] Zhou E, Tan Z, Yang Y, et al. (2007) Synthesis, hole mobility, and photovoltaic properties of cross-linked polythiophenes with vinylene-terthiophene-vinylene as conjugated bridge. Macromolecules 40: 1831–1837. doi: 10.1021/ma062633p
    [14] Tu G, Bilge A, Adamczyk S, et al. (2007) The influence of interchain branches on solid state packing, hole mobility and photovoltaic properties of poly(3-hexylthiophene) (P3HT). Macromol Rapid Comm 28: 1781–1785. doi: 10.1002/marc.200700239
    [15] Mangold HS, Richter TV, Link S, et al. (2012) Optoelectronic properties of hyperbranched polythiophenes. J Phys Chem B 116: 154–159.
    [16] Yang SH, Lin TS, Huang YZ, et al. (2014) Synthesis of hyperbranched polythiophenes containing tetrachloroperylene bisimide as bridging moiety for polymer solar cells. Polymer 55: 6058–6068. doi: 10.1016/j.polymer.2014.09.046
    [17] Li W, Hendriks KH, Roelofs WSC, et al. (2013) Efficient small bandgap polymer solar cells with high fill factors for 300 nm thick films. Adv Mater 25: 3182–3186. doi: 10.1002/adma.201300017
    [18] Tan H, Deng X, Yu J, et al. (2013) A novel benzo[1,2-b:4,5-b']dithiophene-based conjugated polymer with a pendant diketopyrrolopyrrole unit for high-performance solar cells. Macromolecules 46: 113–118. doi: 10.1021/ma302243r
    [19] Kanimozhi C, Balraju P, Sharma GD, et al. (2010) Synthesis of diketopyrrolopyrrole containing copolymers: a study of their optical and photovoltaic properties. J Phys Chem B 114: 3095–3103. doi: 10.1021/jp909183x
    [20] Huo L, Hou J, Chen HY, et al. (2009) 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 42: 6564–6571. doi: 10.1021/ma9012972
    [21] Roncali J (1997) Synthetic principles for bandgap control in linear π-conjugated systems. Chem Rev 97: 173–205. doi: 10.1021/cr950257t
    [22] Stefan MC, Javier AE, Osaka I, et al. (2009) Grignard metathesis method (GRIM): toward a universal method for the synthesis of conjugated polymers. Macromolecules 42: 30–32. doi: 10.1021/ma8020823
    [23] Chen WC, Chen PY, Yang SH (2017) Solution-processed hybrid light emitting and photovoltaic devices comprising zinc oxide nanorod arrays and tungsten trioxide layers. AIMS Mater Sci 4: 551–560. doi: 10.3934/matersci.2017.3.551
    [24] Huang WJ, Huang PH, Yang SH (2016) PCBM doped with fluorene-based polyelectrolytes as electron transporting layers for improving the performance of planar heterojunction perovskite solar cells. Chem Commun 52: 13572–13575. doi: 10.1039/C6CC07062K
    [25] Nunez CM, Chiou BS, Andrady AL, et al. (2000) Solution Rheology of Hyperbranched Polyesters and Their Blends with Linear Polymers. Macromolecules 33: 1720–1726. doi: 10.1021/ma991044z
    [26] Lee JU, Jung JW, Emrick T, et al. (2010) Synthesis of C60-end capped P3HT and its application for high performance of P3HT/PCBM bulk heterojunction solar cells. J Mater Chem 20: 3287–3294.
    [27] Stevens DM, Qin Y, Hillmyer MA, et al. (2009) Enhancement of the morphology and open circuit voltage in bilayer polymer/fullerene solar cells. J Phys Chem C 113: 11408–11415.
    [28] Zhao Y, Yuan G, Roche P (1995) A calorimetric study of the phase transitions in poly(3-hexylthiophene). Polymer 36: 2211–2214.
    [29] Zhao J, Swinnen A, Assche GV, et al. (2009) Phase diagram of P3HT/PCBM blends and its implication for the stability of morphology. J Phys Chem B 113: 1587–1591.
    [30] Kim Y, Choulis SA, Nelson J, et al. (2005) Device annealing effect in organic solar cells with blends of regioregular poly(3-hexylthiophene) and soluble fullerene. Appl Phys Lett 86: 063502. doi: 10.1063/1.1861123
    [31] Liao HC, Chantarat N, Chen SY, et al. (2011) Annealing effect on photovoltaic performance of hybrid P3HT/In-Situ grown CdS nanocrystal solar cells. J Electrochem Soc 158: E67–E72. doi: 10.1149/1.3585668
    [32] Li G, Shrotriya V, Yao Y, et al. (2007) Manipulating regioregular poly(3-hexylthiophene): [6,6]-phenyl-C61-butyric acid methyl ester blends-route towards high efficiency polymer solar cells. J Mater Chem 17: 3126–3140.
    [33] Brown PJ, Thomas DS, Köhler A, et al. (2003) Effect of interchain interactions on the absorption and emission of poly(3-hexylthiophene). Phys Rev B 67: 064203. doi: 10.1103/PhysRevB.67.064203
    [34] Noriega R, Rivnay J, Vandewal K, et al. (2013) A general relationship between disorder, aggregation and charge transport in conjugated polymers. Nat Mater 12: 1038–1044. doi: 10.1038/nmat3722
    [35] Yao K, Chen L, Li F, et al. (2012) Cooperative assembly donor–acceptor system induced by intermolecular hydrogen bonds leading to oriented nanomorphology for optimized photovoltaic performance. J Phys Chem C 116: 714–721.
    [36] Kim JY, Lee K, Coates NE, et al. (2007) Efficient tandem polymer solar cells fabricated by all-solution processing. Science 317: 222–225. doi: 10.1126/science.1141711
    [37] Scharber MC, Mühlbacher D, Koppe M, et al. (2006) Design rules for donors in bulk-heterojunction solar cells-towards 10% energy-conversion efficiency. Adv Mater 18: 789–794. doi: 10.1002/adma.200501717
    [38] Zhu R, Jiang CY, Liu B, et al. (2009) Highly efficient nanoporous TiO2-polythiophene hybrid solar cells based on interfacial modification using a metal-free organic dye. Adv Mater 21: 994–1000. doi: 10.1002/adma.200802388
    [39] Hou J, Chen TL, Zhang S, et al. (2009) An easy and effective method to modulate molecular energy level of poly(3-alkylthiophene) for high-VOC polymer solar cells. Macromolecules 42: 9217–9219. doi: 10.1021/ma902197a
    [40] Sekiguchi H, Sekiguchi T (2014) Molecular ordering effect of regioregular poly(3-hexylthiophene) using sulfur K-edge X-ray absorption spectroscopy. Jpn J Appl Phys 53: 02BB07.
    [41] Chen HY, Hou J, Zhang S, et al. (2009) Polymer solar cells with enhanced open-circuit voltage and efficiency. Nat Photonics 3: 649–653. doi: 10.1038/nphoton.2009.192
    [42] Huang JS, Chou CY, Lin CF (2014) Enhancing performance of organic-inorganic hybrid solar cells using a fullerene interlayer from all-solution processing. Sol Energ Mat Sol C 6: 466–471.
    [43] Tan Z, Li L, Cui C, et al. (2012) Solution-processed tungsten oxide as an effective anode buffer layer for high-performance polymer solar cells. J Phys Chem C 116: 18626–18632.
    [44] Lampande R, Kim GW, Boizot J, et al. (2013) A highly efficient transition metal oxide layer for hole extraction and transport in inverted polymer bulk heterojunction solar cells. J Mater Chem A 1: 6895–6900. doi: 10.1039/c3ta10863e
    [45] Wang F, Tan Z, Li Y (2015) Solution-processable metal oxides/chelates as electrode buffer layers for efficient and stable polymer solar cells. Energ Environ Sci 8: 1059–1091. doi: 10.1039/C4EE03802A
    [46] Ma W, Yang C, Gong X, et al. (2005) Thermally stable, efficient polymer solar cells with nanocontrol of the interpenetrating network morphology. Adv Funct Mater 15: 1617–1622. doi: 10.1002/adfm.200500211
    [47] Li G, Shrotriya V, Yao Y, et al. (2005) Investigation of annealing effects and film thickness dependence of polymer solar cells based on poly(3-hexylthiophene). J Appl Phys 98: 043704. doi: 10.1063/1.2008386
    [48] Peet J, Kim Y, Coates NE, et al. (2007) Efficiency enhancement in low-bandgap polymer solar cells by processing with alkane dithiols. Nat Mater 6: 497–500. doi: 10.1038/nmat1928
  • This article has been cited by:

