Dyes and Pigments 163 (2019) 356–362
Contents lists available at ScienceDirect
Dyes and Pigments journal homepage: www.elsevier.com/locate/dyepig
New-structure perylene diimide oligomers by the linkage of the bay- and imide-position for nonfullerene solar cells
T
Zhenghui Luoa,b, Kailong Wub, Yuan Zhaob, Beibei Qiuc,d, Yongfang Lic,d,∗, Chuluo Yanga,b,∗∗ a
Shenzhen Key Laboratory of Polymer Science and Technology, College of Materials Science and Engineering, Shenzhen University, Shenzhen, 518060, China Hubei Key Lab on Organic and Polymeric Optoelectronic Materials, Department of Chemistry, Wuhan University, Wuhan, 430072, China c CAS Research/Education Center for Excellence in Molecular Sciences, CAS Key Laboratory of Organic Solids, Institute of Chemistry, Chinese Academy of Sciences, Beijing, 100190, China d School of Chemical Science, University of Chinese Academy of Sciences, Beijing, 100049, China b
A R T I C LE I N FO
A B S T R A C T
Keywords: Perylene diimide Polymer solar cells Non-fullerene Photovoltaic performance
A new family of perylene diimide (PDI) oligomers, T-PDI and H-PDI, were designed and synthesized via dehydration condensation reactions with two PDI monomers linked together at bay- and imide-positions. The as-cast non-fullerene polymer solar cells (PSCs) based on H-PDI as acceptor achieved a power conversion efficiency (PCE) of 1.91%. In comparison with the device based on H-PDI, the as-cast T-PDI based PSC gave a PCE of 3.50% with higher open-circuit voltage (VOC), short-circuit density (JSC) and fill factor (FF). The higher VOC of the PSC based on T-PDI was consistent well with its shallower LUMO energy level compared with that of H-PDI. Meanwhile, the higher JSC and FF for the T-PDI-based device should be mainly attributed to the stronger light absorption, more balanced carrier transport and favorable morphology relative to those of the device based on H-PDI. These results reveal the relationship between the PDI oligomers and the photovoltaic performance.
1. Introduction Over the past two decades, bulk-heterojunction (BHJ) polymer solar cells (PSCs) have been studied extensively [1–4]. The promise of flexibility, light weight and low cost has attracted a number of researchers [5–10]. In generally, the active layers of BHJ PSCs include a fullerene derivative as acceptor and a polymer as donor [11–13]. However, in recent three years, non-fullerene acceptors have shown great potential in the PSCs [14–19], especially 3,9-bis(2-methylene-(3-(1,1-dicyanomethylene)-indanone)-5,5,11,11-tetrakis(4-hexylphenyl)-dithieno-[2,3d:2′,3′-d′]-s-indaceno[1,2-b:5,6-b′]-dithiophene (ITIC) and perylene diimide (PDI) derivatives [20–40], which can be comparable or even superior to the fullerene derivatives [41,42]. PDI dimers and trimers, a very pivotal type of PDI derivatives, are developed by PDI monomers coupling via chemical bonds (singly-, doubly-, and triply-linked), without other bulky bridge-blocks such as spirofluorene [35,43], tetraphenyl ethylene [36] and benzene [32,33,44]. PDI dimers and trimers play an important role in the nonfullerene acceptors due to the desirable combined properties of good
film-forming, high molar extinction coefficient and excellent electron mobilities [40,45–48]. As known in Fig. 1, the PDI monomer includes three different reaction positions: ortho- (α-), bay- (β-) and imide position. Most reported PDI dimers and trimers are the homologous ones, which are usually synthesized by linking two PDI monomers via β- or the imide positions (Fig. 1a and b) [40,45,46]. Until recently, the first heterologous PDI arrays (di-PDI and tri-PDI) were synthesized by Suzuki cross-coupling reactions between the β-position bromide of PDI monomer and the α-position borate of PDI monomer, which exhibited unique photophysical and electrochemical properties relative to the homologous ones [49]. And the PSC based on tri-PDI as acceptor achieved the high power conversion efficiency (PCE) of 4.55%. These results indicate that developing heterologous PDI oligomers provides an alternative way to investigate fullerene-free electron acceptors for high efficient PSCs. Inspired by the above considerations, we present two heterologous PDI molecules (T-PDI and H-PDI) in Fig. 1c. The shapes of T-PDI and HPDI resemble the capital letters ‘‘T’’ and ‘‘H’’, respectively. The two heterologous molecules were efficiently synthesized through
∗ Corresponding author. CAS Research/Education Center for Excellence in Molecular Sciences, CAS Key Laboratory of Organic Solids, Institute of Chemistry, Chinese Academy of Sciences, Beijing, 100190, China. ∗∗ Corresponding author. Shenzhen Key Laboratory of Polymer Science and Technology, College of Materials Science and Engineering, Shenzhen University, Shenzhen, 518060, China. E-mail addresses:
[email protected] (Y. Li),
[email protected] (C. Yang).
https://doi.org/10.1016/j.dyepig.2018.12.015 Received 22 October 2018; Received in revised form 10 December 2018; Accepted 10 December 2018 Available online 11 December 2018 0143-7208/ © 2018 Elsevier Ltd. All rights reserved.
Dyes and Pigments 163 (2019) 356–362
Z. Luo et al.
Fig. 1. a) Structures of PDI dimers and trimers via β-β positions; b) Structures of PDI dimers and trimers via the imide-imide positions; c) Structures of T-PDI and HPDI via the imide-β positions.
dehydration condensation reactions between the β-position amino group of one PDI monomer and the imide position of the other PDI monomer. The two acceptors possess quite twisted structures, which can suppress aggregations of PDIs. In comparison with H-PDI, T-PDI shows higher LUMO level, stronger ability of light absorption and more favorable orbital distribution, which make T-PDI a more promising electron acceptor. PSCs based on PBDB-T:T-PDI without thermal annealing and additive treatment achieved a PCE of 3.50%. A lower PCE of 1.91% for the as-cast device of PBDB-T:H-PDI was obtained mainly due to its poor JSC and VOC.
chloroform (CF).
2.2. Optical properties and electronic energy levels The optical properties of T-PDI and H-PDI were characterized by measuring their ultraviolet–visible (UV–vis) absorption spectra in thin films and in chloroform solution (10−5 M). As can be seen in Fig. 2a, in dilute chloroform solution, T-PDI exhibits strong absorption peaks in the wavelength range of 410–560 nm with a maximum extinction coefficient (εmax) of 1.43 × 105 M−1 cm−1 at 529 nm, which is higher than that of H-PDI (1.11 × 105 M−1 cm−1 at 523 nm). Neat T-PDI and H-PDI films show similar absorption spectrum to the corresponding solution one, which is similar to the reported spiro-fused PDI dimer (SDP) [48], indicative of less aggregation tendency among PDI cores. However, for the single chromophore PDI, red-shifted and broader absorption spectrum was observed relative to that of T-PDI and H-PDI (Fig. S1), which is recognized as H-aggregate formation. In addition, the energy levels of T-PDI and H-PDI were also investigated by employing electrochemical cyclic voltammetry (CV) method, and the corresponding data were outline in Table 1. T-PDI and H-PDI show multiple reduction waves. The two different reduction potentials were observed for T-PDI, with the higher attributed to the PDI with ‘PDI’ at the imide position due to the weaker electron donating ability of a
2. Results and discussions 2.1. Synthesis and characterization The synthetic routes of the two heterologous PDI molecules (T-PDI and H-PDI) were outlined in Scheme 1. The reduction of PDI-NO2 via stannous chloride reduction method yielded the compound PDI-NH2. Then dehydration condensation reactions between PDI-NH2 and PDAC6/PDA generated T-PDI/H-PDI, respectively. The new compounds were fully characterized by 1H NMR, 13C NMR, MALDI-TOF and elemental analysis. They show quite good solubility in common organic solvents, such as o-dichlorobenzene (o-DCB), chlorobenzene (CB) and
Scheme 1. Synthesis of T-PDI, H-PDI. 357
Dyes and Pigments 163 (2019) 356–362
Z. Luo et al.
Fig. 2. a) UV–vis absorption spectra of T-PDI and H-PDI in chloroform solution (10−5 M); b) Normalized absorption spectra of T-PDI, H-PDI and PBDB-T in films; c) Energy level diagram of PBDB-T, T-PDI and H-PDI.
achieved a higher VOC of 0.837 V, a JSC of 7.64 mA cm−2 and a FF of 0.547, thereby an enhanced PCE of 3.50% was obtained. The poor VOC for the device of H-PDI is consistent with its lower LUMO energy level relative to that of T-PDI. Meanwhile, the higher JSC and FF for the TPDI-based device could be ascribed to the stronger ability of light absorption and more favorable morphology compared with the device of H-PDI. Narayan and coworkers reported that monomer PDI and a PDI dimer achieved PCEs of 0.13% and 2.78% [47], respectively. The worse PCE of monomer PDI resulted from the oversize phase separation. Xiao's group also reported a 2.84% efficiency by using heterologous di-PDI [49]. The significantly higher PCE of T-PDI is mainly ascribed to the larger VOC compared to the two mentioned PDI dimer above. The external quantum efficiency (EQE) curves of the PSCs based on T-PDI and H-PDI are shown in Fig. 4d. Both of the PBDB-T:T-PDI and PBDB-T:H-PDI-based devices displayed a broad response from 300 to 700 nm. The calculated current density values integrated from the EQE spectra for the PBDB-T:T-PDI and PBDB-T:H-PDI-based devices are 7.65 and 5.16 mA/cm2, respectively, which are in good agreement with the JSC values measured from J−V curves.
single alkyl chain compared with the PDI with imide alkyl chains, and the lower assigned to the other PDI. Similarly, for H-PDI, the highest reduction potential is ascribed to the PDI with ‘PDI’ at the imide position”, and the other two are ascribed to the PDI with imide alkyl chains. Also, the HOMO and LUMO energy levels of T-PDI were estimated to be −5.85 eV and −3.95 eV from the onset oxidation and reduction potentials. Lower HOMO energy level (−5.92 eV) and LUMO energy level (−4.02 eV) were obtain for H-PDI. The higher LUMO energy level for TPDI facilitates the realization of higher VOC. 2.3. Theoretical analysis Density functional theory (DFT) calculations were employed to obtain the analogous models of T-PDI and H-PDI (Fig. 3). Both T-PDI and H-PDI display quite twisted molecular geometries, which can suppress intermolecular aggregation. For T-PDI, the dihedral angle between the two PDIs is calculated to be 74°. A smaller dihedral angle between adjacent two PDIs for H-PDI is 66°. As shown in Fig. 3, the imide positions of the terminal PDIs have two alkyl chains with weak electron donating ability, while the imide positions of the core-PDI are connected to the β positions of the strong electron-withdraw PDI, which make the LUMO locate on the core-PDI. The HOMO/LUMO values of TPDI and H-PDI are −5.98/-3.66 eV and −6.02/-3.83 eV through DFT calculations, respectively. The calculated values are in good agreement with the values by the CV measurements.
2.5. Charge carrier mobility Charge transport property has an important effect on the device performance. Thus we employed space-charge-limited-current (SCLC) method to survey the electron and hole mobilities (see Fig. S2), and the corresponding parameters were summarized in Table 1. The hole and electron mobility of PBDB-T:H-PDI blend films were 1.02 × 10−4 and 5.17 × 10−5 cm2 V−1 s−1 (μh/μe = 1.97), respectively. The higher hole and electron mobility (1.10 × 10−4 cm2 V−1 s−1/ 8.42 × 10−5 cm2 V−1 s−1) and more balanced μh/μe (μh/μe = 1.31) were obtained for the PBDB-T:T-PDI-based blend films, which can be a reason for the higher JSC and FF in the corresponding device. In addition, electron mobilities of neat T-PDI and H-PDI films were also measured by employing SCLC method (Fig. S3). In comparison with HPDI, T-PDI gives higher electron mobility (9.87 × 10−5 cm2 V−1 s−1), which contributes to the carrier transport.
2.4. Photovoltaic properties In order to estimate the photovoltaic performance of T-PDI and HPDI, PSCs were constructed with a conventional device configuration architecture of ITO/PEDOT:PSS poly(3,4-ethylenedioxythiophene):poly (styrenesulfonate)/PBDB-T: acceptors/PDINO (perylene diimide functionalized with amino N-oxide)/Al (Fig. 4b). The current density-voltage (J-V) curves of PBDB-T:T-PDI and PBDB-T:H-PDI-based devices without thermal annealing and additives are shown in Fig. 4c, and the corresponding photovoltaic data are presented in Table 2. The as-cast solar cell based on PBDB-T:H-PDI exhibited a PCE of 1.91% with a VOC of 0.734 V, JSC of 5.14 mA cm−2 and FF of 0.506. In comparison with the device of H-PDI, the as-cast T-PDI-based device Table 1 Optoelectronic data of T-PDI and H-PDI. Acceptor
εmaxa (105 M−1 cm−1)
λmaxfilm (nm)
λonsetfilm (nm)
b
T-PDI H-PDI
1.43 1.11
529 523
571 577
2.18 2.19
a b
Egopt (eV)
HOMODFT (eV)
LUMODFT (eV)
HOMOCV (eV)
LUMOCV (eV)
μe (cm2 V−1 s−1)
−5.98 −6.02
−3.66 −3.83
−5.85 −5.92
−3.95 −4.02
9.87 × 10−5 7.80 × 10−5
In chloroform solution. Calculated from Egopt = 1240/λonset. 358
Dyes and Pigments 163 (2019) 356–362
Z. Luo et al.
Fig. 3. Optimal molecular geometries, HOMO and LUMO electron density distributions using density functional theory (DFT) by Gaussian 09 at B3LYP/6-31G(d) level.
active layer, which is favorable for the charge collection and transport. In the TEM image, more evident and interconnected phase-separated domain sizes are clearly observed in the PBDB-T: T-PDI blend films, which match well with the AFM phase image. These results should be responsible for higher FF and JSC in the PBDB-T:T-PDI-based device.
2.6. Morphology characterization To gain further insight into the effect of surface morphology on the photovoltaic performance, transmission electron microscopy (TEM) and atomic force microscopy (AFM) were utilized to investigate the morphology of the active layers (Fig. 5). The blend films of PBDB-T:T-PDI and PBDB-T:H-PDI exhibited smooth and uniform morphologies with root-mean-square (RMS) surface roughness of 1.38 and 1.30 nm, respectively. And the higher RMS value for the blend film of PBDB-T and T-PDI enlarges the contact area between the cathode interface and
3. Conclusions In summary, a new family of PDI oligomers, heterologous PDI dimers and trimers, were designed and efficiently synthesized via
Fig. 4. a) Chemical structure of donor PBDB-T; b) Device configuration of the studied PSCs; c) J-V characteristics curves and (d) EQE spectra of the best performing PSCs of PBDB-T:T-PDI (1:1 w/w) and PBDB-T:H-PDI (1:1 w/w). 359
Dyes and Pigments 163 (2019) 356–362
Z. Luo et al.
Table 2 Summary of photovoltaic parameters of PSCs based on PBDB-T/acceptors under standard AM 1.5G illumination, 100 mW/cm2. Devices PBDB-T:T-PDI PBDB-T:H-PDI a
Voca (V) 0.837 0.734
Jsca (mA cm−2) 7.64 5.14
FFa % 54.7 50.6
PCEa % 3.50 (3.36 ± 0.07) 1.91 (1.81 ± 0.06)
μh (cm2 v1 s−1) −4
1.10 × 10 1.02 × 10−4
μe (cm2 v−1 s−1) −5
8.42 × 10 5.17 × 10−5
μh/μe 1.31 1.97
The values in parentheses are average values obtained from 10 devices.
dehydration condensation reactions between the β-position of one PDI monomer and the imide position of the other PDI monomer. The absorption spectra indicate that the intermolecular aggregations in the solid state of the two PDI oligomers were effectively suppressed due to the highly twisted molecular configurations. Meanwhile, T-PDI displayed higher molar extinction coefficient and LUMO levels relative to that of H-PDI. Employing the two PDI oligomers as electron acceptors, the as-cast non-fullerene PSC based on H-PDI achieved a PCE of 1.91% with a VOC of 0.734 V, a JSC of 5.14 mA cm−2 and a FF of 0.506. In comparison with the H-PDI-based device, the as-cast T-PDI-based device gave a satisfactory PCE of 3.50% with a VOC of 0.837 V, a JSC of 7.64 mA cm−2 and a FF of 0.547. The higher VOC for the device of T-PDI is consistent well with the shallower LUMO energy level compared with that of H-PDI. Meanwhile, the higher JSC and FF for the T-PDI-based device should be mainly attributed to the stronger ability of light absorption, more balanced carrier transport and favorable morphology relative to that of the device based on H-PDI. Though the device performance is not ideal, the new-structure PDI oligomers provide an alternative way to study fullerene-free electron acceptors for PSCs.
200 mL of 2 M Na2CO3, then extracted by CHCl3 three times, next washed with water for two times, and dried over Na2SO4. The crude product was purified by silicon chromatography (petroleum ether (PE):ethyl acetate (EA), v/v(5:1)) to get pure product (1.30 g, 85%). 1H NMR (CDCl3, 400 MHz): δ [ppm]: 8.88 (d, J = 8.0 Hz, 1H), 8.57–8.45 (m, 5H), 8.27 (d, J = 8.0 Hz, 1H), 5.30–5.25 (m, 4H), 2.27–2.18 (m, 4H), 1.88–1.82 (m, 4H), 1.36–1.21 (m, 32H), 0.84–0.81 (m, 12H); 13C NMR (100 MHz, CDCl3): δ [ppm]: 165.76, 144.35, 136.03, 132.75, 130.84, 129.23, 128.45, 127.46, 126.95, 125.31, 124.50, 123.93, 122.22, 118.51, 54.57, 48.17, 32.37, 30.98, 29.25, 29.22, 26.93, 26.90, 22.66, 22.59, 14.07. MS (MALDI-TOF): calcd. for (C50H63N3O4), 769.5; found, 770.6. Elemental anal. calcd for C50H63N3O4: C, 77.99; H, 8.25; N 5.46. Found: C, 77.52; H, 8.33; N 5.83. Synthesis of T-PDI: Compound PDI-NH2 (770 mg, 1.00 mmol), Compound PDA-C6 (570 mg, 1.00 mmol), zinc acetate (0.11 g, 0.6 mmol), and quinoline (10 mL) was stirred at 180 °C for 12 h under Ar atmosphere. After cooling to room temperature, the reaction mixture was added with HCl aqueous solution (2 M, 200 mL), then extracted by CHCl3 three times, and dried over Na2SO4. The crude product was purified by silicon chromatography (petroleum ether (PE):CHCl3, v/v (1:3)) to get pure product (1.11 g, 80%).1H NMR (CDCl3, 400 MHz): δ [ppm]: 8.78–8.74 (m, 12H), 8.67 (d, J = 8.1 Hz, 1H), 8.58 (d, J = 12.0 Hz, 1H), 8.41 (d, J = 8.0 Hz, 1H), 5.20–5.04 (m, 3H), 2.30–2.20 (m, 6H), 1.99–1.79 (m, 6H), 1.45–1.20 (m, 48H), 0.85–0.77 (m, 18H). 13C NMR (CDCl3, 100 MHz): δ [ppm]: 163.50, 136.32, 134.04, 133.93, 132.97, 132.65, 130.20, 129.47, 129.29, 129.25, 128.25, 126.99, 126.80, 126.38, 125.23, 123.91, 123.77, 123.33, 123.26, 122.27, 119.29, 54.90, 54.68, 32.36, 31.77, 31.76, 31.70,
4. Experimental section The synthesis of PDI-NO2 and PDA-C6 were synthesized according to literature [40,50]. Synthesis of PDI-NH2: Compound PDI-NO2 (1.6 g, 2.0 mmol), HCl (4 mL), SnCl2•2H2O (1.80 g, 8.0 mmol) were added into 50 mL tetrahydrofuran (THF), after stirring for 12 h at 90 °C. Then reaction mixture was cooled to room temperature, the reaction mixture was poured into
Fig. 5. Morphology images of the blend films: a) AFM height, b) AFM phase, and c) TEM images for PBDB-T:T-PDI (1:1, w/w) blend film; d) AFM height, e) AFM phase, and f) TEM images for PBDB-T:H-PDI (1:1, w/w) blend film. 360
Dyes and Pigments 163 (2019) 356–362
Z. Luo et al.
29.25, 29.24, 29.17, 26.89, 26.82, 22.61, 22.55, 18.45, 14.07, 14.03. MS (MALDI-TOF): calcd for (C87H96N4O8), 1324.7; found, 1324.5. Elemental anal. calcd for C87H96N4O8: C, 78.82; H, 7.30; N 4.23. Found: C, 79.21; H, 7.39; N 4.58. Synthesis of H-PDI: Compound PDI-NH2 (770 mg, 1.00 mmol), PDA (120 mg, 0.30 mmol), zinc acetate (0.11 g, 0.6 mmol) and quinoline (10 mL) was stirred at 180 °C for 24 h under Ar atmosphere. After cooling to room temperature, the reaction mixture was added to HCl aqueous solution (2 M, 200 mL), then extracted by CHCl3 three times, and dried over Na2SO4. The crude product was purified by silicon chromatography (petroleum ether (PE):CHCl3, v/v(1:4)) to get pure product (360 mg, 66%). 1H NMR (CDCl3, 400 MHz): δ [ppm]: 8.89–8.74 (m, 16H), 8.72–8.69 (m, 2H), 8.61–8.57 (m, 2H), 8.49–8.46 (m, 2H), 5.23–5.04 (m, 4H), 2.32–2.20 (m, 8H), 1.99–1.78 (m, 8H), 1.45–1.11 (m, 64H), 0.85–0.70 (m, 24H). 13C NMR (CDCl3, 100 MHz): δ [ppm]: 163.45, 163.40, 139.80, 135.83, 135.75, 134.06, 132.79, 130.14, 130.08, 129.84, 129.33, 128.90, 128.28, 127.85, 127.07, 126.99, 126.84, 126.69, 125.26, 124.14, 123.84, 123.42, 123.04, 122.78, 58.48, 54.92, 54.73, 32.28, 31.76, 31.71, 29.26, 29.18, 26.89, 26.83, 22.62, 22.56, 18.44, 14.08, 14.05. MS (MALDI-TOF): calcd for (C124H130N6O12), 1894.9; found, 1895.2. Elemental anal. calcd for C124H130N6O12: C, 78.53; H, 6.91; N 4.43. Found: C, 78.24; H, 7.22; N 4.71.
[12] Arias AC, MacKenzie JD, McCulloch I, Rivnay J, Salleo A, Salleo A. Materials and applications for large area electronics: solution-based approaches. Chem Rev 2010;110:3–24. [13] Li W, Hendriks KH, Furlan A, Wienk MM, Janssen RA. High quantum efficiencies in polymer solar cells at energy losses below 0.6 eV. J Am Chem Soc 2015;137:2231–4. [14] Li H, Hwang YJ, Courtright BAE, Eberle FN, Subramaniyan S, Jenekhe SA. Finetuning the 3D structure of nonfullerene electron acceptors toward high-performance polymer solar cells. Adv Mater 2015;27:3266–72. [15] Kwon OK, Uddin MA, Park JH, Park SK, Nguyen TL, Woo HY, Park SY. A high efficiency nonfullerene organic solar cell with optimized crystalline organizations. Adv Mater 2016;28:910–6. [16] Holliday S, Ashraf RS, Wadsworth A, Baran D, Yousaf SA, Nielsen CB, Tan C, Dimitrov SD, Shang Z, Gasparini N, Alamoudi M, Laquai F, Brabec CJ, Salleo A, Durrant JR, McCulloch I. High-efficiency and air-stable P3HT-based polymer solar cells with a new non-fullerene acceptor. Nat Commun 2016;7:11585. [17] Holliday S, Ashraf RS, Nielsen CB, Kirkus M, Röhr JA, Tan C, Collado-Fregoso E, Knall AC, Durrant JR, Nelson J, McCulloch I. A rhodanine flanked nonfullerene acceptor for solution-processed organic photovoltaics. J Am Chem Soc 2015;137:898–904. [18] Wu X, Fu W, Xu Z, Shi M, Liu F, Chen H, Wan J, Russell TP. Spiro linkage as an alternative strategy for promising nonfullerene acceptors in organic solar cells. Adv Funct Mater 2015;25:5954–66. [19] Liu F, Zhou Z, Zhang C, Vergote T, Fan H, Liu F, Zhu X. A thieno[3,4-b]thiophenebased non-fullerene electron acceptor for high-performance bulk-heterojunction organic solar cells. J Am Chem Soc 2016;138:15523–6. [20] Lin Y, Wang J, Zhang Z, Bai H, Li Y, Zhu D, Zhan X. An electron acceptor challenging fullerenes for efficient polymer solar cells. Adv Mater 2015;27:1170–4. [21] Lin Y, Zhan X. Non-fullerene acceptors for organic photovoltaics: an emerging horizon. Mater Horiz 2014;1:470–88. [22] Zhao W, Li S, Yao H, Zhang S, Zhang Y, Yang B, Hou J. Molecular optimization enables over 13% efficiency in organic solar cells. J Am Chem Soc 2017;139:7148–51. [23] Kan B, Feng H, Wan X, Liu F, Ke X, Wang Y, Wang Y, Zhang H, Li C, Hou J. Smallmolecule acceptor based on the heptacyclic benzodi(cyclopentadithiophene) unit for highly efficient nonfullerene organic solar cells. JAm Chem Soc 2017;139:4929–34. [24] Bin H, Gao L, Zhang Z-G, Yang Y, Zhang Y, Zhang C, Chen S, Xue L, Yang C, Xiao M, Li Y. 11.4% Efficiency non-fullerene polymer solar cells with trialkylsilyl substituted 2D-conjugated polymer as donor. Nat Commun 2016;7:13651. [25] Xie D, Liu T, Gao W, Zhong C, Huo L, Luo Z, Wu K, Xiong W, Liu F, Sun Y, Yang C. A novel thiophene-fused ending group enabling an excellent small molecule acceptor for high-performance fullerene-free polymer solar cells with 11.8% efficiency. Sol RRL 2017;1:1700044. [26] Liu Y, Zhang Z, Feng S, Li M, Wu L, Hou R, Xu X, Chen X, Bo Z. Exploiting noncovalently conformational locking as a design strategy for high performance fusedring electron acceptor used in polymer solar cells. J Am Chem Soc 2017;139:3356–9. [27] Qiu N, Zhang H, Wan X, Li C, Ke X, Feng H, Kan B, Zhang H, Zhang Q, Lu Y. A new nonfullerene electron acceptor with a ladder type backbone for high-performance organic solar cells. Adv Mater 2017;29:1604960. [28] Zhang G, Yang G, Yan H, Kim JH, Ade H, Wu W, Xu X, Duan Y, Peng Q. Efficient nonfullerene polymer solar cells enabled by a novel wide bandgap small molecular acceptor. Adv Mater 2017;29:1606054. [29] Shivanna R, Shoaee S, Dimitrov S, Kandappa SK, Rajaram S, Durrant JR, Narayan K. Charge generation and transport in efficient organic bulk heterojunction solar cells with a perylene acceptor. Energy Environ Sci 2014;7:435–41. [30] Meng D, Sun D, Zhong C, Liu T, Fan B, Huo L, Li Y, Jiang W, Choi H, Kim T, Kim JY, Sun Y, Wang Z, Heeger AJ. High-performance solution-processed non-fullerene organic solar cells Based on selenophene-containing perylene bisimide acceptor. J Am Chem Soc 2016;138:375–80. [31] Wu Q, Zhao D, Schneider AM, Chen W, Yu L. Covalently bound clusters of alphasubstituted PDI-rival electron acceptors to fullerene for organic solar cells. J Am Chem Soc 2016;138:7248–51. [32] Meng D, Fu H, Xiao C, Meng X, Winands T, Ma W, Wei W, Fan B, Huo L, Doltsinis NL, Wang Z. Three-bladed rylene propellers with three-dimensional network assembly for organic electronics. J Am Chem Soc 2016;138:10184–90. [33] Duan Y, Xu X, Yan H, Wu W, Li Z, Peng Q. Pronounced effects of a triazine core on photovoltaic performance-efficient organic solar cells enabled by a PDI trimerbased small molecular acceptor. Adv Mater 2017;29:1605115. [34] Yan Q, Cai K, Zhao D. Supramolecular aggregates with distinct optical properties from PDI oligomers of similar structures. Phys Chem Chem Phys 2016;18:1905–10. [35] Lee J, Singh R, Sin DH, Kim HG, Song KC, Cho K. A nonfullerene small molecule acceptor with 3D interlocking geometry enabling efficient organic solar cells. Adv Mater 2016;28:69–76. [36] Liu Y, Mu C, Jiang K, Zhao J, Li Y, Zhang L, Li Z, Lai JY, Hu H, Ma T, Hu R, Yu D, Huang X, Tang BZ, Yan H. A tetraphenyl ethylene core-based 3D structure small molecular acceptor enabling efficient non-fullerene organic solar cells. Adv Mater 2015;27:1015–20. [37] Pho TV, Toma FM, Tremolet de Villers BJ, Wang S, Treat ND, Eisenmenger ND, Su GM, Coffin RC, Douglas JD, Fréchet JMJ, Bazan GC, Wudl F, Chabinyc ML. Decacyclene Triimides: paving the road to universal non-fullerene acceptors for organic photovoltaics. Adv Energy Mater 2014;4:1502109. [38] Hartnett PE, Timalsina A, Matte HS, Zhou N, Guo X, Zhao W, Facchetti A, Chang RP, Hersam MC, Wasielewski MR, Marks TJ. Slip-stacked perylenediimides as an alternative strategy for high efficiency nonfullerene acceptors in organic
Conflicts of interest There are no conflicts to declare. Acknowledgements This work was financially supported by the National Natural Science Foundation of China (NFSC) (No. 21572171) and the Shenzhen Peacock Plan (KQTD20170330110107046). Appendix A. Supplementary data Supplementary data to this article can be found online at https:// doi.org/10.1016/j.dyepig.2018.12.015. References [1] Guo X, Xin H, Kim FS, Liyanage ADT, Jenekhe SA, Watson MD. Thieno[3,4-c]pyrrole-4,6-dione-based donor-acceptor conjugated polymers for solar cells. Macromolecules 2011;44:269–77. [2] Lim DC, Kim KD, Park SY, Hong EM, Seo HO, Lim JH, Lee KH, Jeong Y, Song C, Lee E, Kim YD, Cho S. Towards fabrication of high-performing organic photovoltaics: new donor-polymer, atomic layer deposited thin buffer layer and plasmonic effects. Energy Environ Sci 2012;5:9803–7. [3] Klein MFG, Pasker F, Wettach H, Gadaczek I, Bredow T, Zilkens P, Vohringer P, Lemmer U, Hoger S, Colsmann A. Oligomers comprising 2-phenyl-2H-benzotriazole building blocks for solution-processable organic photovoltaic devices. J Phys Chem C 2012;116:16358–64. [4] Zhao J, Li Y, Yang G, Jiang K, Lin H, Ade H, Ma W, Yan H. Efficient organic solar cells processed from hydrocarbon solvents. Nat Energy 2016;1:15027. [5] You J, Dou L, Yoshimura K, Kato T, Ohya K, Moriarty T, Emery K, Chen C, Gao J, Li G, Yang Y. A polymer tandem solar cell with 10.6% power conversion efficiency. Nat Commun 2013;4:1446. [6] Yu G, Gao J, Hummelen J, Wudl F, Heeger AJ. Polymer photovoltaic cells-enhanced efficiencies via a network of internal donor-acceptor heterojunctions. Science 1995;270:1789–91. [7] Heeger AJ. 25th anniversary article: bulk heterojunction solar cells: understanding the mechanism of operation. Adv Mater 2014;26:10–28. [8] Vandewal K, Tvingstedt K, Gadisa A, Inganäs O, Manca JV. On the origin of the open-circuit voltage of polymer-fullerene solar cells. Nat Mater 2009;8:904–9. [9] Chen C, Chang W, Yoshimura K, Ohya K, You J, Gao J, Hong Z, Yang Y. An efficient triple-junction polymer solar cell having a power conversion efficiency exceeding 11%. Adv Mater 2014;26:5670–7. [10] Huo L, Liu T, Sun X, Cai Y, Heeger AJ, Sun Y. Single-junction organic solar cells based on a novel wide-bandgap polymer with efficiency of 9.7%. Adv Mater 2015;27:2938–44. [11] Zhang Z-G, Qi B, Jin Z, Chi D, Qi Z, Li Y, Wang J. Perylene diimides: a thicknessinsensitive cathode interlayer for high performance polymer solar cells. Energy Environ Sci 2014;7:1966–73.
361
Dyes and Pigments 163 (2019) 356–362
Z. Luo et al.
[44] Li S, Liu W, Li C-Z, Liu F, Zhang Y, Shi M, Chen H, Russell TP. A simple perylene diimide derivative with a highly twisted geometry as an electron acceptor for efficient organic solar cells. J Mater Chem A 2016;4:10659–65. [45] Zhen Y, Wang C, Wang Z. Bay-linked perylene bisimides as promising non-fullerene acceptors for organic solar cells. Chem Commun 2010;46:1926–8. [46] Zang Y, Li C, Chueh CC, Williams ST, Jiang W, Wang Z, Yu J, Jen AKY. Integrated molecular, interfacial, and device engineering towards high-performance non-fullerene based organic solar cells. Adv Mater 2014;26:5708–14. [47] Rajaram S, Shivanna R, Kandappa SK, Narayan K. Nonplanar perylene diimides as potential alternatives to fullerenes in organic solar cells. J Phys Chem Lett 2012;3:2405–8. [48] Gao G, Liang N, Geng H, Jiang W, Fu H, Feng J, Hou J, Feng X, Wang Z. Spiro-fused perylene diimide arrays. J Am Chem Soc 2017;139:15914–20. [49] Wang H, Chen L, Xiao Y. Heterologous perylene diimide arrays: potential nonfullerene acceptors in organic solar cells. J Mater Chem C 2017;5:8875–82. [50] Luo Z, Xiong W, Liu T, Cheng W, Wu K, Sun Y, Yang C. Triphenylamine-cored starshape compounds as non-fullerene acceptor for high-efficiency organic solar cells: tuning the optoelectronic properties by S/Se-annulated perylene diimide. Org Electron 2017;41:166–72.
photovoltaics. J Am Chem Soc 2014;136:16345–56. [39] Hendsbee AD, Sun J-P, Law WK, Yan H, Hill IG, Spasyuk DM, Welch GC. Synthesis, self-assembly, and solar cell performance of N-Annulated perylene diimide nonfullerene acceptors. Chem Mater 2016;28:7098–109. [40] Liang N, Sun K, Zheng Z, Yao H, Gao G, Meng X, Wang Z, Ma W, Hou J. Perylene diimide trimers based bulk heterojunction organic solar cells with efficiency over 7%. Adv. Energy Mater. 2016;6:1600060. [41] Luo Z, Bin H, Liu T, Zhang ZG, Yang Y, Zhong C, Qiu B, Li G, Gao W, Xie D, Wu K, Sun Y, Liu F, Li Y, Yang C. Fine-tuning of molecular packing and energy level through methyl substitution enabling excellent small molecule acceptors for nonfullerene polymer solar cells with efficiency up to 12.54%. Adv Mater 2018;30:1706124. [42] Luo Z, Sun C, Chen S, Zhang Z-G, Wu K, Qiu B, Yang C, Li Y, Yang C. Side-chain impact on molecular orientation of organic semiconductor acceptors: high performance nonfullerene polymer solar cells with thick active layer over 400 nm. Adv Energy Mater 2018;8:1800856. [43] Yan Q, Zhou Y, Zheng Y-Q, Pei J, Zhao D. Towards rational design of organic electron acceptors for photovoltaics: a study based on perylenediimide derivatives. Chem Sci 2013;4:4389–94.
362