    1. Ayesha Kausar, 2021, 9780128224632, 129, 10.1016/B978-0-12-822463-2.00002-6
    2. Ayesha Kausar, Epitome of Fullerene in Conducting Polymeric Nanocomposite—Fundamentals and Beyond, 2023, 62, 2574-0881, 618, 10.1080/25740881.2022.2121223
    3. Ayesha Kausar, 2023, 9780323995153, 65, 10.1016/B978-0-323-99515-3.00003-1
    4. Zhaoqiong Zhou, Yan Li, Pinyu Chen, Jinqiu Meng, Nan Luo, Liang Luo, Feng He, Xiangfeng Shao, Hao‐Li Zhang, Zitong Liu, Hyperbranched Diketopyrrolopyrrole‐based Polymers Constructed via Linear Side‐Chains towards Organic Field‐Effect Transistors, 2023, 0947-6539, 10.1002/chem.202203361
    5. Zhaoqiong Zhou, Nan Luo, Xiangfeng Shao, Hao‐Li Zhang, Zitong Liu, Hyperbranched Polymers for Organic Semiconductors, 2023, 88, 2192-6506, 10.1002/cplu.202300261
    6. Aleksei Maksimov, Bulat Yarullin, Kharlampii Kharlampidi, Gennadii Kutyrev, Advances in hyperbranched polymer chemistry, 2024, 1026-1265, 10.1007/s13726-024-01379-6
  • Reader Comments
  • © 2017 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(5872) PDF downloads(1049) Cited by(6)

Article outline

Figures and Tables

Figures(8)  /  Tables(3)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog