Recent advances, challenges and prospects in ternary organic solar cells

Congcong Zhao a, Jiuxing Wang *a, Xuanyi Zhao a, Zhonglin Du *a, Renqiang Yang *b and Jianguo Tang *a
aInstitute of Hybrid Materials, National Center of International Joint Research for Hybrid Materials Technology, National Base of International Sci. & Tech. Cooperation on Hybrid Materials, College of Materials Science and Engineering, Qingdao University, 308 Ningxia Road, Qingdao 266071, China. E-mail: jiuxingwang@qdu.edu.cn
bKey Laboratory of Optoelectronic Chemical Materials and Devices (Ministry of Education), School of Chemical and Environmental Engineering, Jianghan University, Wuhan 430056, China. E-mail: yangrq@qibebt.ac.cn

Received 31st October 2020 , Accepted 22nd December 2020

First published on 23rd December 2020


Abstract

The past decade has seen a tremendous development of organic solar cells (OSCs). To date, high-performance OSCs have boosted power conversion efficiencies (PCEs) over 17%, showing bright prospects toward commercial applications. Compared with binary OSCs, ternary OSCs, by introducing a third component as a second donor or acceptor into the active layer, have great potential in realizing outstanding photovoltaic performance. Herein, a comprehensive review of the recent advances of ternary solar cells is presented. According to the chemical components of active layer materials, we classify the ternary systems into four categories, including polymer/small molecule/small molecule, polymer/polymer/small molecule, all-polymer and all-small-molecule types. The relationships among the photovoltaic materials structure and weight ratio, active layer morphology and photovoltaic performance are systematically analyzed and summarized. The features and design strategies of each category are also discussed and summarized. Key issues and challenges faced in ternary OSCs are pointed out, and potential strategies and solutions are proposed. This review may provide guidance for the field of ternary OSCs.


image file: d0nr07788g-p1.tif

Congcong Zhao

Congcong Zhao received his B.S. degree in Materials Science and Engineering from Liaocheng University in 2018. He is currently pursuing his M.S. degree at the National Center of International Joint Research for Hybrid Materials Technology of Qingdao University under the supervision of Dr Jiuxing Wang. His research interests are the design and synthesis of efficient organic and polymer materials in solar cells.

image file: d0nr07788g-p2.tif

Jiuxing Wang

Jiuxing Wang received his B.S. degree in Polymer Materials and Engineering from Sichuan University in 2011 and Ph.D. degree in Chemical Engineering with Prof. Renqiang Yang from the University of Chinese Academy of Sciences in 2016. He has been an assistant professor at the National Center of International Joint Research for Hybrid Materials Technology of Qingdao University since 2016. His research interests focus on organic and polymer materials for photovoltaic applications.

image file: d0nr07788g-p3.tif

Zhonglin Du

Zhonglin Du received his Ph.D. degree from East China University of Science and Technology (ECUST) in 2017 under the supervision of Professor Xinhua Zhong. He worked as a postdoctoral fellow with professor Dongling Ma at the Institut National de la Recherche Scientifique (INRS) of Canada from 2019 to 2021. Currently, he is an associate professor at the National Center of International Joint Research for Hybrid Materials Technology of Qingdao University. His main research interests include photoelectronic functional organic–inorganic hybrid materials and their application in solar energy conversion, especially photocatalysis and solar cells.

image file: d0nr07788g-p4.tif

Renqiang Yang

Renqiang Yang received his Ph.D. degree from South China University of Technology in 2004 and then conducted his postdoctoral work at the University of California Santa Barbara with Prof. Guillermo C. Bazan. In 2009–2019, he worked as a full professor at Qingdao Institute of Bioenergy and Bioprocess Technology, Chinese Academy of Sciences. Currently he is a full professor at the School of Chemical and Environmental Engineering, Jianghan University. His research interests include molecular design and synthesis of advanced organic opto-electronic materials and device applications thereof.

image file: d0nr07788g-p5.tif

Jianguo Tang

Jianguo Tang received his Ph.D. degree from Shanghai Jiaotong University in 2000, after which he worked at the University of Bayreuth as a visiting professor (2001). Then he joined Prof. L. A. Belfiore's group at Colorado State University as a research fellow (2002–2003). Currently he is the director of the National Center of International Joint Research for Hybrid Materials at Qingdao University. His research interests include hybrid organic–inorganic luminescent materials and solar cells.


1. Introduction

Due to the increasing global consumption of fossil fuels and serious problems of energy shortage, exploiting renewable and environmentally friendly energy resources has become an urgent issue around the world. Solar energy is considered as an ideal alternative energy resource, due to its clean and rich reserve advantages. Silicon-based solar cells have already been commercialized, but complex manufacture and high cost limit their further applications. In the past decades, organic solar cells (OSCs) have become a hot-spot topic in photovoltaic fields owing to their unique advantages such as light weight, low cost, mechanical flexibility and solution processability.1–4 The rapid development of various donor and acceptor materials greatly improves the PCEs of OSCs.5–11

OSC device configurations include single-layer, bilayer, bulk-heterojunction (BHJ), tandem and ternary structures. The first single-layer organic photovoltaic cell (Fig. 1a) was reported by Kearns and Calvin in 1958.12 The maximum power output was only 3 × 10−12 W, due to low-efficiency charge separation. Thereafter, much research attention has been paid to this emerging field.13–15 In 1986, Tang reported the first bilayer solar cell (Fig. 1b).16 However, limited interfacial area and insufficient charge transport restricted the PCE to 1%. In 1995, the concept of BHJ was introduced by Heeger and coworkers to overcome the difficulties of short exciton diffusion length and limited exciton lifetime.17 In a BHJ structure (Fig. 1c), both donor and acceptor materials are mixed together to form a bi-continuous interpenetrating network with large interfacial area for efficient exciton dissociation. To date, the BHJ structure has become the most widely used architecture in OSCs. With the rapid development of materials design and device optimization, the PCEs of single-junction OSCs have exceeded 17%, demonstrating great potential for large-area manufacture and commercialization in the future.18


image file: d0nr07788g-f1.tif
Fig. 1 Schematic of OSC structures: (a) single-layer OSCs, (b) bilayer OSCs, (c) binary BHJ OSCs, (d) tandem OSCs, and (e) ternary OSCs.

Generally, the active layer in the OSC is very thin (∼100 nm), in order to achieve efficient charge collection. However, such a thin blend film leads to insufficient photon harvesting, which limits further enhancement in the performance of single-junction OSCs. To address this issue, several strategies focusing on how to absorb a broad range of the solar spectrum in solar cell devices have been developed. Constructing tandem (Fig. 1d) and ternary device structures (Fig. 1e) is an efficient strategy to extend light absorption and improve PCE. Tandem OSCs consist of two or more single-junction solar cells, which are linked either in series or in parallel.19 The individual sub-cells work independently to absorb photons in different ranges of the solar spectrum. To fully utilize sunlight, the wide- and low-bandgap sub-cells are usually used as front and rear cells, respectively. However, the complicated fabrication of tandem OSCs may increase the cost.27

The active layer of binary OSCs is composed of one donor (D) and one acceptor (A). The working mechanism of binary OSCs can be summarized as follows. Firstly, the photoexcitation of photovoltaic materials in the active layer generates electron–hole pairs, known as excitons. Secondly, the excitons diffuse to the donor–acceptor interface. Thirdly, the excitons dissociate at the donor–acceptor interface to form free electrons and holes. Finally, free holes transport to the anode through the donor material and free electron transport to the cathode through the acceptor material. Ternary OSCs have a composition of either one donor/two acceptors (D[thin space (1/6-em)]:[thin space (1/6-em)]A1[thin space (1/6-em)]:[thin space (1/6-em)]A2) or two donors/one acceptor (D1[thin space (1/6-em)]:[thin space (1/6-em)]D2[thin space (1/6-em)]:[thin space (1/6-em)]A) in the active layer. In recent years, ternary OSCs have emerged as a competitive candidate to achieve high photovoltaic performances. As shown in Fig. 2, the ternary OSCs usually outperform their binary counterparts. Until now, various third components including polymers, fullerene derivatives and small molecules have been used in ternary OSCs. Compared with binary OSCs, ternary OSCs have distinct advantages. (i) The third component can enhance the absorption, which is favourable for increasing the short circuit current density (Jsc).28 However, some studies indicated that it is not always the case unless the ternary blend has quite narrow absorption.29 (ii) The third component can also increase exciton dissociation and charge mobility by building efficient charge transport pathways, resulting in high fill factors (FF) in OSCs.30 (iii) The morphologies of active layers in ternary OSCs could be optimized by the addition of the third component.31 (iv) The Voc in ternary OSCs varies as a function of the blend ratio, and it could be more easily and finely tuned than that in binary devices.32,33 The strategies for improving the photovoltaic performance of binary OSCs, such as tuning energy levels, adjusting donor[thin space (1/6-em)]:[thin space (1/6-em)]acceptor blend ratios and controlling phase separation, can also be applied in ternary OSCs. However, the disadvantages of ternary OSCs should also be noted. For some ternary systems, the introduction of the third component might lead to a more complex morphology, which may give rise to difficulty in morphology control. The Voc value of the ternary OSCs is in between those of the two original binary OSCs, lower than the higher one of the two binary OSCs. Despite these disadvantages, ternary OSCs show great potential for high-performance solar cells. The highest PCE of over 17% of ternary OSCs has been reported recently, showing great potential for constructing highly efficient OSCs.21


image file: d0nr07788g-f2.tif
Fig. 2 The comparison of PCEs of representative categories of binary and ternary OSCs. (a) Blue bar: PBDB-TF[thin space (1/6-em)]:[thin space (1/6-em)]Y6,20 red bar: PBDB-TF[thin space (1/6-em)]:[thin space (1/6-em)]Y6[thin space (1/6-em)]:[thin space (1/6-em)]PC61BM20 (polymer donor/non-fullerene acceptor/fullerene acceptor). (b) Blue bar: PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6,21 red bar: PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6[thin space (1/6-em)]:[thin space (1/6-em)]BTP-M21 (polymer donor/non-fullerene acceptor/non-fullerene acceptor). (c) Blue bar: PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM,22 red bar: PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]DR3TSBDT[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM22 (polymer donor/small-molecule donor/fullerene acceptor). (d) Blue bar: PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]FOIC,23 red bar: PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]TR[thin space (1/6-em)]:[thin space (1/6-em)]FOIC23 (polymer donor/small-molecule donor/non-fullerene acceptor). (e) Blue bar: PM6[thin space (1/6-em)]:[thin space (1/6-em)]Br-ITIC,24 red bar: PM6[thin space (1/6-em)]:[thin space (1/6-em)]J71[thin space (1/6-em)]:[thin space (1/6-em)]Br-ITIC24 (polymer/polymer/small molecule). (f) Blue bar: PTzBI[thin space (1/6-em)]:[thin space (1/6-em)]N2200,25 red bar: PTzBI[thin space (1/6-em)]:[thin space (1/6-em)]PBTA-BO[thin space (1/6-em)]:[thin space (1/6-em)]N220025 (all-polymer). (g) Blue bar: DPP-2TTP[thin space (1/6-em)]:[thin space (1/6-em)]DR3TBDTTF,26 red bar: DPP-2TTP[thin space (1/6-em)]:[thin space (1/6-em)]DR3TBDTTF[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM26 (all small molecule).

Typically, four working mechanisms could be involved in a ternary OSC: charge transfer, energy transfer, parallel-like model and alloy model. In the charge transfer mechanism (Fig. 3a), the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) energy levels of the third component are located between those of the host donor and acceptor, forming a cascade energy level alignment.34 This cascade energy level can facilitate charge separation and charge transfer at the D[thin space (1/6-em)]:[thin space (1/6-em)]A interface because the third component can function as a charge relay for electron and hole transport.35 As for the energy transfer mechanism (Fig. 3b), the third component does not directly generate free charges but transfers the energy to either a donor or an acceptor through Dexter or Förster resonance energy transfer (FRET). In some cases, the mechanisms of charge transfer and energy transfer can simultaneously exist in a ternary blend. In the parallel-like model (Fig. 3c), the third component forms its independent hole-transport channels.36 In the alloy model (Fig. 3d), the third component electronically couples with the host donor to form a new charge transfer state.37


image file: d0nr07788g-f3.tif
Fig. 3 Schematic of the working mechanisms in ternary OSCs: (a) charge transfer, (b) energy transfer, (c) parallel-like model and (d) alloy model.

According to the possible combinations of three components, all the ternary OSCs can be classified into four categories including polymer/small molecule/small molecule, polymer/polymer/small molecule, all-polymer and all-small-molecule types (Fig. 4). In this review, we have summarized the recent progress of ternary OSCs and analyzed the relationships among the molecular structure, active layer morphology and photovoltaic performance. Especially, we have emphasized the critical roles of the third component in ternary blends. Finally, the potential challenges and promising outlook for ternary OSCs are also discussed.


image file: d0nr07788g-f4.tif
Fig. 4 The classification of ternary OSCs.

2. Recent advances in ternary organic solar cells

2.1 Polymer/small molecule/small-molecule OSCs

In this type of ternary OSC, a small molecule as a third component adding to the active layer can not only extend the absorption which is beneficial for harvesting more photons, but also improve the crystallinity of the polymer and tune active layer morphology through the effect of small molecules.38–40 Two non-fullerene acceptors or mixed acceptors composed of a fullerene acceptor and an non-fullerene acceptor could be used in this type of ternary device, and thereby it can be classified into four categories, which are polymer donor/non-fullerene acceptor/fullerene acceptor, polymer donor/non-fullerene acceptor/non-fullerene acceptor, polymer donor/small-molecule donor/fullerene acceptor, and polymer donor/small-molecule donor/non-fullerene acceptor types.
(a) Polymer donor/non-fullerene acceptor/fullerene acceptor OSCs. Polymer donor/non-fullerene acceptor/fullerene acceptor ternary OSCs combine the advantages of the high electron mobility of the fullerene acceptor and broad absorption of the narrow bandgap non-fullerene acceptor. Over the past few years, many effective polymer donor/non-fullerene acceptor/fullerene acceptor ternary OSCs with PCEs over 10% have been developed.41–44 In these systems, fullerene derivatives such as phenyl-C61-butyric acid methyl ester (PC61BM) and phenyl-C71-butyric acid methyl ester (PC71BM) have been commonly used as the third component to strengthen the light-harvesting capability, facilitate charge dissociation and tune the morphology of active layers, and thus lead to enhanced Jsc and PCE.45–48 The structures of donors and acceptors for polymer donor/non-fullerene acceptor/fullerene acceptor ternary OSCs are shown in Fig. 5, and the photovoltaic parameters are summarized in Table 1.
image file: d0nr07788g-f5.tif
Fig. 5 The structures of donors and acceptors for polymer donor/non-fullerene acceptor/fullerene acceptor ternary OSCs.
Table 1 Photovoltaic parameters of ternary OSCs with polymer donor/non-fullerene acceptor/fullerene acceptor blends
Polymer (D) Non-fullerene A1 Fullerene A2 (third component) Weight ratio (D[thin space (1/6-em)]:[thin space (1/6-em)]A1[thin space (1/6-em)]:[thin space (1/6-em)]A2) V oc (V) J sc (mA cm−2) FF (%) PCE (%) Working mechanism Ref.
PBDB-TF Y6 PC61BM 1[thin space (1/6-em)]:[thin space (1/6-em)]1.2[thin space (1/6-em)]:[thin space (1/6-em)]0.2 0.845 25.4 77 16.5 20
PM6 Y6 PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2 0.828 23.57 72.03 14.06 Energy transfer 50
PM6 Y6 PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2 0.861 25.1 77.2 16.7 Energy transfer 51
PBT1-C IT-2F PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2 0.892 18.19 75.1 12.19 Energy transfer 52
PBT1-C ITIC-2Cl IC60BA 1[thin space (1/6-em)]:[thin space (1/6-em)]0.8[thin space (1/6-em)]:[thin space (1/6-em)]0.2 0.89 19.58 76.8 13.4 Energy transfer 53
PTB7-Th 3TT-FIC PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]1.2[thin space (1/6-em)]:[thin space (1/6-em)]0.3 0.669 27.73 73 13.54 54
PBDTTPD-HT ITIC PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]0.9[thin space (1/6-em)]:[thin space (1/6-em)]0.6 0.95 17.38 74 12.09 Parallel-like model 55
PTBTz-2 IT-M PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2 0.928 18.7 70.78 12.28 Energy transfer 56
P3TCO-1 ITIC PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2 0.941 18.16 66.78 11.41 Charge transfer 57
J52 IEICO-4F PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]0.9[thin space (1/6-em)]:[thin space (1/6-em)]0.6 0.698 22.70 67.4 10.68 58
PTBTz-2 ITIC PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.35 0.89 20.75 60.94 11.26 59
PBDB-T ITM Bis-PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2 0.952 17.39 73.7 12.20 Energy transfer 60
PTB7-Th COi8DFIC PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]1.05[thin space (1/6-em)]:[thin space (1/6-em)]0.45 0.70 16.21 60.2 14.08 61
PPBDTBT ITIC PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]1.2[thin space (1/6-em)]:[thin space (1/6-em)]0.8 0.898 16.82 68.26 10.41 Charge transfer 62


In 2019, Zou's group first reported the non-fullerene acceptor Y, which promotes the progress of OSCs with a high PCE over 15%, and Y6-based systems have been widely applied and investigated to date.49 To further improve the PCEs of Y6-based OSCs, Hou's group incorporated PC61BM into a PBDB-TF[thin space (1/6-em)]:[thin space (1/6-em)]Y6 blend to fabricate ternary OSCs.20 After the incorporation of 14% PC61BM, the hole mobilities (μh) and electron mobilities (μe) of the PBDB-TF[thin space (1/6-em)]:[thin space (1/6-em)]Y6[thin space (1/6-em)]:[thin space (1/6-em)]PC61BM (1[thin space (1/6-em)]:[thin space (1/6-em)]1.2[thin space (1/6-em)]:[thin space (1/6-em)]0.2) blend were 2.5 and 2.1 × 10−4 cm2 V−1 s−1, respectively, leading to a balanced charge transport process. Furthermore, PC61BM can inhibit the high tendency to crystallize and aggregate in Y6, which resulted in a high electroluminescence quantum efficiency of 1.9 × 10−4 and a reduced non-radiative energy loss (Eloss) of 0.22 eV. On introducing PC61BM to the PBDB-TF[thin space (1/6-em)]:[thin space (1/6-em)]Y6 blend, atomic force microscopy (AFM) measurements showed that the optimal ternary blend exhibited a root-mean-square (RMS) roughness of 2.64 nm, indicating that the large aggregates in the PBDB-TF[thin space (1/6-em)]:[thin space (1/6-em)]Y6 blend were decreased. Benefiting from these advantages, an open-circuit voltage (Voc) of 0.845 V, a Jsc of 25.4 mA cm−2, an FF of 77.0% and a high PCE of 16.5% were achieved. Ge et al. reported a flexible ITO-free ternary OSC based on a PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM blend.50 As the 2D GIWAXS patterns show in Fig. 6, the introduction of the PC71BM component into PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6 blends resulted in obvious (010) stacking peaks in an out-of-plane (OOP) direction at 1.744 Å−1 (d = 3.60 Å), which demonstrated that the face-on orientation was still maintained in the ternary blend film. The AFM image of the PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM blend showed a well-mixed morphology with an RMS roughness of 1.25 nm, which was slightly higher than that of the PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6 blend (0.964 nm). As shown in Fig. 7d, the flexible devices exhibited excellent mechanical performance. When the cylinder radius was 7.5 mm in the bending test, the PCE of the flexible devices could still remain at ∼98%. The optimal photovoltaic performances were: Voc = 0.828 V, Jsc = 23.57 mA cm−2, FF = 72.03%, and PCE = 14.06%. Zhan et al. added PC71BM into the PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6 system to increase the Voc and phase purity of the devices.51 The lowest unoccupied molecular orbital (LUMO) energy level of PC71BM was higher than that of Y6, and lower than that of PM6. On blending PC71BM with a ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2 for PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM, the device Voc was increased from 0.848 to 0.861 V. Transmission electron microscopy (TEM) and AFM measurements demonstrated that the phase purity was also increased from 0.883 in the binary blend to ∼1 in the ternary blend. Due to the stronger absorption of PC71BM in the 300–700 nm wavelength region, a Jsc of 25.1 mA cm−2, an FF of 77.2%, and an impressive PCE of 16.7% were achieved.


image file: d0nr07788g-f6.tif
Fig. 6 GIWAXS images of PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM blend films with different ratios: (a) 1[thin space (1/6-em)]:[thin space (1/6-em)]1.2[thin space (1/6-em)]:[thin space (1/6-em)]0; (b) 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2; and (c) 1[thin space (1/6-em)]:[thin space (1/6-em)]0[thin space (1/6-em)]:[thin space (1/6-em)]1.2; (d) GIWAXS intensity profiles along the in-plane (dotted line) and out-of-plane (solid line) directions; and AFM images (2 μm × 2 μm) of films with blend ratios of (e) 1[thin space (1/6-em)]:[thin space (1/6-em)]1.2[thin space (1/6-em)]:[thin space (1/6-em)]0 and (f) 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2. Reproduced with permission.50 Copyright 2019, Wiley.

image file: d0nr07788g-f7.tif
Fig. 7 (a) The schematic structure of the flexible device; (b) JV curves; (c) EQE spectrum; and (d) bending performance with different bending diameters (10, 5, and 2 mm) based on the ternary devices with the PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM photoactive layer in an optimized ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2. Reproduced with permission.50 Copyright 2019, Wiley.

Apart from Y6-based systems, fullerene derivatives were applied in conjunction with indacenodithiophene (IDT)-based acceptors in ternary OSCs. In 2019, Sun's group reported a high efficiency ternary OSC with improved ambient stability by adding PC71BM into a PBT1-C[thin space (1/6-em)]:[thin space (1/6-em)]IT-2F blend.52 The three materials showed complementary absorptions across the whole visible region. AFM images indicated that the crystalline aggregates of IT-2F were suppressed after adding PC71BM with decreased RMS roughness from 1.69 to 1.11 nm, leading to a reduced trap-assisted recombination. The Voc of the ternary device was improved because of the higher lying LUMO energy level of PC71BM than that of IT-2F. At a ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2 of PBT1-C[thin space (1/6-em)]:[thin space (1/6-em)]IT-2F[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM, a maximum PCE of 12.19% was achieved, together with a Voc of 0.892 V, a Jsc of 18.19 mA cm−2, and an FF of 75.1%. To study the morphology of ternary OSCs, the same group used an indene-C60 bisadduct (IC60BA) as the second acceptor in the PBT1-C[thin space (1/6-em)]:[thin space (1/6-em)]ITIC-2Cl blend.53 The addition of 20% IC60BA disturbed the intermolecular π–π stacking of crystalline ITIC-2Cl and then improved the film morphology. As a consequence, the RMS roughness of the film decreased from 1.70 to 1.28 nm (Fig. 8), leading to balanced charge mobility. The optimal device exhibited an impressive PCE of 13.4%, with a Jsc of 19.58 mA cm−2, a Voc of 0.89 V, and a high FF of 76.8%. Chen et al. reported that the introduction of PC71BM could increase the PCE of the PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]3TT-FIC binary OSC from 12.21% to 13.54%.54 The enhancement of PCE was mainly ascribed to the improved external quantum efficiency (EQE) response in the visible range of 300–700 nm with the addition of PC71BM. Meanwhile, the μe and μh of the ternary device were 2.28 × 10−4 and 2.16 × 10−4 cm2 V−1 s−1, respectively, which were slightly larger and more balanced than those of the binary device (1.67 × 10−4 and 2.09 × 10−4 cm2 V−1 s−1). Jang et al. achieved high-performance ternary devices based on PBDTTPD-HT[thin space (1/6-em)]:[thin space (1/6-em)]ITIC[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM blends.55 The maximum charge generation rate (Gmax) value of 1.09 × 1028 for the ternary device was higher than that of the PBDTTPD-HT[thin space (1/6-em)]:[thin space (1/6-em)]ITIC binary device (9.06 × 1027 m−3 s−1) and the PBDTTPD-HT[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM binary device (7.25 × 1027 m−3 s−1). In addition, the incorporated PC71BM enhanced the ordering of ITIC and was uniformly dispersed with sufficient continuity. This favorable morphology resulted in significantly higher charge generation and extraction of the ternary device. As a result, a Voc of 0.95 V, a Jsc of 17.38 mA cm−2, an FF of 74% and a PCE of 12.09% were achieved. Recently, Xia et al. studied the synergistic effect of PC71BM and IT-M in a PTBTz-2[thin space (1/6-em)]:[thin space (1/6-em)]IT-M[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM ternary OSC.56 2D GIWAXS patterns showed that the diffraction signal at 3.50 nm−1 was enhanced by doping 20% PC71BM in the blend film, indicating an increased crystallinity of IT-M. Meanwhile, TEM measurements suggested that a favorable phase separation was achieved, which was beneficial for charge generation and collection in OSCs. Therefore, the ternary OSCs exhibited a desirable PCE of up to 12.28%, a Voc of 0.928 V, a Jsc of 18.70 mA cm−2 and an FF of 70.78%. An et al. synthesized a new polymer, P3TCO-1, with a twisted conjugated backbone to enhance the solubility in organic solvents at room temperature.57 P3TCO-1 has a wide optical bandgap of 1.90 eV and a deep HOMO energy level of −5.39 eV. When the two acceptors of ITIC and PC71BM were combined, the P3TCO-1-based ternary OSC exhibited a high PCE of 11.41% with a simultaneously improved Jsc of 18.16 mA cm−2 and an FF of 66.78%. Cao et al. fabricated ternary OSCs based on a J52[thin space (1/6-em)]:[thin space (1/6-em)]IEICO-4F[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM ternary blend.58 The addition of PC71BM into the J52[thin space (1/6-em)]:[thin space (1/6-em)]IEICO-4F binary blend contributed to the suppressed trap-assisted recombination and enhanced charge extraction. Therefore, the PCE of the ternary OSCs reached up to 10.68%, accompanied by a Jsc of 22.70 mA cm−2 and an FF of 67.4%. Yang et al. used PTBTz-2 as a polymer donor, and ITIC and PC71BM as two acceptors to fabricate ternary OSCs.59 The performance of ternary devices was enhanced when 35% PC71BM was added into the PTBTz-2[thin space (1/6-em)]:[thin space (1/6-em)]ITIC blend, resulting in an optimized PCE of 11.26% with a high Jsc of 20.75 mA cm−2. The study showed that the addition of PC71BM can not only affect the structural character of the donor but also enhance the crystallization of ITIC in the in-plane direction. Hou and coworkers introduced bis-PC71BM as the third component into a PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]IT-M blend and fabricated ternary OSCs.60 At a ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2 for PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]IT-M[thin space (1/6-em)]:[thin space (1/6-em)]bis-PC71BM, the ternary device exhibited an outstanding PCE of 12.2%. The enhanced PCE was due to the complementary light absorption of bis-PC71BM in the short wavelength range from 380 to 550 nm and effective exciton diffusion at the PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]bis-PC71BM and PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]IT-M blend interface. Ding et al. reported a highly efficient ternary OSC based on PTB7-Th, COi8DFIC and PC71BM.61 The PC71BM enlarged light absorption at short wavelengths, facilitated electron transfer from PTB7-Th to COi8DFIC and optimized the morphology of the active layer, leading to a high Jsc of 16.21 mA cm−2 and an FF of 60.2%. When the ratio of PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]COi8DFIC[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM was 1[thin space (1/6-em)]:[thin space (1/6-em)]1.05[thin space (1/6-em)]:[thin space (1/6-em)]0.45, the ternary solar cells gave a PCE of 14.08%. Bo and coworkers employed PC71BM and ITIC as small-molecule acceptors to fabricate PPBDTBT-based ternary devices.62 PC71BM can disrupt the molecule stacking of ITIC to reduce its aggregation, leading to increased D/A interface area in the blend films. Meanwhile, the cascade-like energy levels of the three components can facilitate charge transfer at the D/A interfaces. Consequently, at the PPBDTBT[thin space (1/6-em)]:[thin space (1/6-em)]ITIC[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1.2[thin space (1/6-em)]:[thin space (1/6-em)]0.8, the ternary solar cell gave a PCE of 10.41%, a Voc of 0.898 V, a Jsc of 16.82 mA cm−2 and an FF of 68.26%.


image file: d0nr07788g-f8.tif
Fig. 8 (a)–(e) TEM and (f)–(j) AFM height images of ternary blends with different ICBA contents. Reproduced with permission.53 Copyright 2018, Wiley.

As analyzed above, the use of fullerene derivatives as the third component in the non-fullerene blends can often improve the device performance, which can be ascribed to: (a) enhanced absorption in the short-wavelength region by the fullerene derivatives; (b) the intrinsic advantages of high electron affinity, high electron mobility and isotropic charge transport, enabling fullerene derivatives to promote efficient charge separation in blend films; and (c) the addition of fullerene derivatives often inhibiting the aggregation of acceptors, showing a decreased phase separation size.

(b) Polymer donor/non-fullerene acceptor/non-fullerene acceptor OSCs. Most recently, great efforts have been devoted to the non-fullerene ternary OSCs due to the advantages of non-fullerene acceptors, such as easily tunable energy levels, strong light absorption, and high stability.22,63–67 Thus many highly efficient non-fullerene acceptors have been developed, such as indacenodithiophene (IDT)-based, arylene imide-based and diketopyrrolopyrrole-based small molecules.68,69 The new emerging highly efficient acceptors show various absorption spectra and energy levels, suggesting great potential in constructing high-performance ternary OSCs. In a non-fullerene ternary OSC, the two acceptors with similar structures are usually selected to form an alloy-like composite due to their good compatibility.70–74 The energy levels and crystallinity of the composite can be further adjusted by changing the ratios of blends. The structures of donors and acceptors for polymer donor/non-fullerene acceptor/non-fullerene acceptor ternary OSCs are shown in Fig. 9, and the photovoltaic parameters are summarized in Table 2.
image file: d0nr07788g-f9.tif
Fig. 9 The structures of donors and acceptors for polymer donor/non-fullerene acceptor/non-fullerene acceptor ternary OSCs.
Table 2 Photovoltaic parameters of ternary OSCs with polymer donor/non-fullerene acceptor/non-fullerene acceptor blends
Polymer (D) Non-fullerene A1 Non-fullerene A2 (third component) Weight ratio (D[thin space (1/6-em)]:[thin space (1/6-em)]A1[thin space (1/6-em)]:[thin space (1/6-em)]A2) V oc (V) J sc (mA cm−2) FF (%) PCE (%) Working mechanism Ref.
PM6 Y6 BTP-M 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2 0.875 26.56 73.46 17.03 Energy transfer 21
PM6 Y6 BTF 1[thin space (1/6-em)]:[thin space (1/6-em)]1.2[thin space (1/6-em)]:[thin space (1/6-em)]0.1 0.853 26.11 74.22 16.53 Energy transfer 75
J71 ITIC BTF 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.1 0.952 18.48 70.03 12.35 Energy transfer 75
PM6 Y6 3TP3T-4F 1[thin space (1/6-em)]:[thin space (1/6-em)]1.02[thin space (1/6-em)]:[thin space (1/6-em)]0.18 0.85 26.10 75.4 16.7 Energy transfer 76
PM6 Y6 3TP3T-IC 1[thin space (1/6-em)]:[thin space (1/6-em)]1.02[thin space (1/6-em)]:[thin space (1/6-em)]0.18 0.86 25.2 71.6 15.6 Energy transfer 76
PBDB-T-SF Y6 ITCT 1[thin space (1/6-em)]:[thin space (1/6-em)]1.1[thin space (1/6-em)]:[thin space (1/6-em)]0.1 0.885 24.75 73.67 16.14 77
PBDB-T-2Cl IXIC-4Cl IT-4F 1[thin space (1/6-em)]:[thin space (1/6-em)]0.3[thin space (1/6-em)]:[thin space (1/6-em)]0.7 0.847 23.65 74.7 14.96 Energy transfer 79
J71 T6Me IT-4F 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 0.82 22.65 70.82 13.16 Energy transfer 80
PBDB-T-2F IT-2F FTTCN 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2 0.94 18.18 76.03 12.99 Alloy model 78
PBDB-T IEICF-DMOT IEICO-4F 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.1 0.86 23.31 69.75 14.00 81
PBDB-T IE4F-S IBC-F 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2 0.887 22.83 74.4 15.06 Energy transfer 6
PBDB-T IT-M DTF-IC 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.1 0.94 17.95 72 12.14 Energy transfer 82
PBDB-TF IT4F O6T-4F 1[thin space (1/6-em)]:[thin space (1/6-em)]0.85[thin space (1/6-em)]:[thin space (1/6-em)]0.15 0.83 21.60 74.6 13.37 Energy transfer 65
PBDBT-SF IDIC IT4F 1[thin space (1/6-em)]:[thin space (1/6-em)]0.6[thin space (1/6-em)]:[thin space (1/6-em)]0.4 0.917 16.89 74.35 11.52 84
PBDB-T Y16 MeIC1 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.15 0.909 22.76 68.22 14.11 Energy transfer 85
PBDB-T ITIC ITIF 1[thin space (1/6-em)]:[thin space (1/6-em)]0.95[thin space (1/6-em)]:[thin space (1/6-em)]0.05 0.865 17.31 70.3 10.53 Energy transfer 83
PBDB-T IEICO-4F 4TIC 1[thin space (1/6-em)]:[thin space (1/6-em)]0.6[thin space (1/6-em)]:[thin space (1/6-em)]0.4 0.82 21.52 65.9 11.63 Energy transfer 86
PBTA-PS IDIC-C4Ph 6TIC 1[thin space (1/6-em)]:[thin space (1/6-em)]0.7[thin space (1/6-em)]:[thin space (1/6-em)]0.3 0.84 22.33 75.94 14.24 Energy transfer 87
PBTA-PS ITIC IT4F 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2 0.91 19.60 74.45 13.27 Energy transfer 88
PM6 C8-SF Y-MODF 1[thin space (1/6-em)]:[thin space (1/6-em)]0.9[thin space (1/6-em)]:[thin space (1/6-em)]0.3 0.845 20.88 75.87 13.39 Energy transfer 89
J61 BT-IC IDIC 0.9[thin space (1/6-em)]:[thin space (1/6-em)]0.8[thin space (1/6-em)]:[thin space (1/6-em)]0.2 0.88 17.91 68.37 10.76 Energy transfer 90
PTB7-Th COi8DFIC BDTThIT-4F 1[thin space (1/6-em)]:[thin space (1/6-em)]0.8[thin space (1/6-em)]:[thin space (1/6-em)]0.2 0.721 24.88 72.9 13.08 91
PTB7-Th F8IC IDT-2BR 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.2 0.687 25.1 69.9 12.1 Energy transfer 92
PM6 IT-4F N7IT 1[thin space (1/6-em)]:[thin space (1/6-em)]0.7[thin space (1/6-em)]:[thin space (1/6-em)]0.3 0.890 22.54 74.8 15.02 Energy transfer 93


In several studies, benzothiadiazole (BT)-based non-fullerene acceptors were rationally employed as the third components in ternary OSCs. In 2020, Chen's group reported a new non-fullerene acceptor named BTP-M as the third component for the PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6 binary system.21 As an analogue of Y6, electron-donating methyl groups were introduced at the terminals of BTP-M to elevate the LUMO energy level of the acceptor alloy (Y6[thin space (1/6-em)]:[thin space (1/6-em)]BTP-M). With the addition of 20% BTP-M, the Voc was increased from 0.844 to 0.875 V, and Eloss was reduced from 0.49 to 0.45, due to the higher lying LUMO level of the acceptor alloy (Fig. 10). Meanwhile, the Jsc values were increased from 24.67 to over 26.5 mA cm−2, which could be ascribed to the optimized morphology. As for the FF, ternary devices retained similar values (74.96%) to that of the PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6 binary device, implying that the acceptor alloy had good charge transport properties. The ternary device also presented enhanced EQE over 86% at around 630 nm, much better than those of the binary devices. As a result, the optimal ternary device exhibited the best PCE of 17.03%. Moreover, the authors found that ternary devices can maintain a PCE over 14% when the thickness of the active layers was varied from 120 to 300 nm. By incorporating a small molecule of 4,7-bis(5-bromothiophen-2-yl)-5,6-difluorobenzo[c][1,2,5]thiadiazole (BTF) into J71[thin space (1/6-em)]:[thin space (1/6-em)]ITIC and PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6 binary blends, Zheng et al.75 fabricated efficient ternary OSCs with improved PCEs of 12.35% and 16.53%, respectively. The introduction of 10% BTF into the J71[thin space (1/6-em)]:[thin space (1/6-em)]ITIC blend improved light harvesting in the wavelength region from 300 to 500 nm, which resulted in enhanced EQEs for the ternary OSCs. Morphological studies illustrated that BTF can enhance the crystallinity of J71 to form more distinct and denser interpenetrating nano-fibrillar networks. As a result, balanced μh (1.41 × 10−4 cm2 V−1 s−1) and μe (1.39 × 10−4 cm2 V−1 s−1) with a μh/μe ratio of 1.01 were observed, which contributed to the high Jsc (18.48 mA cm−2) and FF (70.03%). Moreover, the Voc of J71[thin space (1/6-em)]:[thin space (1/6-em)]ITIC[thin space (1/6-em)]:[thin space (1/6-em)]BTF ternary devices was enhanced from 0.941 to 0.952 V, owing to the higher lying LUMO energy level of BTF than that of ITIC. When BTF was added into the PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6 system, and the binary device also showed an excellent PCE of 16.53% with an increased Voc of 0.853 V, a Jsc of 26.11 mA cm−2 and an FF of 74.22%.


image file: d0nr07788g-f10.tif
Fig. 10 Energy level diagram of materials used in ternary OSCs (Y6[thin space (1/6-em)]:[thin space (1/6-em)]BTP-M = 4[thin space (1/6-em)]:[thin space (1/6-em)]1, by wt). Reproduced with permission.21 Copyright 2020, Royal Society of Chemistry.

In many studies, IDT-based acceptors were employed in non-fullerene systems. Sun et al. incorporated two non-fullerene small-molecule acceptors 3TP3T-4F and 3TP3T-IC as the third component into PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6 binary blends, respectively.76 Both the guest acceptors formed energy cascades with PM6 and Y6, which could promote charge transfer in ternary OSCs. Compared with 3TP3T-IC, 3TP3T-4F showed a more excellent compatibility with PM6 and Y6. AFM measurements indicated that the incorporation of 3TP3T-4F led to a tendency of forming well-mixed phases in the ternary blend without disturbing the favorable morphology. The 2D GIWAXS patterns showed that the intermolecular π–π stacking distance of the PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6[thin space (1/6-em)]:[thin space (1/6-em)]3TP3T-4F ternary blend was shorter than that of the PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6 film, leading to more favorable electron transport. At the optimal blend ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1.02[thin space (1/6-em)]:[thin space (1/6-em)]0.18 for PM6[thin space (1/6-em)]:[thin space (1/6-em)]Y6[thin space (1/6-em)]:[thin space (1/6-em)]3TP3T-4F, the corresponding ternary device showed a PCE of 16.7%, which was higher than that (15.6%) of the device containing 3TP3T-IC. In 2019, Zhan et al. achieved a new ternary system based on PBDB-T-SF[thin space (1/6-em)]:[thin space (1/6-em)]Y6[thin space (1/6-em)]:[thin space (1/6-em)]ITCT, which showed an impressive efficiency of 16.14%.77 ITCT has a higher LUMO energy level than Y6, which contributed to a higher Voc when blended with the PBDB-T-SF[thin space (1/6-em)]:[thin space (1/6-em)]Y6 binary. Therefore, Voc values of 0.98 V and 0.86 V were obtained from the ITCT and Y6 binary devices, respectively. After a small amount of Y6 was replaced with ITCT (PBDB-T-SF[thin space (1/6-em)]:[thin space (1/6-em)]Y6[thin space (1/6-em)]:[thin space (1/6-em)]ITCT = 1[thin space (1/6-em)]:[thin space (1/6-em)]1.1[thin space (1/6-em)]:[thin space (1/6-em)]0.1), larger phase domains and increased phase crystallinity of the active layer were observed. The fine film morphology enabled balanced charge mobilities and reduced recombination, which resulted in improved FF (73.67%) and Jsc (24.75 mA cm−2) values. When a non-fullerene acceptor IT-4F was introduced as the third component into the PBDB-T-2Cl[thin space (1/6-em)]:[thin space (1/6-em)]IXIC-4Cl binary blend, a high PCE of 14.96% was achieved by Chen et al.79 IT-4F showed an absorption edge of ∼800 nm, complementing the absorption spectrum of PBDB-T-2Cl and IXIC-4Cl. More importantly, IT-4F can suppress the excessive crystallization of IXIC-4Cl and decrease the pure phase domain of the active layer, which reduced charge recombination and balanced charge transport. By using different ratios of the ternary blend, the enhancement in Voc showed an almost linear dependence on the content of IT-4F, indicating the suitable miscibility of the two alloy-like acceptors. At a weight ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]0.3[thin space (1/6-em)]:[thin space (1/6-em)]0.7 for PBDB-T-2Cl[thin space (1/6-em)]:[thin space (1/6-em)]IXIC-4Cl[thin space (1/6-em)]:[thin space (1/6-em)]IT-4F, the optimal ternary device exhibited a high PCE of 14.96%, together with a Voc of 0.847 V, a Jsc of 23.655 mA cm−2 and an FF of 74.7%.

Zhang's group also incorporated IT-4F into the J71[thin space (1/6-em)]:[thin space (1/6-em)]T6Me blend to fabricate a series of ternary OSCs.80 The EQE values of ternary OSCs in the wavelength range from 600 to 750 nm increased with the increase of IT-4F content, which was derived from the contribution of IT-4F on photon harvesting and optimized phase separation. Furthermore, the similar LUMO energy levels of IT-4F (−3.99 V) and T6Me (−3.94 V) could be beneficial for electron transfer in the ternary active layers. After adding 50% IT-4F into the J71[thin space (1/6-em)]:[thin space (1/6-em)]T6Me blend, the RMS roughness values of the blend films slightly decreased from 1.04 to 0.76 nm. As a result, the PCE of ternary OSCs reached 13.16% with a constant Voc of 0.82 V, resulting from the simultaneously optimized Jsc of 22.65 mA cm−2 and FF of 70.82%. Zhang's group incorporated a non-fullerene acceptor FTTCN into the PBDB-T-2F[thin space (1/6-em)]:[thin space (1/6-em)]IT2F blend to fabricate ternary OSCs.78 The HOMO and LUMO energy levels of the blend films (Fig. 11b) increased monotonically along with the increase of FTTCN content, indicating that an FTTCN and IT-2F alloyed acceptor (Fig. 11d) may be formed due to their good compatibility. The increased LUMO energy levels can well explain the increased Voc of ternary OSCs, illustrating the formation of mixed energy levels for two acceptors. Meanwhile, the nano-fibrillar structure was obviously increased in the ternary blend films because the incorporation of FTTCN content can finely adjust the molecular arrangement of PBDB-T-2F. By blending 20% FTTCN in binary blends, the optimized ternary OSCs achieved a high PCE of 12.99% along with a Jsc of 18.18 mA cm−2, a Voc of 0.940 V and an FF of 76.03%.


image file: d0nr07788g-f11.tif
Fig. 11 (a) Vocs of PSCs’ dependence on FTTCN content in acceptors; (b) HOMO and LUMO levels of FTTCN[thin space (1/6-em)]:[thin space (1/6-em)]IT-2F blend films’ dependence on FTTCN content; (c) PL spectra of IT-2F[thin space (1/6-em)]:[thin space (1/6-em)]FTTCN blend films with varied FTTCN contents in acceptors under 700 nm light excitation; and (d) schematic charge transport process in ternary active layers. Reproduced with permission.78 Copyright 2019, Elsevier.

Narrow bandgap non-fullerene acceptors could provide complementary absorption and improve light harvesting ability for ternary OSCs. Zhan and coworkers reported a high-performance ternary OSC with a PCE of 14% by incorporating IEICO-4F into a PBDBT[thin space (1/6-em)]:[thin space (1/6-em)]IEICF-DMOT binary blend.81 Compared with IEICO-4F, IEICF-DMOT possessed two shorter side chains (methoxyl) on the π-bridge. As a result, IEICF-DMOT exhibited much broadened absorption with a larger absorption coefficient (2.9 × 10−5 M−1 cm−1), a higher LUMO energy level (−3.85 eV), a larger optical bandgap (1.39 eV) and reduced crystallinity. When using the crystalline IEICO-4F as the near infrared (NIR) absorber, the PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]IEICF-DMOT[thin space (1/6-em)]:[thin space (1/6-em)]IEICO-4F (1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.1) ternary blend shows increased crystallinity for both donor and acceptor phases, achieving an increased Jsc of 23.31 mA cm−2, an FF of 69.75%, and a promising PCE of 14%. Li et al. found that the addition of a new small-molecule IBC-F into a PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]IE4F-S blend can improve the efficiency and stability of the resulting ternary devices.6 2D-GIWAXS patterns showed that the incorporation of 20% IBC-F could enhance the molecular interactions to facilitate face-on orientation in the blend film. AFM and TEM measurements illustrated that the ternary blend exhibited more fibrillar textures than the binary blend. Therefore, enhanced thermal and photoinduced stability was realized in the ternary device. The PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]IE4F-S[thin space (1/6-em)]:[thin space (1/6-em)]IBC-F based ternary device retained 81.0% and 85.5% of its initial PCE (15.06%) after continuous thermal aging and illumination under an inert atmosphere for 1000 h, respectively, while the PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]IE4F-S based binary device retained 71.7% and 73.6% of its initial PCE (13.70%), respectively. Ge et al. introduced a non-fullerene acceptor DTF-IC as the third component into a PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]IT-M system.82 The addition of DTF-IC broadened the absorption of the active layer in a range of 730–800 nm. Moreover, the third component provided a cascaded energy between the host donor and acceptor, which was favorable for efficient charge transfer between the acceptors and facilitating exciton dissociation. Consequently, the PCE of the ternary device increased from 10.90% to 12.14% when the content of DTF-IC was 10%.

In 2019, Xie et al. fabricated ternary OSCs consisting of one polymer donor PBDB-TF and two non-fullerene acceptors IT-4F and O6T-4F.65 The third component O6T-4F exhibited strong absorption in the near infrared region and could be used to extend the absorption spectrum. The incorporation of 15% O6T-4F into the PBDB-TF[thin space (1/6-em)]:[thin space (1/6-em)]IT-4F blend optimized the interpenetrating D/A morphology and improved the exciton dissociation by energy transfer from IT-4F to O6T-4F. The ternary OSC showed a Voc of 0.83 V, a Jsc of 21.60 mA cm−2 and an FF of 74.6%, leading to a remarkable PCE of 13.37%. Chen et al. selected a non-fullerene acceptor IDIC and its fluorinated acceptor ID4F to combine with a polymer donor PBDBT-SF in the construction of a ternary OSC.84 The structural similarity between IDIC and ID4F enabled the formation of the acceptor alloy, which was beneficial for charge transport. At the optimized ratio of PBDBT-SF[thin space (1/6-em)]:[thin space (1/6-em)]IDIC[thin space (1/6-em)]:[thin space (1/6-em)]IT4F(1[thin space (1/6-em)]:[thin space (1/6-em)]0.6[thin space (1/6-em)]:[thin space (1/6-em)]0.4), the ternary devices achieved a high PCE of 11.52%, a Voc of 0.917 V, a Jsc of 16.89 mA cm−2 and an FF of 74.35%. The ternary OSCs also showed good thickness tolerance with a PCE > 10% from 85 to 250 nm and better stability than the binary OSCs. Zhang et al. reported a series of high-performance ternary OSCs based on one polymer donor PBDB-T and two non-fullerene acceptors Y16 and MeIC1.85 The compatible Y16 and MeIC1 preferred to form alloyed acceptors with mixed energy levels, which was beneficial for the formation of electron transport pathways. The phase separation of the active layers can also be optimized by incorporating 15% MeIC1 as a morphology regulator. As a result, the champion device showed a high PCE of 14.11%, a Jsc of 22.76 mA cm−2 and an FF of 68.22%. A fluorine-substituted small-molecule acceptor ITIF was synthesized and used in a ternary OSC based on PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]ITIF[thin space (1/6-em)]:[thin space (1/6-em)]ITIC blends by Chen et al.83 Due to the fluorinated and nonplanar structure on the asymmetric terminal group, ITIF showed a higher LUMO level, and a lower HOMO level than ITIC, leading to enhanced absorption at 500–600 nm. As shown in Fig. 12a, the asymmetric terminal groups led to a nonplanar structure of ITIF, which could reduce its aggregation. The X-ray diffraction (XRD) pattern of the blends (Fig. 12b) demonstrated the co-crystal behavior of ITIC and ITIF, owing to their structural similarity. Moreover, the decreased melting point of ITIF can be observed in the differential scanning calorimetry (DSC) curve (Fig. 12c and d), indicating that ITIF could reduce the aggregation of ITIC. As a result, the ternary device loading 5% of ITIF exhibited a PCE of 10.53%, a Voc of 0.865 V, a Jsc of 17.31 mA cm−2 and an FF of 70.3%. Wang et al. introduced a narrow bandgap acceptor 4TIC as the third component into a PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]IEICO-4F binary blend.86 The efficient energy transfer from PBDB-T to the alloyed acceptor enhanced the exciton utilization of the ternary active layer. 4TIC can act as a regulator to adjust the molecular arrangement of PBDB-T, which suppressed charge carrier recombination and optimized charge carrier transport. The PCE of ternary OSCs increased from 10.02% to 11.63% when 40% IEICO-4F was replaced with 4TIC, together with an enhanced Voc of 0.82 V, a Jsc of 21.52 mA cm−2 and an FF of 65.9%. Yang et al. reported a ternary OSC system using a wide bandgap polymer donor (PBTA-PS) and two non-fullerene small-molecule acceptors (IDIC-C4Ph and 6TIC).87 When 30% of 6TIC was incorporated into a PBTA-PS[thin space (1/6-em)]:[thin space (1/6-em)]IDIC-C4Ph blend, the ternary OSC showed an outstanding PCE of 14.24% with a high Voc of 0.84 V, an improved Jsc of 22.33 mA cm−2, and an excellent FF of 75.94%. The superior performance was mainly due to the complementary absorption of the active layer, suppressed energy loss in the charge generation process and efficient charge transfer between acceptors. Yang and coworkers used two structurally similar acceptors (ITIC and IT-4F) to construct ternary OSCs combining with a polymer donor PBTA-PS.88 They found that the crystalline IT-4F could function as a morphology regulator to enhance the molecular packing. Moreover, efficient energy transfer from ITIC to IT-4F was observed, which could promote the exciton dissociation process. After incorporating 20% IT-4F into a PBTA-PS[thin space (1/6-em)]:[thin space (1/6-em)]ITIC system, the ternary device exhibited an outstanding PCE of 13.27% with a Jsc of 19.60 mA cm−2 and an FF of 74.45%. Li and coworkers designed and synthesized two small-molecule acceptors C8-SF and Y-MODF for the construction of ternary OSCs.89 The addition of 25% Y-MODF into a PM6[thin space (1/6-em)]:[thin space (1/6-em)]C8-SF system could suppress monomolecular recombination and improve the utilization of photons in the 450–600 nm range. The optimal device achieved a high PCE of 13.39%, with an enhanced Voc of 0.845 V, a Jsc of 20.88 mA cm−2 and an FF of 75.87%. Li and coworkers fabricated ternary OSCs by introducing a highly crystalline IDIC into a low crystalline blend of J61 and BT-IC.90 The introduction of 20% IDIC can enhance the crystallinity of J61 in the blend film, yielding higher hole mobility and achieving higher Jsc and FF than those of binary OSCs based on a J61[thin space (1/6-em)]:[thin space (1/6-em)]BT-IC blend. At a J61[thin space (1/6-em)]:[thin space (1/6-em)]IDIC[thin space (1/6-em)]:[thin space (1/6-em)]BT-IC ratio of 0.9[thin space (1/6-em)]:[thin space (1/6-em)]0.2[thin space (1/6-em)]:[thin space (1/6-em)]0.8, the optimized device achieved a PCE of 10.76%, a Voc of 0.88 V, a Jsc of 17.91 mA cm−2 and an FF of 68.37%. Zhang and coworkers incorporated a medium bandgap small-molecule acceptor BDTThIT-4F as the third component to blend with a PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]COi8DFIC system.91 The introduction of BDTThIT-4F could adjust the molecular arrangement of COi8DFIC to enhance photon harvesting in the long wavelength range, leading to improved EQE values from 550 to 1000 nm. By the incorporation of 20% of BDTThIT-4F, the PCE of the ternary devices reached 13.08%, which was higher than 11.47% for COi8DFIC-based binary OSCs and 9.25% for BDTThIT-4F-based binary OSCs. Zhan et al. fabricated ternary OSCs by combining F8IC and IDT-2BR acceptors to improve device performance.92 The higher LUMO energy level of IDT-2BR preferred to form an electron transport channel, which was beneficial for exciton separation and reducing energy loss in the charge transfer process. The ternary device with 20% IDT-2BR content in the active layer yielded a PCE of 12.1%, a Voc of 0.687 V, a Jsc of 25.1 mA cm−2 and an FF of 69.9%.


image file: d0nr07788g-f12.tif
Fig. 12 (a) Molecular geometry of ITIF and ITIC calculations by density functional theory (DFT) at the B3LYP/6-31G(d,p) level. (b) X-ray diffraction (XRD) patterns of ITIC, ITIF and the mixed acceptors with 5% ITIF in solid powders; DSC thermograms of ITIF, ITIC and the mixed acceptors with 5%: (c) heat-only thermograms and (d) cool-only thermograms. Reproduced with permission.83 Copyright 2019, Wiley.

Tang and coworkers employed a small-molecule acceptor N7IT as the third component added into the PM6[thin space (1/6-em)]:[thin space (1/6-em)]IT-4F blend to fabricate ternary OSCs.93 The incorporation of N7IT into IT-4F decreased the PL intensity of IT-4F, and increased the PL intensity in the low energy region. This obviously indicated the energy transfer from IT-4F to N7IT. Compared with PM6[thin space (1/6-em)]:[thin space (1/6-em)]IT-4F, the crystal coherence length of the PM6[thin space (1/6-em)]:[thin space (1/6-em)]IT-4F[thin space (1/6-em)]:[thin space (1/6-em)]N7IT blend was reduced, resulting in a higher electron mobility of 4.90 × 10−4 cm2 V−1 s−1. AFM images of the ternary blend showed clear nanofiber network structures with an RMS of 1.66 nm, which demonstrated good morphology compatibility between the acceptors of IT-4F and N7IT. As a result, the optimal ternary device (PM6[thin space (1/6-em)]:[thin space (1/6-em)]IT-4F[thin space (1/6-em)]:[thin space (1/6-em)]N7IT = 1[thin space (1/6-em)]:[thin space (1/6-em)]0.7[thin space (1/6-em)]:[thin space (1/6-em)]0.3) gave an outstanding PCE of 15.02%, which was higher than those of the binary devices based on PM6[thin space (1/6-em)]:[thin space (1/6-em)]IT-4F (13.62%) and PM6[thin space (1/6-em)]:[thin space (1/6-em)]N7IT (13.91%).

In conclusion, when a non-fullerene small-molecule acceptor is introduced as the third component into host binary devices, the resulting ternary devices often exhibit enhanced photovoltaic performances. Firstly, compared with the host acceptor, a slightly higher LUMO energy level of the guest acceptor is beneficial for achieving a high Voc. Secondly, a narrow bandgap non-fullerene acceptor as the NIR absorber usually complements the absorption spectra of binary active layers, which may contribute to an increased Jsc. Thirdly, a small amount of the third component often acts as the morphology regulator to optimize the phase separation of ternary blend films, resulting in enhanced FFs of ternary OSCs.

(c) Polymer donor/small-molecule donor/fullerene acceptor OSCs. Polymer donor/small-molecule donor/fullerene acceptor ternary systems have been investigated in recent years. The introduction of a small-molecule donor into a polymer/fullerene binary blend could tune the molecular orientation, decrease π–π stacking distance and increase coherence length. The structures of donors for polymer donor/small-molecule donor/fullerene acceptor ternary OSCs are shown in Fig. 13, and the photovoltaic parameters are summarized in Table 3.
image file: d0nr07788g-f13.tif
Fig. 13 The structures of donors for polymer donor/small-molecule donor/fullerene acceptor ternary OSCs.
Table 3 Photovoltaic parameters of ternary OSCs with polymer donor/small-molecule donor/fullerene acceptor blends
D1 D2 (third component) A Weight ratio (D1[thin space (1/6-em)]:[thin space (1/6-em)]D2[thin space (1/6-em)]:[thin space (1/6-em)]A) V oc (V) J sc (mA cm−2) FF (%) PCE (%) Working mechanism Ref.
PTB7-Th p-DTS(FBTTH2)2 PC71BM 0.85[thin space (1/6-em)]:[thin space (1/6-em)]0.15[thin space (1/6-em)]:[thin space (1/6-em)]1.1 0.755 18.44 75.27 10.50 94
PTB7-Th DR3TSBDT PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]0.25[thin space (1/6-em)]:[thin space (1/6-em)]1.5 0.772 23.31 70.44 12.10 Energy transfer 95
PTB7-Th DR3TBDTT PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]0.1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 0.790 18.33 63.89 9.25 96
PTB7-Th BTR PC71BM 0.9[thin space (1/6-em)]:[thin space (1/6-em)]0.1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 0.78 19.23 72.21 10.83 Energy transfer 97
PTB7-Th BTR PC71BM 0.75[thin space (1/6-em)]:[thin space (1/6-em)]0.25[thin space (1/6-em)]:[thin space (1/6-em)]1.1 0.751 21.4 70.0 11.40 Energy transfer 98
PBDTTPD-HT BDT-3T-CNCOO PC71BM 0.6[thin space (1/6-em)]:[thin space (1/6-em)]0.4[thin space (1/6-em)]:[thin space (1/6-em)]1 0.969 12.17 71.23 8.40 99
PBDTTT-C-T n-BDTT-3T-CNCOO ICBA 0.75[thin space (1/6-em)]:[thin space (1/6-em)]0.25[thin space (1/6-em)]:[thin space (1/6-em)]1 0.98 11.43 49.4 5.51 100
PTB7-Th PDT2FBT-ID PC71BM 0.8[thin space (1/6-em)]:[thin space (1/6-em)]0.2[thin space (1/6-em)]:[thin space (1/6-em)]1.1 0.770 18.92 76.14 11.10 Energy transfer 101
PBDB-T tBTI-IDT PC61BM 1[thin space (1/6-em)]:[thin space (1/6-em)]0.2[thin space (1/6-em)]:[thin space (1/6-em)]1 0.86 13.53 71.8 8.40 Energy transfer 43
Si-PCPDTBT SMPV1 PC71BM 0.96[thin space (1/6-em)]:[thin space (1/6-em)]0.04[thin space (1/6-em)]:[thin space (1/6-em)]1.5 0.56 18.22 63.1 6.44 Energy transfer 102


In 2015, Wei et al. reported a high-performance ternary OSC by introducing a small molecule 7,7-(4,4-bis(2-ethylhexyl)-4H-silolo[3,2-b:4,5-b′]dithiophene-2,6-diyl)bis(6-fluoro-4-(5′-hexyl-[2,2′-bithiophene]-5-yl)benzo[c][1,2,5]thiadiazole) (p-DTS(FBTTH2)2) into a PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM system.94 The results of resonant soft X-ray scattering (R-SoXS) revealed that the position of the scattering peaks at q = 0.2 nm−1 were unchanged when 15% of p-DTS(FBTTH2)2 was incorporated into the PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM blend film, indicating that p-DTS(FBTTH2)2 did not influence bi-continuous phase separation in ternary systems. GIWAXS measurements demonstrated that a face-on π–π stacking peak became more obvious in the 2D images through incorporating p-DTS(FBTTH2)2 into PTB7-Th, indicating the enhanced face-on orientation in the ternary blend. Consequently, the hole mobilities were enhanced and the charge recombination was decreased. With the incorporation of 15% of p-DTS(FBTH2)2 in the ternary blend, the Jsc increased from 17.53 to 18.44 mA cm−2, the FF increased from 65.26% to 75.27%, and the PCE increased from 9.2% to 10.5%. Yang et al. incorporated a benzo[1,2-b;4,5-b′]dithiophene-based small molecule DR3TSBDT into a PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM blend to improve PCE in ternary devices.95 AFM measurements revealed that incorporating up to 30% of DR3TSBDT into the host blend could help achieve uniform nano-scale phase separation. In addition, GIWAXS measurements indicated that the addition of DR3TSBDT resulted in the formation of (010) π–π stacking peaks in both in-plane and out-of-plane directions, demonstrating the formation of mixed face-on and edge-on orientations. At a low DR3TSBDT loading of 25% into the host system, the ternary device exhibited a Jsc of 23.31 mA cm−2, a Voc of 0.772 V, an FF of 70.44%, and a high PCE of 12.10%. Yang's group also incorporated DR3TBDTT into a PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM blend to fabricate ternary devices.96 The results of AFM indicated that a highly ordered molecular compatibility could be achieved, and the addition of DR3TBDTT improved the phase separation of the active layer. Therefore, the ordered nano-morphology improved the hole mobility from 5.385 × 10−5 cm2 V−1 s−1 to 1.221 × 10−4 cm2 V−1 s−1. With 10% DR3TBDTT in the ternary system, the PCE of ternary devices increased from 7.53% to 9.25%.

Zhang et al. introduced a liquid-crystalline small-molecule donor BTR into a PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM system as the third component.97 With the incorporation of 10% BTR in the donors, the correlation length of PTB7-Th crystallites increased from 11.23 to 15.22 nm, indicating that the crystallinity of PTB7-Th was enhanced. In addition to enhancing the photon harvesting of the active layers, Zhang et al. found that BTR could also act as a nucleating agent to adjust the PTB7-Th molecular arrangement and optimize the phase separation. Incorporating 10% BTR in the donors, the PCE of ternary devices increased from 10.08% to 10.83%, resulting from the enhanced Jsc of 19.23 mA cm−2 and FF of 72.21%. Cao's group also reported ternary OSCs based on the same ternary blends.98 It was found that the addition of BTR into the PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM system could decrease the π–π stacking distance, enlarge the coherence length, and enhance the domain purity of the blend film, which were beneficial for efficient charge transport and reduced bimolecular recombination. Consequently, the hole mobility was significantly increased up to 6.33 × 10−3 cm2 V−1 s−1 with 25% BTR content. When the active layer thickness was 250 nm, the ternary device exhibited a remarkable PCE of 11.40%, a Voc of 0.751 V, a Jsc of 21.4 mA cm−2 and an FF of 70.0%.

In 2015, Wei et al. reported a small molecule BDT-3T-CNCOO as an additional donor component to blend with a PBDTTPD-HT[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM binary system.99 In the ternary system, the diffraction peaks were gradually strengthened with the increase of BDT-3T-CNCOO weight ratios, indicating that the small molecule BDT-3T-CNCOO could promote the crystallization and phase separation of donor materials. Therefore, the optimized domain size enabled efficient charge transport through favorable nano-structures. With the increase in small-molecule ratios, hole mobility gradually became more balanced, leading to reduced bimolecular recombination. As a result, when 40% BDT-3T-CNCOO was introduced to the binary blend, the PCE of the ternary device increased from 6.9% to 8.4%. In 2017, Wei's group incorporated n-BDT-3T-CNCOO into a PBDTTT-C-T[thin space (1/6-em)]:[thin space (1/6-em)]ICBA system.100 Compared with a PC71BM-based ternary system, the Voc of the ternary device significantly increased from 0.77 to 0.98 V due to the higher LUMO energy level of ICBA. The results of GIWAXS revealed that the scattering intensity gradually increased with the increase in the n-BDT-3T-CNCOO weight ratio, indicating that the addition of the small molecule promoted crystallization of the donors in the films. Such an addition also enhanced the aggregation of ICBA, resulting in enhanced charge carrier dissociation and improved charge transport. The PCE was enhanced from 5.01% in binary OSCs to 5.51% in ternary OSCs.

Another small molecule PDT2FBT-ID was applied in a PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]PDT2FBT-ID[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM ternary system by Wei et al.101 PDT2FBT-ID exhibited a broad and strong emission range from 650 to 850 nm covering the absorption range of PTB7-Th, which indicated the potential energy transfer from PDT2FBT-ID to PTB7-Th. There were no obvious PDT2FBT-ID crystalline peaks in GIWAXS results, which demonstrated that the small amount of PDT2FBT-ID was likely to disperse in the PTB7-Th amorphous range, and induce the crystallinity of PTB7-Th in the face-on direction. In addition, the alloy-like structure and energy transfer mechanism coexisted in the ternary system. As a result, the FF improved from 66.72% to 76.14% and the Jsc was improved from 17.90 to 18.92 mA cm−2. A maximum PCE of 11.1% was obtained in the ternary device. Welch et al. incorporated a benzothioxanthene-based small-molecule donor tBTI-IDT as the third component to blend with a PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]PC61BM system.43 The EQE spectra of the ternary blend film indicated that the addition of tBTI-IDT increased optical absorption from 400 to 450 nm. The Voc values were slightly enhanced from 0.80 to 0.95 V with increasing concentration of tBTI-IDT, indicating that the ternary system was likely to work as a series connection of OSCs. Addition of 10% of the third component did not significantly alter the film morphology. The optimal device achieved a PCE of 8.40%, a Voc of 0.86 V, a Jsc of 13.53 mA cm−2, and an FF of 71.8%. Tang et al. reported a ternary system with a low bandgap polymer Si-PCPDTBT as the host donor and a small molecule SMPV1 as an additional donor.102 SMPV1 exhibited strong absorption in the range of 450–650 nm, which was complementary to the absorption of a Si-PCPDTBT[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM blend film. TEM measurements showed that the nano-fibrillar networks of the active layers became more and more obvious along with the increase of SMPV1 weight ratios up to 4%. The results indicated that the addition of SMPV1 could adjust the molecular arrangement of Si-PCPDTBT to form the more efficient bi-continuous transport channels in the ternary active layers. As a result, the champion PCE of ternary devices was increased from 5.73% to 6.44% by adding 4% of SMPV1 in donors with solvent vapor annealing treatment.

In general, the conjugated small molecules can exhibit broad absorption spectra in ternary blend films, resulting in improved light harvesting ability. Owing to the high crystallinity and hole mobility of the conjugated small molecules, the incorporation of small molecules could enhance the crystallinity of the polymer donor and hence suppress bimolecular recombination. As a result, the better morphology of ternary OSCs could be achieved through the synergistic effect of small molecules and polymers.

(d) Polymer donor/small-molecule donor/non-fullerene acceptor OSCs. To overcome the weak absorption of fullerene derivatives, polymer donor/small-molecule donor/non-fullerene acceptor OSCs were investigated. The structures of donors and acceptors for polymer donor/small-molecule donor/non-fullerene acceptor ternary OSCs are shown in Fig. 14, and the photovoltaic parameters are summarized in Table 4.
image file: d0nr07788g-f14.tif
Fig. 14 The structures of donors and acceptors for polymer donor/small-molecule donor/non-fullerene acceptor OSCs.
Table 4 Photovoltaic parameters of ternary OSCs with polymer donor/small-molecule donor/fullerene acceptor blends
D1 D2 (third component) A Weight ratio (D1[thin space (1/6-em)]:[thin space (1/6-em)]D2[thin space (1/6-em)]:[thin space (1/6-em)]A) V oc (V) J sc (mA cm−2) FF (%) PCE (%) Working mechanism Ref.
PTB7-Th TR FOIC 1[thin space (1/6-em)]:[thin space (1/6-em)]0.25[thin space (1/6-em)]:[thin space (1/6-em)]1.5 0.734 25.1 70.9 13.1 Energy transfer 23
PTB7-Th DIBC IEICO-4F 1[thin space (1/6-em)]:[thin space (1/6-em)]0.1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 0.70 26.80 71.36 13.53 Energy transfer 103


In 2019, Zhan and coworkers reported the first example of ternary OSCs based on a polymer donor/small-molecule donor/non-fullerene acceptor, in which a mid-bandgap small-molecule donor TR was introduced as a third component into a PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]FOIC system.23 The high crystallinity of TR could enhance the ordered molecular packing and hole mobility of PTB7-Th by regulating the morphology of blend films. The optimized PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]TR[thin space (1/6-em)]:[thin space (1/6-em)]FOIC blend exhibited an increased hole mobility of 6.8 × 10−4 cm2 V−1 s−1, which was higher than those of PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]FOIC (4.0 × 10−4 cm2 V−1 s−1) and TR[thin space (1/6-em)]:[thin space (1/6-em)]FOIC (3.4 × 10−4 cm2 V−1 s−1) binary blend films. The results of AFM measurements showed that the RMS roughness of the ternary blend was reduced from 3.52 to 1.41 nm, which was beneficial for efficient charge transport and suppressing bimolecular recombination. GIWAXS measurements demonstrated the good miscibility between PTB7-Th and TR, due to the lack of the (010) peak of PTB7-Th in the ternary blend film. When the weight ratio of TR was 25% in the ternary system, the best-performing device yielded a remarkable PCE of 13.1%, with a Voc of 0.734 V, a Jsc of 25.1 mA cm−2 and an FF of 70.9%. Tao and coworkers introduced a small molecule DIBC as the third component into a PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]IEICO-4F system.103 The DIBC exhibited strong absorption in the range of 400–600 nm, which was complementary to the absorption of PTB7-Th. When 10% of DIBC was doped in the PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]IEICO-4F blend, the absorption spectrum of the ternary film showed an obvious enhancement in the range of 300–900 nm, which was desirable for harvesting more photons. The PL spectrum of DIBC overlapped with the absorption spectrum of PTB7-Th, suggesting efficient energy transfer between DIBC and PTB7-Th. More importantly, the formation of an intermolecular hydrogen bond between DIBC and IEICO-4F was confirmed from the Fourier transform infrared (FT-IR) spectra. This hydrogen-bond interaction can enlarge the π–π stacking distance and suppress oversize aggregation of PTB7-Th, leading to an improved morphology of the ternary film. As a result, the PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]IEICO-4F[thin space (1/6-em)]:[thin space (1/6-em)]DIBC-based ternary device showed a high PCE of 13.53% with excellent shelf-life. After 90 days of storage, the ternary device still retained 85.02% of the original PCE, which was almost two times higher than that for the PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]IEICO-4F-based binary device.

The use of crystalline small molecules as the third component could enhance the crystallinity of the polymer donor and optimize phase separation. The high mobility of small molecules could also provide another channel for efficient charge transport and thereby reduce bimolecular recombination. In particular, the formation of hydrogen bonds between the small molecule and non-fullerene acceptor could suppress the excessive aggregation of the polymer donor, which improved the morphology of the ternary blend.

2.2. Polymer/polymer/small-molecule OSCs

Due to the narrow absorption spectra of conjugated polymers, the use of two donor polymers can broaden the light absorption of the active layer and facilitate charge dissociation and transport via the cascade energy structure and optimized morphology.104 Thus, much research work has been focused on the utilization of polymer donors as guest donors in fullerene-free ternary OSCs aiming at improving the performance of the binary blend system. The structures of donors and acceptors for polymer/polymer/small-molecule ternary OSCs are shown in Fig. 15, and the photovoltaic parameters are summarized in Table 5.
image file: d0nr07788g-f15.tif
Fig. 15 The structures of donors and acceptors for polymer/polymer/small-molecule ternary OSCs.
Table 5 Photovoltaic parameters of ternary OSCs with polymer/polymer/small-molecule blends
D1 D2 (third component) A Weight ratio (D1[thin space (1/6-em)]:[thin space (1/6-em)]D2[thin space (1/6-em)]:[thin space (1/6-em)]A) V oc (V) J sc (mA cm−2) FF (%) PCE (%) Working mechanism Ref.
PF2 J71 Y6 0.7[thin space (1/6-em)]:[thin space (1/6-em)]0.3[thin space (1/6-em)]:[thin space (1/6-em)]1.5 0.75 24.97 64.7 12.12 Energy transfer 105
PM6 J71 Br-ITIC 0.8[thin space (1/6-em)]:[thin space (1/6-em)]0.2[thin space (1/6-em)]:[thin space (1/6-em)]1 0.930 19.39 78.4 14.13 Energy transfer 24
FTAZ PBDB-T ITM 0.8[thin space (1/6-em)]:[thin space (1/6-em)]0.2[thin space (1/6-em)]:[thin space (1/6-em)]1 0.95 18.1 73.6 13.2 106
PTB7-Th PBDTm-T1 FOIC 0.8[thin space (1/6-em)]:[thin space (1/6-em)]0.2[thin space (1/6-em)]:[thin space (1/6-em)]1.5 0.762 24.0 73.5 13.4 Energy transfer 107
PTB7-Th PBDB-T O-IDTBR 0.7[thin space (1/6-em)]:[thin space (1/6-em)]0.3[thin space (1/6-em)]:[thin space (1/6-em)]1.5 1.02 17.07 66.52 11.58 108
PBDB-T PTB7-Th FOIC 0.5[thin space (1/6-em)]:[thin space (1/6-em)]0.5[thin space (1/6-em)]:[thin space (1/6-em)]1 0.73 24.61 65.51 12.02 Energy transfer 109
PBDB-T PTB7-Th IEICO-4F 0.8[thin space (1/6-em)]:[thin space (1/6-em)]0.2[thin space (1/6-em)]:[thin space (1/6-em)]1 0.74 24.14 65.03 11.62 Energy transfer 110
J51 PTB7-Th ITIC 0.8[thin space (1/6-em)]:[thin space (1/6-em)]0.2[thin space (1/6-em)]:[thin space (1/6-em)]1 0.81 17.75 67.82 9.70 Energy transfer 111
PBDB-T PTB7-Th SFBRCN 0.3[thin space (1/6-em)]:[thin space (1/6-em)]0.7[thin space (1/6-em)]:[thin space (1/6-em)]1 0.93 17.86 73.9 12.27 Energy transfer 112
PDCBT PTB7-Th PC71BM 0.5[thin space (1/6-em)]:[thin space (1/6-em)]0.5[thin space (1/6-em)]:[thin space (1/6-em)]1.5 0.78 17.53 74 10.12 Energy transfer 113
PTB7-Th PBT1-S PC71BM 0.9[thin space (1/6-em)]:[thin space (1/6-em)]0.1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 0.82 16.8 72.2 10.3 Energy transfer 114
PTB7-Th PffBT4T-2OD PC71BM 0.85[thin space (1/6-em)]:[thin space (1/6-em)]0.15[thin space (1/6-em)]:[thin space (1/6-em)]1.2 0.776 19.02 72.62 10.72 Energy transfer 115
PTB7-Th PBDTTS-FTAZ PC71BM 0.8[thin space (1/6-em)]:[thin space (1/6-em)]0.2[thin space (1/6-em)]:[thin space (1/6-em)]1.5 0.816 16.4 67.3 9.2 Energy transfer 116
PPDT2FBT PPDT2CNBT PC71BM 0.9[thin space (1/6-em)]:[thin space (1/6-em)]0.1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 0.76 18.7 67 9.46 Energy transfer 117


Zhang's group designed ternary OSCs by introducing a polymer donor J71 into a blend of PF2 and Y6.105 They found that the Elosss of OSCs gradually decreased from 0.68 to 0.53 eV with increasing J71 content in donors, which may be an important factor for PCE improvement of ternary OSCs. The photoluminescence (PL) spectra (Fig. 16a) of the ternary blend showed that the PF2 emission intensity increased monotonically with the increase of J71 content, indicating the existence of energy transfer from J71 to PF2. The energy transfer can be further confirmed from the decreased 635 nm emission lifetime and increased 745 nm emission lifetime in blend films, as shown in time-resolved photoluminescence (TRPL) spectra (Fig. 16b). When the J71 content was up to 30% in donors, the Jsc increased from 23.17 to 24.97 mA cm−2 and the FF increased from 62.35% to 64.70%. As a result, a PCE of 12.12% was obtained. The same group also introduced J71 as the second donor into the PM6[thin space (1/6-em)]:[thin space (1/6-em)]J71 binary blend to fabricate ternary OSCs.24 PM6 and J71 may form an alloyed donor due to the good compatibility, which was beneficial for optimizing photon harvesting and phase separation of ternary active layers. When using TEM measurements to characterize the morphology of blend films, the fibrillar structures were found to be well maintained in ternary blend films with 20% J71 content, suggesting an ideal phase separation for sufficient charge transport. It was found that some excitons on J71 may transfer their energy onto PM6 and then are dissociated into free carriers at PM6/Br-ITIC interfaces. Consequently, a PCE of 14.13% was achieved in the optimized ternary OSCs with improved Jsc (19.39 mA cm−2) and FF (78.4%).


image file: d0nr07788g-f16.tif
Fig. 16 (a) PL spectra of neat PF2, J71 and blend PF2[thin space (1/6-em)]:[thin space (1/6-em)]J71 films under 540 nm light excitation. (b) TRPL spectra of neat PF2, J71 and blend PF2[thin space (1/6-em)]:[thin space (1/6-em)]J71 films obtained by probing 635 nm and 745 nm light emission. Reproduced with permission.105 Copyright 2019, Royal Society of Chemistry.

To study the morphological and mechanical stability of ternary OSCs, Ade et al. reported a highly efficient ternary OSC based on a non-fullerene acceptor IT-M and two polymer donors (FTAZ and PBDB-T).106 Due to the rigid polymer chains of PBDB-T, the crystallization behavior of ITM was significantly suppressed in the FTAZ[thin space (1/6-em)]:[thin space (1/6-em)]PBDBT[thin space (1/6-em)]:[thin space (1/6-em)]IT-M (0.8[thin space (1/6-em)]:[thin space (1/6-em)]0.2[thin space (1/6-em)]:[thin space (1/6-em)]1) blend film even after thermal annealing at 180 °C. Consequently, the corresponding ternary device still retained 90% (Fig. 17) of the initial efficiency (13.2%) after a storage time of ∼1000 h in a glovebox. Another ternary OSC was fabricated by employing PTB7-Th as the donor, FOIC as the acceptor, and a PBDTm-T1 polymer donor as the third component.107 Incorporating 20% PBDTm-T1 in the host PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]FOIC blend led to more balanced charge mobilities (μh = 8.9 × 10−4 cm2 V−1 s−1, μe = 6.6 × 10−4 cm2 V−1 s−1). GIWAXS patterns showed that the incorporation of PBDTm-T1 had negligible influence on the π–π stacking and face-on orientation of the blends. As a result, an improved PCE of 13.8% with a decreased non-radiative recombination loss of 0.271 eV in ternary OSCs was achieved. Chen et al. designed a ternary blend based on two polymer donors of PTB7-Th and PBDB-T with comparable HOMO energy levels and a non-fullerene acceptor O-IDTBR, which exhibited a high Voc of 1.02 V.108 With an optimized ratio of 0.7[thin space (1/6-em)]:[thin space (1/6-em)]0.3[thin space (1/6-em)]:[thin space (1/6-em)]1.5 for a PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]O-IDTBR blend, the ternary device displayed a PCE of 11.58%. The ternary device retained good PCEs of 9.37% after 168 h thermal annealing at 85 °C, much higher than 6.27% and 5.33% for PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]O-IDTBR and PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]O-IDTBR binary devices, respectively. In addition, the authors also studied the photovoltaic performances of larger area devices of 0.37, 0.57, and 0.91 cm2. The corresponding devices exhibited PCEs of 10.12%, 9.51%, and 9.01%, respectively.


image file: d0nr07788g-f17.tif
Fig. 17 Normalized PCE of devices based on FTAZ[thin space (1/6-em)]:[thin space (1/6-em)]IT-M (1[thin space (1/6-em)]:[thin space (1/6-em)]1) and FTAZ[thin space (1/6-em)]:[thin space (1/6-em)]PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]IT-M (0.8[thin space (1/6-em)]:[thin space (1/6-em)]0.2[thin space (1/6-em)]:[thin space (1/6-em)]1) (annealed at 150 °C for 10 min) as a function of storage time in nitrogen under darkness. Reproduced with permission.106 Copyright 2019, Wiley.

Ma et al. fabricated ternary OSCs based on PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]FOIC blends by blade coating in ambient environment.109 Upon the addition of PTB7-Th into PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]FOIC blends, the crystallinity of FOIC decreased owing to the better miscibility of PTB7-Th and FOIC than that of PBDB-T and FOIC, which exhibited more efficient charge transport. Consequently, the optimal ternary device based on PBDBT[thin space (1/6-em)]:[thin space (1/6-em)]PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]FOIC blends (0.5[thin space (1/6-em)]:[thin space (1/6-em)]0.5[thin space (1/6-em)]:[thin space (1/6-em)]1) provided the best PCE of 12.02% with a high FF of 65.78%. By incorporating 20% PTB7-Th into a PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]IEICO-4F binary blend, Tang et al. fabricated efficient ternary OSCs with PCEs as large as 11.62%.110 The combination of three materials can cover a large photon harvesting range from 300 to 1000 nm. The energy transfer between PBDB-T and PTB7-Th may enhance exciton utilization efficiency. Thus, the two factors resulted in a high Jsc of 24.14 mA cm−2. Due to the similar HOMO energy levels of the two donors, the Voc of 0.74 V for ternary OSCs can remain unchanged. Moreover, PTB7-Th can also act as a regulator to adjust the PBDB-T molecular arrangement for more efficient charge transport, and a high FF of 65.03% was obtained. Li et al. investigated ternary OSCs with PTB7-Th as the second donor in J51[thin space (1/6-em)]:[thin space (1/6-em)]ITIC blends.111 When the blend ratio of J51[thin space (1/6-em)]:[thin space (1/6-em)]PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]ITIC was 0.8[thin space (1/6-em)]:[thin space (1/6-em)]0.2[thin space (1/6-em)]:[thin space (1/6-em)]1, the ternary device achieved a PCE of 9.7% together with a Jsc of 17.75 mA cm−2. The enhanced absorption in the wavelength region of ∼650–750 nm, reduced phase separation and energy transfer between J51 and PTB7-Th were the reasons for the improved Jsc and high efficiency. Peng et al. reported ternary OSCs based on PBDBT[thin space (1/6-em)]:[thin space (1/6-em)]PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]SFBRCN blends.112 Such ternary devices exhibited a broad composition tolerance over the whole D1[thin space (1/6-em)]:[thin space (1/6-em)]D2 weight ratios (from 1[thin space (1/6-em)]:[thin space (1/6-em)]0 to 0[thin space (1/6-em)]:[thin space (1/6-em)]1), due to the good miscibility and desirable nano-phase morphology of the two donor materials when blending with acceptors. To explore the device stability with different ternary compositions, the OSCs were treated thermally from room temperature to 140 °C. From the results, OSCs based on the ternary blend with a 0.5 ratio of PTB7-Th showed the best thermal stability (over 0.8 at 140 °C). At the optimized weight ratio of 0.3[thin space (1/6-em)]:[thin space (1/6-em)]0.7[thin space (1/6-em)]:[thin space (1/6-em)]0.8 for a PBDB-T[thin space (1/6-em)]:[thin space (1/6-em)]PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]SFBRCN ternary blend, the resulting devices exhibited an improved Voc of 0.93 V, a Jsc of 17.86 mA cm−2 and an FF of 73.9%, delivering a high PCE of 12.27%.

Ameri et al. introduced PTB7-Th as the third component into a PDCBT[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM system.113 The high intermolecular affinity between PDCBT and PTB7-Th endowed the two polymer donors with a particular blend microstructure, which notably improved the charge transport mechanism. The resulting ternary device yielded a PCE of 10.12% with a high FF of 74%. Sun et al. fabricated semitransparent ternary OSCs using a wide bandgap polymer donor PBT1-S as the third component.114 PTB7-Th and PBT1-S showed complementary absorption, which contributed to extending the absorption of active layers. Moreover, the addition of 10% PBT1-S did not significantly change its average visible transmittance. The champion ternary devices showed a high PCE of 10.3%. A ternary system based on PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]PffBT4T-2OD[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM was reported by Wang and coworkers.115 It was found that the energy transfer from PffBT4T-2OD to PTB7-Th enhanced light harvesting. After incorporating 15% PffBT4T-2OD, the blend film showed hierarchical phase separation within 25 nm and improved phase purity, which was beneficial for achieving sufficient exciton separation and charge transport. Therefore, an improved PCE of 10.72% and Jsc (19.02 mA cm−2) as well as FF (72.62%) were obtained in the ternary device. Fahlman et al. used two polymer donors PBDTTS-FTAZ and PTB7-Th to fabricate fullerene-based ternary OSCs.116 Due to the well-matched PL and absorption spectra of the two polymers, the energy transfer from PBDTTS-FTAZ to PTB7-Th was realized.

Moreover, adding PBDTTS-FTAZ into the ternary blend could promote the exciton dissociation of PC71BM and the donors. The ternary device with 20% PBDTTS-FTAZ exhibited a PCE of 9.2%, a Voc of 0.816 V, a Jsc of 16.4 mA cm−2 and an FF of 67.3%. Kim et al. reported high-performance ternary OSCs by adding PPDT2CNBT into a PPDT2FBT[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM binary blend.117 The fluorine atoms on the benzothiadiazole moiety of PPDT2CNBT were replaced with nitrile groups, which overcame the poor solubility in processing solvents. The additional donor extended the light absorption of the active layer in a range of 700–850 nm and improved the Voc because of its deep HOMO energy level. As a result, the best-performing devices showed a PCE of 9.46%, a Jsc of 18.7 mA cm−2, a Voc of 0.76 V, and an FF of 67%.

The use of two compatible polymer donors may suppress the crystallization behavior of acceptors, and the energy transfer between them will enhance the exciton utilization of the active layer. However, the introduction of the additional polymer donor usually leads to interrupted crystallization of the host donor. Consequently, controlling the interactions between two polymer donors should be further investigated.

2.3. All-polymer ternary OSCs

To overcome the limited sunlight absorption in most of the binary all-polymer solar cells, all-polymer ternary OSCs have been gradually developed in recent years. The introduction of the third component will lead to difficulty in controlling the morphology of ternary blend films. Therefore, careful selection of the ternary components is important for optimizing all the photovoltaic parameters. The structures of donors and acceptors for all-polymer ternary OSCs are shown in Fig. 18, and the photovoltaic parameters are summarized in Table 6.
image file: d0nr07788g-f18.tif
Fig. 18 The structures of donors and acceptors for all-polymer ternary OSCs.
Table 6 Photovoltaic parameters of ternary OSCs with all-polymer blends
D1 D2 (third component) A Weight ratio (D1[thin space (1/6-em)]:[thin space (1/6-em)]D2[thin space (1/6-em)]:[thin space (1/6-em)]A) V oc (V) J sc (mA cm−2) FF (%) PCE (%) Working mechanism Ref.
PTB7-Th PBDD-ff4T N2200 1.5[thin space (1/6-em)]:[thin space (1/6-em)]0.25[thin space (1/6-em)]:[thin space (1/6-em)]1 0.82 15.7 56 7.2 Charge transfer 118
PBDTTT-EF-T PCDTBT N2200 1[thin space (1/6-em)]:[thin space (1/6-em)]0.1[thin space (1/6-em)]:[thin space (1/6-em)]1 0.790 14.4 58.3 6.65 Energy transfer 119
PTB7-Th PBDTTS-FTAZ PNDI-T10 1[thin space (1/6-em)]:[thin space (1/6-em)]0.15[thin space (1/6-em)]:[thin space (1/6-em)]1 0.84 14.4 74 9.0 Energy transfer 120
PTzBI-Si PBTA-Si N2200 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 0.85 14.89 75.65 9.56 Energy transfer 121
PEG-2% PTB7-Th N2200 1.4[thin space (1/6-em)]:[thin space (1/6-em)]0.6[thin space (1/6-em)]:[thin space (1/6-em)]1 0.83 16.89 66.13 9.27 122
PBTA-BO PNTB N2200 1.7[thin space (1/6-em)]:[thin space (1/6-em)]0.3[thin space (1/6-em)]:[thin space (1/6-em)]1 0.84 15.77 74.98 10.09 Energy transfer 123
PTzBI PBTA-BO N2200 1.7[thin space (1/6-em)]:[thin space (1/6-em)]0.3[thin space (1/6-em)]:[thin space (1/6-em)]1 0.836 15.64 78.33 10.12 Energy transfer 25
J51 PTB7-Th N2200 1.4[thin space (1/6-em)]:[thin space (1/6-em)]0.6[thin space (1/6-em)]:[thin space (1/6-em)]1 0.81 17.17 66.36 9.6 124
PTB7-Th PCDTBT N2200 1[thin space (1/6-em)]:[thin space (1/6-em)]0.15[thin space (1/6-em)]:[thin space (1/6-em)]1 0.78 11.44 56 5.11 125
PTB7-Th PF12TBT P(NDI2OD-T2) 1[thin space (1/6-em)]:[thin space (1/6-em)]0.15[thin space (1/6-em)]:[thin space (1/6-em)]1 0.81 12.28 61 6.07 Energy transfer 126
PTB7-Th PBClT NDP-V-C7 0.85[thin space (1/6-em)]:[thin space (1/6-em)]0.15[thin space (1/6-em)]:[thin space (1/6-em)]1 0.78 16.77 68.07 9.03 127


Introducing a second polymer donor with appropriate absorption, energy levels and crystallinity into a binary all-polymer system is one of the promising strategies for achieving high-performance all-polymer ternary OSCs. In 2016, Li and coworkers incorporated a moderate bandgap polythiophene derivative PBDD-ff4T into a low bandgap PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]N2200 blend to complement the absorption of the host blend.118 PBDD-ff4T also exhibited high crystallinity, which can promote charge transport and suppress geminate recombination in all-polymer ternary OSCs. The introduction of 10% PBDD-ff4T in the ternary blends significantly improved the μh from 5.47 × 10−4 to 1.23 × 10−3 cm2 V−1 s−1, increasing the PCE from 5.9% to 7.2%. Another example based on PBDTTT-EF-T[thin space (1/6-em)]:[thin space (1/6-em)]N2200 and a wide bandgap polymer donor PCDTBT was reported by Ito et al.119 When the PCDTBT loading amount in the ternary blend was 10%, the EQEs at visible wavelengths were successfully increased up to 65–70%. In addition, the excitons of PCDTBT can be directly transported to PBDTTT-EF-T and N2200 via resonant Förster energy transfer, contributing to efficient charge generation. Therefore, the ternary device achieved an improved Jsc of 14.4 mA cm−2 and a PCE of 6.65%. Wang and coworkers improved the photovoltaic performance of PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]PNDI-T10-based all-polymer solar cells by introducing a high bandgap polymer PBDTTS-FTAZ as the second donor.120 PBDTTS-FTAZ served as a sensitizer in the visible region of 450–650 nm, broadening the absorption of the ternary blend film. The PL spectrum of PBDTTS-FTAZ was found to overlap well with the absorption spectra of PTB7-Th and PNDI-T10, suggesting that a Förster energy transfer from PBDTTS-FTAZ to PTB7-Th and PNDI-T10 can occur at D/A interfaces and thereby improve charge generation in all-polymer ternary OSCs. These advantages enabled an outstanding PCE of 9.0% in the optimized 1[thin space (1/6-em)]:[thin space (1/6-em)]0.15[thin space (1/6-em)]:[thin space (1/6-em)]1 ternary device.

In 2018, Cao et al. performed a series of excellent research studies on all-polymer ternary OSCs. For example, they synthesized a wide bandgap polymer donor PBTA-Si to blend with a PTzBI-Si[thin space (1/6-em)]:[thin space (1/6-em)]N2200 system.121 PBTA-Si contained a difluorobenzotriazole building block with a siloxane-terminated side chain, which can be dissolved in green solvent 2-methyltetrahydrofuran (MeTHF) at room temperature. The good compatibility between PBTA-Si and PTzBI-Si could drive the molecular stacking of N2200 in the ternary blend, leading to reduced structure defects and enhanced charge transport. Therefore, the ternary blend with a ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 for PBTA-Si[thin space (1/6-em)]:[thin space (1/6-em)]PTzBI-Si[thin space (1/6-em)]:[thin space (1/6-em)]N2200 showed a slightly higher μe (1.02 × 10−3 cm2 V−1 s−1) than the PBTA-Si[thin space (1/6-em)]:[thin space (1/6-em)]N2200 blend (μe = 8.76 × 10−4 cm2 V−1 s−1) and the PTzBI-Si[thin space (1/6-em)]:[thin space (1/6-em)]N2200 blend (μe = 5.63 × 10−4 cm2 V−1 s−1). The corresponding ternary devices exhibited a high PCE of 9.56% with an impressive FF of 75.7%. Further investigations showed that the all-polymer OSCs achieved a PCE of 9.17% with an active layer of 350 nm and maintained a PCE of 8.34% with a 420 nm thickness active layer. A wide bandgap copolymer PEG-2%, which had the benzodithiophene-alt-difluorobenzotriazole as the backbone and a polyethylene glycol (PEG) modified side chain, was designed by Cao et al. to fabricate all-polymer ternary OSCs.122 The flexible side chains made PEG-2% soluble in the green solvent of MeTHF. At the optimal weight ratio (1.4[thin space (1/6-em)]:[thin space (1/6-em)]0.6[thin space (1/6-em)]:[thin space (1/6-em)]1) of the PEG-2%[thin space (1/6-em)]:[thin space (1/6-em)]PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]N2200 blend, the resulting device achieved the highest μh (9.51 × 10−4 cm2 V−1 s−1) and μe (5.68 × 10−4 cm2 V−1 s−1) with a balanced μh/μe of 1.67, contributing to its enhanced Jsc (16.89 mA cm−2) and favorable FF (66.13%). AFM measurements suggested that the incorporation of PTB7-Th increased the miscibility of PEG-2% and N2200, resulting in a reduced RMS roughness (0.90 nm) compared with the PEG-2%[thin space (1/6-em)]:[thin space (1/6-em)]N2200 blend film (RMS = 1.30 nm). As a result, the PEG-2%[thin space (1/6-em)]:[thin space (1/6-em)]N2200 binary device's PCE of 5.98% was increased to 9.27% in the PEG-2%[thin space (1/6-em)]:[thin space (1/6-em)]PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]N2200 all-polymer ternary OSC. Cao's group used a narrow bandgap conjugated copolymer (PNTB) containing the naphtho[1,2-c:5,6-c′]bis([1,2,5]thiadiazole) unit to blend with the PBTA-BO[thin space (1/6-em)]:[thin space (1/6-em)]N2200 film.123 The increased weight ratio of PNTB gradually improved the EQE response in the 650–800 nm region, and the ternary blend film containing 30% PNTB exhibited the optimum EQE response, contributing to the highest Jsc of 15.77 mA cm−2. Upon adding PNTB to a PBTA-BO[thin space (1/6-em)]:[thin space (1/6-em)]N2200 binary blend, a reduced RMS roughness of 1.4 nm was observed, demonstrating a smoother surface morphology than that of the binary blend film (RMS = 4.67 nm). A larger (100) crystal coherence length was also observed from ternary blends compared with two binary blends, indicating improved crystallinity. Such an optimized morphology enabled good charge transport of ternary blends, resulting in a high PCE of 10.09% and a desirable FF of 74.98%. Another high-performance all-polymer ternary OSC achieving a PCE of 10.12% and a high FF of 78% was reported by Cao et al.25 The ternary systems comprised PTzBI[thin space (1/6-em)]:[thin space (1/6-em)]PBTA-BO donors and N2200 as an acceptor. AFM showed that the PTzBI[thin space (1/6-em)]:[thin space (1/6-em)]N2200 binary blend films exhibited granular aggregates with an RMS roughness of 1.37 nm, while the ternary blend films with 30% PBTA-BO formed smoother fibrous structures with a decreased RMS roughness of 1.29 nm. In addition, when the weight ratio of PBTA-BO increased from 0% to 30%, a strong face-on orientation was observed in GIWAXS patterns, which can help reduce the structural disorder and thereby enhance the charge transport. Both enhanced charge mobility and the desirable morphology accounted for the superior PCE and FF.

In 2019, Han and coworkers reported a high-performance all-polymer ternary OSC using PTB7-Th as an additional polymer donor to blend with a J51[thin space (1/6-em)]:[thin space (1/6-em)]N2200 system.124 The EQE of the ternary blend exhibited a dramatic improvement in the range from 640 to 850 nm, suggesting that the addition of PTB7-Th contributed to the light absorption in the long-wavelength regions. The increased PTB7-Th content in the ternary blend may enhance the crystallinity of N2200, resulting in the improvement of electron transport. At a ratio of 1.4[thin space (1/6-em)]:[thin space (1/6-em)]0.6[thin space (1/6-em)]:[thin space (1/6-em)]1 for J51[thin space (1/6-em)]:[thin space (1/6-em)]PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]N2200, the ternary blend film exhibited balanced electron–hole mobility (μh = 3.54 × 10−4 cm2 V−1 s−1, μe = 3.17 × 10−4 cm2 V−1 s−1), which was consistent with an enhanced Jsc of 17.17 mA cm−2 and an FF of 66.3%, and hence the PCE reached 9.60%. More importantly, the authors used a green solvent cyclopentyl methyl ether (CPME) to fabricate the ternary devices, which was profitable for practical applications.

Han et al. chose PCDTBT as a sensitizer of PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]N2200 binary blend to fabricate ternary OSCs.125 The HOMO and LUMO energy levels of PCDTBT were located between those of PTB7-Th and N2200, resulting in an energy cascade level for efficient charge transfer. The addition of PCDTBT also broadened the absorption of PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]N2200 in the 300–420 nm range, which enhanced light harvesting. More importantly, the interaction between PTB7-Th and PCDTBT promoted the aggregation of PTB7-Th, leading to decreased domain sizes and enlarged D/A interfaces. As a result, the ternary device with a 15% PCDTBT content gave a PCE of 5.11%, a Voc of 0.78 V, a Jsc of 11.44 mA cm−2, and an FF of 56%. Liu et al. developed a novel principle of dual FRET in a PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]PF12TBT[thin space (1/6-em)]:[thin space (1/6-em)]P(NDI2OD-T2) system, in which the third component (PF12TBT) could transfer energy to both the donor and acceptor.126 The photocurrent generation process for the PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]P(NDI2OD-T2)[thin space (1/6-em)]:[thin space (1/6-em)]PF12TBT blend system is illustrated in Fig. 19. The introduction of PF12TBT could provide a nucleus to induce more and smaller microcrystals in the blend film, leading to decreased phase separation for efficient photon absorption and exciton dissociation. In addition, 3-hexylthiophene was selected as a solvent additive to improve the dispersity of PF12TBT, enabling efficient dual FRET. The corresponding ternary devices yielded a PCE of 6.07%, a Voc of 0.81 V, a Jsc of 12.28 mA cm−2, and an FF of 61%.


image file: d0nr07788g-f19.tif
Fig. 19 Steps in the photocurrent generation process for a PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]P(NDI2OD-T2)[thin space (1/6-em)]:[thin space (1/6-em)]PF12TBT blend system: ① exciton formation, ② energy transfer, ③ exciton separation, and ④ carrier transport. Reproduced with permission.126 Copyright 2017, Royal Society of Chemistry.

He et al. reported an efficient all-polymer ternary OSC by introducing a chlorinated polymer donor PBClT as the third component into PTB7-Th[thin space (1/6-em)]:[thin space (1/6-em)]NDP-V-C7 blend films.127 Compared with PTB7-Th, the LUMO and HOMO energy levels of PBClT decreased due to the strong electrophilic effect of the two chlorine atoms. The formed cascade energy levels were beneficial for charge transfer. Moreover, the introduction of PBClT could also enhance miscibility and suppress crystallization in ternary blends, resulting in a favorable film morphology. As a result, the best-performing device with 15% PBClT exhibited a PCE of 9.03%, a Voc of 0.78 V, a Jsc of 16.77 mA cm−2, and an FF of 68.07%.

To improve the performance of all-polymer ternary OSCs, wide bandgap polymer donors are often incorporated into the host blend to complement the absorption of narrow bandgap polymer acceptors. But all-polymer blend films often exhibit large phase separation, which may lead to decreased FFs of the resulting devices. Therefore, improving the miscibility of polymers is a crucial issue in all-polymer systems.

2.4. All-small-molecule ternary OSCs

Compared with all-polymer OSCs, all-small-molecule OSCs have attracted increasing attention and achieved significant improvement due to the unique features of small molecules, such as their well-defined structure and molecular weight, high purity, and less batch-to-batch variation.128–133 The structures of donors and acceptors for all-small-molecule ternary OSCs are shown in Fig. 20, and the photovoltaic parameters are summarized in Table 7.
image file: d0nr07788g-f20.tif
Fig. 20 The structures of donors and acceptors for all-small-molecule ternary OSCs.
Table 7 Photovoltaic parameters of ternary OSCs with all-small-molecule blends
D1 D2 (third component) A Weight ratio (D1[thin space (1/6-em)]:[thin space (1/6-em)]D2[thin space (1/6-em)]:[thin space (1/6-em)]A) V oc (V) J sc (mA cm−2) FF (%) PCE (%) Working mechanism Ref.
DRCN5T DR3TSBDT PC71BM 0.7[thin space (1/6-em)]:[thin space (1/6-em)]0.3[thin space (1/6-em)]:[thin space (1/6-em)]0.8 0.92 16.5 67 10.16 Alloy model 134
DR3TBDTT DR3TBDTT-E PC71BM 0.9[thin space (1/6-em)]:[thin space (1/6-em)]0.1[thin space (1/6-em)]:[thin space (1/6-em)]0.8 0.896 14.97 76.5 10.26 Alloy model 135
DR3TBDTT DR3TBDTT-S-E PC71BM 0.95[thin space (1/6-em)]:[thin space (1/6-em)]0.05[thin space (1/6-em)]:[thin space (1/6-em)]0.8 0.91 14.89 76.9 10.38 136
DCAO3TBDTT DR3TBDTT-S-E IDIC 0.95[thin space (1/6-em)]:[thin space (1/6-em)]0.05[thin space (1/6-em)]:[thin space (1/6-em)]1 0.91 16.37 67.4 10.04 136
DPP-2TTP DR3TBDTTF PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]0.2[thin space (1/6-em)]:[thin space (1/6-em)]1.5 0.85 17.89 77.1 11.15 Energy transfer 26
DPPEZnP-BzTBO DPPEZnP-TBO PC61BM 0.8[thin space (1/6-em)]:[thin space (1/6-em)]0.2[thin space (1/6-em)]:[thin space (1/6-em)]1.2 0.78 18.60 70.10 10.17 Charge transfer 137
BTR DIB-SQ PC71BM 0.94[thin space (1/6-em)]:[thin space (1/6-em)]0.06[thin space (1/6-em)]:[thin space (1/6-em)]1 0.90 15.44 73.8 10.3 Energy transfer 138
DRCN5T BTR PC71BM 0.985[thin space (1/6-em)]:[thin space (1/6-em)]0.015[thin space (1/6-em)]:[thin space (1/6-em)]0.8 0.90 16.54 67.5 10.05 Energy transfer 139
DPPEZnP-TEH DPP(TBFu)2 PC61BM 0.8[thin space (1/6-em)]:[thin space (1/6-em)]0.2[thin space (1/6-em)]:[thin space (1/6-em)]1 0.79 14.5 62 7.1 140
DR3 ICC6 PC71BM 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0.4 0.87 16.3 72 10.8 Energy transfer and charge transfer 141


In 2016, Deng et al. fabricated a series of all-small-molecule ternary OSCs with DRCN5T and DR3TSBDT as donors and PC71BM as the acceptor.134 DRCN5T exhibited a wider absorption range than DR3TSBDT, showing a complementary absorption in the range from 300 nm to 800 nm. The two donors with good compatibility preferred to form an alloy state, which would provide efficient hole transport channels in ternary active layers. Moreover, the nano-scale phase separation and bi-continuous interpenetrating network were well maintained in the optimized ternary blend films, which were beneficial for efficient exciton dissociation and charge transport, resulting in improved Jsc and FF. By incorporating 30% DR3TSBDT in a DRCN5T[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM blend, a champion PCE of 10.16% was obtained with a Jsc of 16.5 mA cm−2, an FF of 67.0% and a Voc of 0.92 V. Another all-small-molecule ternary OSC consisting of DR3TBDTT and DR3TBDTT-E as donors and PC71BM as the acceptor was fabricated by Wei and coworkers.135 The energy levels of DR3TBDTT-E were lowered due to the introduction of ester functional side chains, and thus energy cascades were formed in the LUMO and HOMO energy levels of three compounds, which was beneficial for charge transfer. As shown in Fig. 21, after thermal and solvent vapor annealing (TSA), the alloy-like model transformed to a cascade model because DR3TBDTT-E shifted from the mixed region to D/A interfaces, which facilitated the charge separation process. At the optimized ratio of DR3TBDTT[thin space (1/6-em)]:[thin space (1/6-em)]DR3TBDTT-E[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM (0.9[thin space (1/6-em)]:[thin space (1/6-em)]0.1[thin space (1/6-em)]:[thin space (1/6-em)]0.8), the ternary solar cell exhibited a maximum PCE of 10.26%.


image file: d0nr07788g-f21.tif
Fig. 21 Diagrammatic sketch of the phase transformation before and after TSA. Reproduced with permission.135 Copyright 2018, American Chemical Society.

In 2019, the same group synthesized a small molecule DR3TBDTT-S-E as a donor to blend with DR3TBDTT[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM and DCAO3TBDTT[thin space (1/6-em)]:[thin space (1/6-em)]IDIC systems, respectively.136 They found that two ternary devices presented a much higher PCE over 10%, but the related mechanisms are different. In a DR3TBDTT[thin space (1/6-em)]:[thin space (1/6-em)]DR3TBDTT-S-E[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM system, DR3TBDTT-S-E could serve as a charge relay to facilitate hole extraction from PC71BM to DR3TBDTT, derived from the cascade energy levels of three materials, while in a DCAO3TBDTT[thin space (1/6-em)]:[thin space (1/6-em)]DR3TBDTT-S-E[thin space (1/6-em)]:[thin space (1/6-em)]IDIC system, the ternary blend film exhibited a mixed face-on and edge-on orientation by incorporating 5% DR3TBDTT-S-E, indicating that the formed 3D charge pathways could promote efficient charge transport. A small molecule DPP-2TTP containing porphyrin and diketopyrrolopyrrole units was designed by Zhu et al. to fabricate ternary OSCs.26 As expected, the DPP-2TTP exhibited a very strong absorption in ranges of 400–550 nm and 700–900 nm, and the weak absorption in the range of 550–650 nm was compensated by a wide-bandgap small molecule DR3TBDTTF. The large overlap between the emission spectrum of DR3TBDTTF and the absorption spectrum of DPP-2TTP suggested efficient energy transfer from DR3TBDTTF to DPP-2TTP, which was helpful in achieving high photocurrents in ternary OSCs. Moreover, the addition of 20% DR3TBDTTF in a DPP-2TTP[thin space (1/6-em)]:[thin space (1/6-em)]DR3TBDTTF[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM blend film can slightly reduce the RMS from 2.35 to 2.08 nm; such a smoother surface was beneficial for more efficient exciton dissociation and charge transport. Thus the corresponding ternary device showed a remarkable PCE of 11.15% with a significant FF of 77.1%, which was much better than that the DPP-2TTP[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM binary device (PCE = 9.30%).

Cao's group reported a highly efficient ternary OSC using two porphyrin derivatives (DPPEZnP-TBO and DPPEZnP-BzTBO) as donors and PC61BM as the acceptor.137 By replacing the alky-thiophene with alky benzothiophene units at two of the porphyrin meso positions, DPPEZnP-BzTBO showed deeper HOMO and LUMO energy levels than DPPEZnP-TBO, forming a cascade structure to facilitate charge transportation. With 20% incorporation of DPPEZnP-TBO, the EQE values of the ternary devices improved significantly in the region of 630–810 nm and up to 65% in the 700–800 nm wavelength range. Compared with DPPEZnP-BzTBO[thin space (1/6-em)]:[thin space (1/6-em)]PC61BM binary devices, the μh of the resulting ternary device was enhanced from 3.93 × 10−4 to 5.52 × 10−4 cm2 V−1 s−1, and the μh was slightly enhanced from 2.38 × 10−4 cm2 V−1 s−1 to 2.53 × 10−4 cm2 V−1 s−1, contributing to the higher Jsc and FF. As a result, the ternary device achieved a PCE up to 10.17% with an improved Jsc of 18.60 mA cm2 and an FF of 70.10%.

Liu et al. incorporated a small molecule DIB-SQ into a BTR[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM system.138 The 2,4-bis[4-(N,N-diisobutylamino)-2,6-dihydroxyphenyl] squaraine (DIB-SQ) was selected as the third component to enhance photon harvesting of the active layer due to its high extinction coefficient of 4.5 × 105 M−1 cm−1 in the long wavelength range. The PL emission intensity was obviously decreased along with increasing contents of DIB-SQ, indicating efficient energy transfer from BTR to DIB-SQ. The incorporation of 6% DIB-SQ decreased the crystallinity of BTR in the ternary blend film, forming an optimized morphology with a fibrillar structure. Therefore, balanced charge transport (μh/μe = 1.07) can be achieved in the ternary active layer. The optimized ternary device with 6% DIB-SQ in the active layer yielded a champion PCE of 10.3% with a Jsc of 15.44 mA cm−2, a Voc of 0.90 V and an FF of 73.8%. Tang et al. selected a high-crystallinity small molecule BTR as the second donor and morphology regulator to prepare ternary OSCs.139 The EQE values of ternary devices improved in the whole spectral range due to enhanced photon harvesting in the short wavelength range on mixing 1.5% BTR. The crystal BTR may induce the ordered packing of DRCN5T in a DRCN5T[thin space (1/6-em)]:[thin space (1/6-em)]BTR[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM ternary blend, leading to the uniform nano-fibrillar structure for efficient charge transport. Consequently, more balanced charge mobility with a μh/μe of 1.12 was obtained in the ternary active layer, resulting in an increased FF. The optimal ternary device achieved a PCE of 10.05% with a Jsc of 16.54 mA cm−2, a Voc of 0.90 V and an FF of 67.5%.

Zhu et al. improved the performance of all-small-molecule ternary OSCs through controlling the morphology of blend films.140 In their work, a crystalline compound DPP(TBFu)2 was selected as the third component to blend with a DPPEZnP-THE[thin space (1/6-em)]:[thin space (1/6-em)]PC61BM system. The addition of DPP(TBFu)2 could lead to phase separation with PC61BM due to accelerated crystallization under thermal annealing, which enhanced the crystallinity and domain purity of the host donor. The ternary device with a 20% DPP(TBFu)2 content showed the best PCE of 7.1% with a Voc of 0.79 V, a Jsc of 14.5 mA cm−2, and an FF of 62%.

Laquai et al. revealed the complex photophysics of DR3[thin space (1/6-em)]:[thin space (1/6-em)]ICC6[thin space (1/6-em)]:[thin space (1/6-em)]PC71BM-based ternary OSCs for the first time.141 It was found that the excitation of donor DR3 or acceptor ICC6 resulted in fast charge transfer, while the excitation of PC71BM led to fast singlet energy transfer to ICC6. The incorporation of PC71BM increased the crystallinity of ICC6, resulting in an improved charge carrier mobility. This improvement enhanced the EQE across the entire absorption range of the ternary blend. The optimized ternary devices yielded an average PCE of 10.8%, with an FF of 72%, a Voc of 0.87 V, and a Jsc of 16.3 mA cm−2.

As discussed above, the introduction of high-crystallinity small molecules could induce the ordered molecular packing of the host donor, leading to the formation of a uniform fibrillar structure for efficient charge transport in the ternary blend film. However, small molecules would prefer to form large crystalline domains due to the strong molecular interactions, which is unfavorable for exciton dissociation and charge transport. To optimize the morphology of ternary blends, manipulating the complicated molecular interactions plays a key role for enhanced efficiencies.

3. Challenges and strategies

Although ternary OSCs have achieved remarkable PCEs, there are still some key issues and challenges that should be noted and addressed to further enhance the photovoltaic performance.

(i) Designing and synthesizing new donor materials are crucial to promote the performance of ternary OSCs. Narrow bandgap non-fullerene acceptors have strong absorption in the long wavelength region (700–950 nm) with high absorption coefficients. Devices based on non-fullerene acceptors have exhibited very impressive photovoltaic performance with PCEs over 17% in binary solar cells. Therefore, there is an urgent need to develop high-performance medium bandgap and wide bandgap donor materials having matched energy levels and complementary absorption with non-fullerene acceptors.

(ii) Increasing the miscibility of each component is required for optimizing the active layer morphology, which is vital to photovoltaic performance. Due to the high crystallinity of non-fullerene acceptors, the tendency to achieve severe phase separation may prevent efficient exciton dissociation and charge transport. Furthermore, a poor miscibility between polymers and small molecules would limit the composition tolerance of ternary systems. These issues could be addressed through the rational design of photovoltaic materials and the optimization of processing conditions.

(iii) Investigating the Voc behavior of ternary devices is of equal importance. In a binary OSC, the Voc is related to the difference between the HOMO energy level of the donor and the LUMO energy level of the acceptor. But it will become complicated when it comes to ternary solar cells. The Voc of a ternary device is dependent on the ratios of used materials and is sensitive to the morphology of the ternary blend. More studies are needed to provide a deep understanding of the Voc for ternary solar cells.

(iv) Compared with the numerous studies on improving the PCE of OSCs, the research on device stabilities has attracted much less attention. Achieving long-term stability is a crucial issue for realizing commercial applications. Although some studies have reported better stabilities of ternary OSCs than binary devices, there is still a long way to go to achieve long-term stabilities. To improve the device stabilities, the following aspects need to be paid attention. (a) Enhance the chemical stabilities of photovoltaic materials, such as the resistance to photodegradation and heat degradation.142 (b) Improve the morphological stability of the active layer, such as cross-linking to increase intermolecular interactions which may stabilize the active layer morphology.143 (c) Make the encapsulation technology of OSC devices perfect to avoid exposure to oxygen and/or water vapor.144

To realize commercial applications of OSCs, other challenges also need to be focused on. To date, manufacturing large-area OSC devices with high PCE has been a huge challenge.145–147 The efficiency of large area devices is much lower than that of small area counterparts. Roll-to-roll compatible printing techniques are usually used to fabricate large area OSCs, such as inkjet printing,148 screen printing,149 pad printing,150 gravure printing,151 and offset printing.152 The processing techniques are of vital importance to enhance the efficiency of large area OSCs, and they need to be further improved to realize competitive PCEs with small area counterparts. To achieve commercialization, it is also a big challenge to enhance the thickness tolerance of active layers transforming from the lab-scale to industry-scale.

4. Conclusions and prospects

Overall, the recent advances in ternary OSCs based on polymer/small molecule/small molecule, polymer/polymer/small molecule, all-polymer and all-small-molecule classifications were thoroughly studied. The characteristics of each of the four categories of ternary OSCs were systematically summarized and analyzed. The ternary strategy is an efficient approach to enhance the photovoltaic performance of binary OSCs. Incorporating the third component into a binary blend can broaden the absorption spectra, facilitate exciton dissociation and charge transport, as well as adjust the morphology of the blend. As a result, the device performance and stability can be simultaneously enhanced. The main advantages of utilizing the ternary strategy in OSCs are listed as follows.

Firstly, the available solar spectrum for OSCs usually ranges from ultraviolet to near-infrared, and a wider spectral coverage is beneficial for obtaining a higher Jsc. Compared with binary OSCs, the introduction of the third component into ternary OSCs could easily broaden the absorption and improve the light harvesting ability, which is beneficial for enhancing the Jsc. For example, introducing the third component with a wide bandgap may enhance absorption in the short-wavelength region, and introducing the third component with a narrow bandgap may enhance absorption in the long-wavelength region.

Secondly, cascade charge transfer is an important process in ternary OSCs. When the third component with suitable energy levels is introduced into a binary blend to form a cascade energy alignment, it is possible for the third component to transfer holes and electrons to the host donor and acceptor, respectively. As a result, efficient charge transfer and suppressed charge recombination could be realized.

Thirdly, morphology control has a significant impact on the Voc, Jsc, and FF, which determine the photovoltaic performance of OSCs. The third component could act as a morphological modulator to improve miscibility and compatibility with the binary blend, resulting in better stability, optimized phase separation, and appropriate crystallinity of the ternary blend. Moreover, the crystallization, domain size, and purity can easily be tuned by controlling the weight ratio of the third component.

Several key issues toward further development were also pointed out, including rational materials design, optimization of the active layer morphology, and in-depth understanding of complicated mechanisms in ternary systems. To further improve the device performance, efforts should focus on developing high-performance donors and acceptors, increasing the miscibility of the third component with the binary blend, and controlling the favorable morphology with better stability by device engineering. Owing to the advantages of broad and strong absorption, cascade energy levels, and easily adjustable morphology, ternary OSCs have great potential in the field of organic electronics. The ternary strategy could provide bright prospects to boost the development of OSCs for commercial applications.

Conflicts of interest

There are no conflicts of interest to declare.

Acknowledgements

This work was supported by (1) the National Natural Science Foundation of China (51703104, 51773220, 51802169, and 52073122), the Shandong Provincial Natural Science Foundation (ZR2017BEM035), the China Postdoctoral Science Foundation (2017M612198), and the Qingdao Innovation Program Applied Basic Research Fund (18-2-2-8-jch), (2) the State Key Project of International Cooperation Research (2017YFE0108300 and 2016YFE0110800), (3) the Shandong Double-Hundred Project (2018), (4) the National Plan for Introducing Talents of Discipline to Universities (“111” plan), and (5) the 1st Class Discipline Program of Shandong Province.

Notes and references

  1. C. Lee, S. Lee, G. U. Kim, W. Lee and B. J. Kim, Chem. Rev., 2019, 119, 8028–8086 CrossRef CAS .
  2. Z. Genene, W. Mammo, E. Wang and M. R. Andersson, Adv. Mater., 2019, 31, 1807275 CrossRef .
  3. Y. Cui, H. Yao, J. Zhang, T. Zhang, Y. Wang, L. Hong, K. Xian, B. Xu, S. Zhang, J. Peng, Z. Wei, F. Gao and J. Hou, Nat. Commun., 2019, 10, 2515 CrossRef .
  4. P. Cheng, G. Li, X. W. Zhan and Y. Yang, Nat. Photonics, 2018, 12, 131–142 CrossRef CAS .
  5. Y. Cui, H. Yao, L. Hong, T. Zhang, Y. Xu, K. Xian, B. Gao, J. Qin, J. Zhang, Z. Wei and J. Hou, Adv. Mater., 2019, 31, 1808356 CrossRef .
  6. Y. Dong, Y. Zou, J. Yuan, H. Yang, Y. Wu, C. Cui and Y. Li, Adv. Mater., 2019, 31, 1904601 CrossRef CAS .
  7. C. Zhao, J. Wang, J. Jiao, L. Huang and J. Tang, J. Mater. Chem. C, 2020, 8, 28–43 RSC .
  8. H. Li, J. Wang, Y. Wang, F. Bu, W. Shen, J. Liu, L. Huang, W. Wang, L. A. Belfiore and J. Tang, Sol. Energy Mater. Sol. Cells, 2019, 190, 83–97 CrossRef CAS .
  9. C. Yan, S. Barlow, Z. Wang, H. Yan, A. K. Y. Jen, S. R. Marder and X. Zhan, Nat. Rev. Mater., 2018, 3, 18003 CrossRef CAS .
  10. N. Gasparini, A. Salleo, I. McCulloch and D. Baran, Nat. Rev. Mater., 2019, 4, 229–242 CrossRef .
  11. D. Ding, J. Wang, Z. Du, F. Li, W. Chen, F. Liu, H. Li, M. Sun and R. Yang, J. Mater. Chem. A, 2017, 5, 10430–10436 RSC .
  12. D. Kearns and M. Calvin, J. Chem. Phys., 1958, 29, 950–951 CrossRef CAS .
  13. J. Wang, X. Bao, D. Ding, M. Qiu, Z. Du, J. Wang, J. Liu, M. Sun and R. Yang, J. Mater. Chem. A, 2016, 4, 11729–11737 RSC .
  14. P. Bi and X. Hao, Sol. RRL, 2019, 3, 1800263 CrossRef .
  15. J. Wang, M. Xiao, W. Chen, M. Qiu, Z. Du, W. Zhu, S. Wen, N. Wang and R. Yang, Macromolecules, 2014, 47, 7823–7830 CrossRef CAS .
  16. C. W. Tang, Appl. Phys. Lett., 1986, 48, 183–185 CrossRef CAS .
  17. G. Yu, J. Gao, J. C. Hummelen, F. Wudl and A. J. Heeger, Science, 1995, 270, 1789–1791 CrossRef CAS .
  18. Q. Liu, Y. Jiang, K. Jin, J. Qin, J. Xu, W. Li, J. Xiong, J. Liu, Z. Xiao, K. Sun, S. Yang, X. Zhang and L. Ding, Sci. Bull., 2020, 65, 272–275 CrossRef CAS .
  19. L. Meng, Y. Zhang, X. Wan, C. Li, X. Zhang, Y. Wang, X. Ke, Z. Xiao, L. Ding, R. Xia, H. L. Yip, Y. Cao and Y. Chen, Science, 2018, 361, 1094–1098 CrossRef CAS .
  20. R. Yu, H. Yao, Y. Cui, L. Hong, C. He and J. Hou, Adv. Mater., 2019, 31, 1902302 CrossRef .
  21. L. Zhan, S. Li, T.-K. Lau, Y. Cui, X. Lu, M. Shi, C.-Z. Li, H. Li, J. Hou and H. Chen, Energy Environ. Sci., 2020, 13, 635–645 RSC .
  22. T. Yan, J. Ge, T. Lei, W. Zhang, W. Song, B. Fanady, D. Zhang, S. Chen, R. Peng and Z. Ge, J. Mater. Chem. A, 2019, 7, 25894–25899 RSC .
  23. S. Dai, S. Chandrabose, J. Xin, T. Li, K. Chen, P. Xue, K. Liu, K. Zhou, W. Ma, J. M. Hodgkiss and X. Zhan, J. Mater. Chem. A, 2019, 7, 2268–2274 RSC .
  24. Q. An, J. Wang and F. Zhang, Nano Energy, 2019, 60, 768–774 CrossRef CAS .
  25. Z. Li, L. Ying, R. Xie, P. Zhu, N. Li, W. Zhong, F. Huang and Y. Cao, Nano Energy, 2018, 51, 434–441 CrossRef CAS .
  26. V. Piradi, X. Xu, Z. Wang, J. Ali, Q. Peng, F. Liu and X. Zhu, ACS Appl. Mater. Interfaces, 2019, 11, 6283–6291 CrossRef CAS .
  27. H. Huang, L. Yang and B. Sharma, J. Mater. Chem. A, 2017, 5, 11501–11517 RSC .
  28. Z. Hu, J. Wang, Z. Wang, W. Gao, Q. An, M. Zhang, X. Ma, J. Wang, J. Miao, C. Yang and F. Zhang, Nano Energy, 2019, 55, 424–432 CrossRef CAS .
  29. N. Felekidis, E. Wang and M. Kemerink, Energy Environ. Sci., 2016, 9, 257–266 RSC .
  30. X. Guo, D. Li, Y. Zhang, M. Jan, J. Xu, Z. Wang, B. Li, S. Xiong, Y. Li, F. Liu, J. Tang, C. Duan, M. Fahlman and Q. Bao, Org. Electron., 2019, 71, 65–71 CrossRef CAS .
  31. J. Lee, S. M. Lee, S. Chen, T. Kumari, S. H. Kang, Y. Cho and C. Yang, Adv. Mater., 2019, 31, 1804762 CrossRef .
  32. Z. Hu, F. Zhang, Q. An, M. Zhang, X. Ma, J. Wang, J. Zhang and J. Wang, ACS Energy Lett., 2018, 3, 555–561 CrossRef CAS .
  33. P. P. Khlyabich, B. Burkhart and B. C. Thompson, J. Am. Chem. Soc., 2011, 133, 14534–14537 CrossRef CAS .
  34. Q. An, F. Zhang, L. Li, J. Wang, Q. Sun, J. Zhang, W. Tang and Z. Deng, ACS Appl. Mater. Interfaces, 2015, 7, 3691–3698 CrossRef CAS .
  35. B. Fan, W. Zhong, X.-F. Jiang, Q. Yin, L. Ying, F. Huang and Y. Cao, Adv. Energy Mater., 2017, 7, 1602127 CrossRef .
  36. T. Liu, X. Xue, L. Huo, X. Sun, Q. An, F. Zhang, T. P. Russell, F. Liu and Y. Sun, Chem. Mater., 2017, 29, 2914–2920 CrossRef CAS .
  37. Y. Chen, P. Ye, Z. G. Zhu, X. Wang, L. Yang, X. Xu, X. Wu, T. Dong, H. Zhang, J. Hou, F. Liu and H. Huang, Adv. Mater., 2017, 29, 1603154 CrossRef .
  38. D. Baran, T. Kirchartz, S. Wheeler, S. Dimitrov, M. Abdelsamie, J. Gorman, R. S. Ashraf, S. Holliday, A. Wadsworth, N. Gasparini, P. Kaienburg, H. Yan, A. Amassian, C. J. Brabec, J. R. Durrant and I. McCulloch, Energy Environ. Sci., 2016, 9, 3783–3793 RSC .
  39. H. Li, K. Lu and Z. Wei, Adv. Energy Mater., 2017, 7, 1602540 CrossRef .
  40. R. Yu, H. Yao and J. Hou, Adv. Energy Mater., 2018, 8, 1702814 CrossRef .
  41. W. Li, Y. Yan, Y. Gong, J. Cai, F. Cai, R. S. Gurney, D. Liu, A. J. Pearson, D. G. Lidzey and T. Wang, Adv. Funct. Mater., 2018, 28, 1704212 CrossRef .
  42. T. Zhang, X. Zhao, D. Yang, Y. Tian and X. Yang, Adv. Energy Mater., 2018, 8, 1701691 CrossRef .
  43. S. V. Dayneko, A. D. Hendsbee, J. R. Cann, C. Cabanetos and G. C. Welch, New J. Chem., 2019, 43, 10442–10448 RSC .
  44. N. Y. Doumon, F. V. Houard, J. Dong, P. Christodoulis, M. V. Dryzhov, G. Portale and L. J. A. Koster, J. Mater. Chem. C, 2019, 7, 5104–5111 RSC .
  45. A. Tang, B. Xiao, Y. Wang, F. Gao, K. Tajima, H. Bin, Z.-G. Zhang, Y. Li, Z. Wei and E. Zhou, Adv. Funct. Mater., 2018, 28, 1704507 CrossRef .
  46. Z. G. Zhang, Y. Yang, J. Yao, L. Xue, S. Chen, X. Li, W. Morrison, C. Yang and Y. Li, Angew. Chem., Int. Ed., 2017, 56, 13503–13507 CrossRef CAS .
  47. M. Nam, M. Cha, H. H. Lee, K. Hur, K. T. Lee, J. Yoo, I. K. Han, S. J. Kwon and D. H. Ko, Nat. Commun., 2017, 8, 14068 CrossRef CAS .
  48. Z. Wang, Z. Peng, Z. Xiao, D. Seyitliyev, K. Gundogdu, L. Ding and H. Ade, Adv. Mater., 2020, 2005386 CrossRef CAS .
  49. J. Yuan, Y. Q. Zhang, L. Y. Zhou, G. C. Zhang, H. L. Yip, T. K. Lau, X. H. Lu, C. Zhu, H. J. Peng, P. A. Johnson, M. Leclerc, Y. Cao, J. Ulanski, Y. F. Li and Y. P. Zou, Joule, 2019, 3, 1140–1151 CrossRef CAS .
  50. T. Yan, W. Song, J. Huang, R. Peng, L. Huang and Z. Ge, Adv. Mater., 2019, 31, 1902210 CrossRef .
  51. M.-A. Pan, T.-K. Lau, Y. Tang, Y.-C. Wu, T. Liu, K. Li, M.-C. Chen, X. Lu, W. Ma and C. Zhan, J. Mater. Chem. A, 2019, 7, 20713–20722 RSC .
  52. H. Fu, C. Li, P. Bi, X. Hao, F. Liu, Y. Li, Z. Wang and Y. Sun, Adv. Funct. Mater., 2019, 29, 1807006 CrossRef .
  53. Y. Xie, F. Yang, Y. Li, M. A. Uddin, P. Bi, B. Fan, Y. Cai, X. Hao, H. Y. Woo, W. Li, F. Liu and Y. Sun, Adv. Mater., 2018, 30, 1803045 CrossRef .
  54. H. H. Gao, Y. Sun, X. Wan, X. Ke, H. Feng, B. Kan, Y. Wang, Y. Zhang, C. Li and Y. Chen, Adv. Sci., 2018, 5, 1800307 CrossRef .
  55. W. T. Hadmojo, F. T. A. Wibowo, W. Lee, H. K. Jang, Y. Kim, S. Sinaga, M. Park, S. Y. Ju, D. Y. Ryu, I. H. Jung and S. Y. Jang, Adv. Funct. Mater., 2019, 29, 1808731 CrossRef .
  56. Z. Liang, J. Tong, H. Li, Y. Wang, N. Wang, J. Li, C. Yang and Y. Xia, J. Mater. Chem. A, 2019, 7, 15841–15850 RSC .
  57. C. Xue, T. Zhang, K. Ma, P. Wan, L. Hong, B. Xu and C. An, Macromol. Rapid Commun., 2019, 40, 1900246 CrossRef CAS .
  58. H. Shi, R. Xia, G. Zhang, H.-L. Yip and Y. Cao, Adv. Energy Mater., 2019, 9, 1803438 CrossRef .
  59. D. Huang, F. Bian, D. Zhu, X. Bao, C. Hong, P. Zhou, Y. Huang and C. Yang, J. Phys. Chem. C, 2019, 123, 14976–14984 CrossRef CAS .
  60. W. Zhao, S. Li, S. Zhang, X. Liu and J. Hou, Adv. Mater., 2017, 29, 1604059 CrossRef .
  61. Z. Xiao, X. Jia and L. Ding, Sci. Bull., 2017, 62, 1562–1564 CrossRef CAS .
  62. H. Lu, J. Zhang, J. Chen, Q. Liu, X. Gong, S. Feng, X. Xu, W. Ma and Z. Bo, Adv. Mater., 2016, 28, 9559–9566 CrossRef CAS .
  63. X. Xu, Z. Bi, W. Ma, G. Zhang, H. Yan, Y. Li and Q. Peng, J. Mater. Chem. A, 2019, 7, 14199–14208 RSC .
  64. K. Weng, C. Li, P. Bi, H. S. Ryu, Y. Guo, X. Hao, D. Zhao, W. Li, H. Y. Woo and Y. Sun, J. Mater. Chem. A, 2019, 7, 3552–3557 RSC .
  65. B. Wang, Y. Fu, Q. Yang, J. Wu, H. Liu, H. Tang and Z. Xie, J. Mater. Chem. C, 2019, 7, 10498–10506 RSC .
  66. X. Du, T. Heumueller, W. Gruber, A. Classen, T. Unruh, N. Li and C. J. Brabec, Joule, 2019, 3, 215–226 CrossRef CAS .
  67. T. Liu, Y. Guo, Y. Yi, L. Huo, X. Xue, X. Sun, H. Fu, W. Xiong, D. Meng, Z. Wang, F. Liu, T. P. Russell and Y. Sun, Adv. Mater., 2016, 28, 10008–10015 CrossRef CAS .
  68. X. Song, N. Gasparini, M. M. Nahid, S. H. K. Paleti, J.-L. Wang, H. Ade and D. Baran, Joule, 2019, 3, 846–857 CrossRef CAS .
  69. M. Zhang, W. Gao, F. Zhang, Y. Mi, W. Wang, Q. An, J. Wang, X. Ma, J. Miao, Z. Hu, X. Liu, J. Zhang and C. Yang, Energy Environ. Sci., 2018, 11, 841–849 RSC .
  70. R. Yu, S. Zhang, H. Yao, B. Guo, S. Li, H. Zhang, M. Zhang and J. Hou, Adv. Mater., 2017, 29, 1700437 CrossRef .
  71. L. Yang, W. Gu, L. Hong, Y. Mi, F. Liu, M. Liu, Y. Yang, B. Sharma, X. Liu and H. Huang, ACS Appl. Mater. Interfaces, 2017, 9, 26928–26936 CrossRef CAS .
  72. W. Su, Q. Fan, X. Guo, X. Meng, Z. Bi, W. Ma, M. Zhang and Y. Li, Nano Energy, 2017, 38, 510–517 CrossRef CAS .
  73. T. Ameri, P. Khoram, J. Min and C. J. Brabec, Adv. Mater., 2013, 25, 4245–4266 CrossRef CAS .
  74. L. Lu, M. A. Kelly, W. You and L. Yu, Nat. Photonics, 2015, 9, 491–500 CrossRef CAS .
  75. Y. Ma, X. Zhou, D. Cai, Q. Tu, W. Ma and Q. Zheng, Mater. Horiz., 2020, 7, 117–124 RSC .
  76. J. Song, C. Li, L. Zhu, J. Guo, J. Xu, X. Zhang, K. Weng, K. Zhang, J. Min, X. Hao, Y. Zhang, F. Liu and Y. Sun, Adv. Mater., 2019, 31, 1905645 CrossRef CAS .
  77. Y. Chang, T.-K. Lau, M.-A. Pan, X. Lu, H. Yan and C. Zhan, Mater. Horiz., 2019, 6, 2094–2102 RSC .
  78. M. Zhang, R. Ming, W. Gao, Q. An, X. Ma, Z. Hu, C. Yang and F. Zhang, Nano Energy, 2019, 59, 58–65 CrossRef CAS .
  79. R. Ma, Y. Chen, T. Liu, Y. Xiao, Z. Luo, M. Zhang, S. Luo, X. Lu, G. Zhang, Y. Li, H. Yan and K. Chen, J. Mater. Chem. C, 2020, 8, 909–915 RSC .
  80. J. Gao, R. Ming, Q. An, X. Ma, M. Zhang, J. Miao, J. Wang, C. Yang and F. Zhang, Nano Energy, 2019, 63, 103888 CrossRef CAS .
  81. Y. Chang, X. Zhang, Y. Tang, M. Gupta, D. Su, J. Liang, D. Yan, K. Li, X. Guo, W. Ma, H. Yan and C. Zhan, Nano Energy, 2019, 64, 103934 CrossRef CAS .
  82. L. Wu, L. Xie, H. Tian, R. Peng, J. Huang, B. Fanady, W. Song, S. Tan, W. Bi and Z. Ge, Sci. Bull., 2019, 64, 1087–1094 CrossRef CAS .
  83. R. Lv, D. Chen, X. Liao, L. Chen and Y. Chen, Adv. Funct. Mater., 2019, 29, 1805872 CrossRef .
  84. F. Pan, L. Zhang, H. Jiang, D. Yuan, Y. Nian, Y. Cao and J. Chen, J. Mater. Chem. A, 2019, 7, 9798–9806 RSC .
  85. X. Ma, M. Luo, W. Gao, J. Yuan, Q. An, M. Zhang, Z. Hu, J. Gao, J. Wang, Y. Zou, C. Yang and F. Zhang, J. Mater. Chem. A, 2019, 7, 7843–7851 RSC .
  86. Z. Liu and N. Wang, J. Mater. Chem. C, 2019, 7, 10039–10048 RSC .
  87. H. Jiang, X. Li, J. Wang, S. Qiao, Y. Zhang, N. Zheng, W. Chen, Y. Li and R. Yang, Adv. Funct. Mater., 2019, 29, 1903596 CrossRef .
  88. H. Jiang, X. Li, Z. Liang, G. Huang, W. Chen, N. Zheng and R. Yang, J. Mater. Chem. A, 2019, 7, 7760–7765 RSC .
  89. H. Huang, X. Li, S. Chen, B. Qiu, J. Du, L. Meng, Z. Zhang, C. Yang and Y. Li, J. Mater. Chem. A, 2019, 7, 27423–27431 RSC .
  90. L. Zhong, H. Bin, Y. Li, M. Zhang, J. Xu, X. Li, H. Huang, Q. Hu, Z.-Q. Jiang, J. Wang, C. Zhang, F. Liu, T. P. Russell, Z. Zhang and Y. Li, J. Mater. Chem. A, 2018, 6, 24814–24822 RSC .
  91. M. Zhang, Z. Xiao, W. Gao, Q. Liu, K. Jin, W. Wang, Y. Mi, Q. An, X. Ma, X. Liu, C. Yang, L. Ding and F. Zhang, Adv. Energy Mater., 2018, 8, 1801968 CrossRef .
  92. P. Xue, Y. Xiao, T. Li, S. Dai, B. Jia, K. Liu, J. Wang, X. Lu, R. P. S. Han and X. Zhan, J. Mater. Chem. A, 2018, 6, 24210–24215 RSC .
  93. T. Liu, R. Ma, Z. Luo, Y. Guo, G. Zhang, Y. Xiao, T. Yang, Y. Chen, G. Li, Y. Yi, X. Lu, H. Yan and B. Tang, Energy Environ. Sci., 2020, 13, 2115–2123 RSC .
  94. J. Zhang, Y. Zhang, J. Fang, K. Lu, Z. Wang, W. Ma and Z. Wei, J. Am. Chem. Soc., 2015, 137, 8176–8183 CrossRef CAS .
  95. T. Kumari, S. M. Lee, S.-H. Kang, S. Chen and C. Yang, Energy Environ. Sci., 2017, 10, 258–265 RSC .
  96. R. Fan, L. Gu, X. Li, G. Fu and S. Yang, Org. Electron., 2017, 41, 209–214 CrossRef CAS .
  97. X. Ma, F. Zhang, Q. An, Q. Sun, M. Zhang, J. Miao, Z. Hu and J. Zhang, J. Mater. Chem. A, 2017, 5, 13145–13153 RSC .
  98. G. Zhang, K. Zhang, Q. Yin, X. F. Jiang, Z. Wang, J. Xin, W. Ma, H. Yan, F. Huang and Y. Cao, J. Am. Chem. Soc., 2017, 139, 2387–2395 CrossRef CAS .
  99. Y. Zhang, D. Deng, K. Lu, J. Zhang, B. Xia, Y. Zhao, J. Fang and Z. Wei, Adv. Mater., 2015, 27, 1071–1076 CrossRef CAS .
  100. J. Fang, D. Deng, J. Zhang, Y. Zhang, K. Lu and Z. Wei, Mater. Chem. Front., 2017, 1, 1223–1228 RSC .
  101. B. Xia, L. Yuan, J. Zhang, Z. Wang, J. Fang, Y. Zhao, D. Deng, W. Ma, K. Lu and Z. Wei, J. Mater. Chem. A, 2017, 5, 9859–9866 RSC .
  102. M. Zhang, F. Zhang, Q. An, Q. Sun, W. Wang, J. Zhang and W. Tang, Nano Energy, 2016, 22, 241–254 CrossRef CAS .
  103. X. Du, J. Zhao, H. Zhang, X. Lu, L. Zhou, Z. Chen, H. Lin, C. Zheng and S. Tao, J. Mater. Chem. A, 2019, 7, 20139–20150 RSC .
  104. X. Liu, Y. Yan, Y. Yao and Z. Liang, Adv. Funct. Mater., 2018, 28, 1802004 CrossRef .
  105. X. Ma, Q. An, O. A. Ibraikulov, P. Lévêque, T. Heiser, N. Leclerc, X. Zhang and F. Zhang, J. Mater. Chem. A, 2020, 8, 1265–1272 RSC .
  106. H. Hu, L. Ye, M. Ghasemi, N. Balar, J. J. Rech, S. J. Stuard, W. You, B. T. O'Connor and H. Ade, Adv. Mater., 2019, 31, 1808279 CrossRef .
  107. Y. Xie, T. Li, J. Guo, P. Bi, X. Xue, H. S. Ryu, Y. Cai, J. Min, L. Huo, X. Hao, H. Y. Woo, X. Zhan and Y. Sun, ACS Energy Lett., 2019, 4, 1196–1203 CrossRef CAS .
  108. Z. Wang, Y. Nian, H. Jiang, F. Pan, Z. Hu, L. Zhang, Y. Cao and J. Chen, Org. Electron., 2019, 69, 174–180 CrossRef .
  109. L. Zhang, X. Xu, B. Lin, H. Zhao, T. Li, J. Xin, Z. Bi, G. Qiu, S. Guo, K. Zhou, X. Zhan and W. Ma, Adv. Mater., 2018, 30, 1805041 CrossRef .
  110. X. Ma, Y. Mi, F. Zhang, Q. An, M. Zhang, Z. Hu, X. Liu, J. Zhang and W. Tang, Adv. Energy Mater., 2018, 8, 1702854 CrossRef .
  111. L. Zhong, L. Gao, H. Bin, Q. Hu, Z.-G. Zhang, F. Liu, T. P. Russell, Z. Zhang and Y. Li, Adv. Energy Mater., 2017, 7, 1602215 CrossRef .
  112. X. Xu, Z. Bi, W. Ma, Z. Wang, W. C. H. Choy, W. Wu, G. Zhang, Y. Li and Q. Peng, Adv. Mater., 2017, 29, 1704271 CrossRef .
  113. N. Gasparini, S. Kahmann, M. Salvador, J. D. Perea, A. Sperlich, A. Baumann, N. Li, S. Rechberger, E. Spiecker, V. Dyakonov, G. Portale, M. A. Loi, C. J. Brabec and T. Ameri, Adv. Energy Mater., 2019, 9, 1803394 CrossRef .
  114. Y. Xie, L. Huo, B. Fan, H. Fu, Y. Cai, L. Zhang, Z. Li, Y. Wang, W. Ma, Y. Chen and Y. Sun, Adv. Funct. Mater., 2018, 28, 1800627 CrossRef .
  115. F. Zhao, Y. Li, Z. Wang, Y. Yang, Z. Wang, G. He, J. Zhang, L. Jiang, T. Wang, Z. Wei, W. Ma, B. Li, A. Xia, Y. Li and C. Wang, Adv. Energy Mater., 2017, 7, 1602552 CrossRef .
  116. C. Wang, W. Zhang, X. Meng, J. Bergqvist, X. Liu, Z. Genene, X. Xu, A. Yartsev, O. Inganäs, W. Ma, E. Wang and M. Fahlman, Adv. Energy Mater., 2017, 7, 1700390 CrossRef .
  117. T. H. Lee, M. A. Uddin, C. Zhong, S.-J. Ko, B. Walker, T. Kim, Y. J. Yoon, S. Y. Park, A. J. Heeger, H. Y. Woo and J. Y. Kim, Adv. Energy Mater., 2016, 6, 1600637 CrossRef .
  118. W. Su, Q. Fan, X. Guo, B. Guo, W. Li, Y. Zhang, M. Zhang and Y. Li, J. Mater. Chem. A, 2016, 4, 14752–14760 RSC .
  119. H. Benten, T. Nishida, D. Mori, H. Xu, H. Ohkita and S. Ito, Energy Environ. Sci., 2016, 9, 135–140 RSC .
  120. Z. Li, X. Xu, W. Zhang, X. Meng, Z. Genene, W. Ma, W. Mammo, A. Yartsev, M. R. Andersson, R. A. J. Janssen and E. Wang, Energy Environ. Sci., 2017, 10, 2212–2221 RSC .
  121. B. Fan, P. Zhu, J. Xin, N. Li, L. Ying, W. Zhong, Z. Li, W. Ma, F. Huang and Y. Cao, Adv. Energy Mater., 2018, 8, 1703085 CrossRef .
  122. Z. Li, B. Fan, B. He, L. Ying, W. Zhong, F. Liu, F. Huang and Y. Cao, Sci. China: Chem., 2018, 61, 427–436 Search PubMed .
  123. Z. Li, R. Xie, W. Zhong, B. Fan, J. Ali, L. Ying, F. Liu, N. Li, F. Huang and Y. Cao, Sol. RRL, 2018, 2, 1800196 CrossRef .
  124. Q. Zhang, Z. Chen, W. Ma, Z. Xie, J. Liu, X. Yu and Y. Han, ACS Appl. Mater. Interfaces, 2019, 11, 32200–32208 CrossRef CAS .
  125. R. Zhang, H. Yang, K. Zhou, J. Zhang, J. Liu, X. Yu, R. Xing and Y. Han, J. Polym. Sci., Part B: Polym. Phys., 2016, 54, 1811–1819 CrossRef CAS .
  126. J. Liu, B. Tang, Q. Liang, Y. Han, Z. Xie and J. Liu, RSC Adv., 2017, 7, 13289–13298 RSC .
  127. H. Chen, Y. Guo, P. Chao, L. Liu, W. Chen, D. Zhao and F. He, Sci. China: Chem., 2018, 62, 238–244 CrossRef .
  128. Z. Zhou, S. Xu, J. Song, Y. Jin, Q. Yue, Y. Qian, F. Liu, F. Zhang and X. Zhu, Nat. Energy, 2018, 3, 952–959 CrossRef CAS .
  129. R.-Z. Liang, Y. Zhang, V. Savikhin, M. Babics, Z. Kan, M. Wohlfahrt, N. Wehbe, S. Liu, T. Duan, M. F. Toney, F. Laquai and P. M. Beaujuge, Adv. Energy Mater., 2019, 9, 1802836 CrossRef .
  130. N. D. Eastham, A. S. Dudnik, B. Harutyunyan, T. J. Aldrich, M. J. Leonardi, E. F. Manley, M. R. Butler, T. Harschneck, M. A. Ratner, L. X. Chen, M. J. Bedzyk, F. S. Melkonyan, A. Facchetti, R. P. H. Chang and T. J. Marks, ACS Energy Lett., 2017, 2, 1690–1697 CrossRef CAS .
  131. L. Yang, S. Zhang, C. He, J. Zhang, H. Yao, Y. Yang, Y. Zhang, W. Zhao and J. Hou, J. Am. Chem. Soc., 2017, 139, 1958–1966 CrossRef CAS .
  132. Q. An, F. Zhang, J. Zhang, W. Tang, Z. Deng and B. Hu, Energy Environ. Sci., 2016, 9, 281–322 RSC .
  133. H. Wang, J. Cao, J. Yu, Z. Zhang, R. Geng, L. Yang and W. Tang, J. Mater. Chem. A, 2019, 7, 4313–4333 RSC .
  134. Q. An, F. Zhang, X. Yin, Q. Sun, M. Zhang, J. Zhang, W. Tang and Z. Deng, Nano Energy, 2016, 30, 276–282 CrossRef CAS .
  135. Z. Wang, X. Zhu, J. Zhang, K. Lu, J. Fang, Y. Zhang, Z. Wang, L. Zhu, W. Ma, Z. Shuai and Z. Wei, J. Am. Chem. Soc., 2018, 140, 1549–1556 CrossRef CAS .
  136. Y. Chang, Y. Chang, X. Zhu, X. Zhou, C. Yang, J. Zhang, K. Lu, X. Sun and Z. Wei, Adv. Energy Mater., 2019, 9, 1900190 CrossRef .
  137. L. Xiao, T. Liang, K. Gao, T. Lai, X. Chen, F. Liu, T. P. Russell, F. Huang, X. Peng and Y. Cao, ACS Appl. Mater. Interfaces, 2017, 9, 29917–29923 CrossRef CAS .
  138. M. Zhang, J. Wang, F. Zhang, Y. Mi, Q. An, W. Wang, X. Ma, J. Zhang and X. Liu, Nano Energy, 2017, 39, 571–581 CrossRef CAS .
  139. M. Zhang, F. Zhang, Q. An, Q. Sun, W. Wang, X. Ma, J. Zhang and W. Tang, J. Mater. Chem. A, 2017, 5, 3589–3598 RSC .
  140. W.-Y. Tan, K. Gao, J. Zhang, L.-L. Chen, S.-P. Wu, X.-F. Jiang, X.-B. Peng, Q. Hu, F. Liu, H.-B. Wu, Y. Cao and X.-H. Zhu, Sol. RRL, 2017, 1, 1600003 CrossRef .
  141. S. Karuthedath, Y. Firdaus, R. Z. Liang, J. Gorenflot, P. M. Beaujuge, T. D. Anthopoulos and F. Laquai, Adv. Energy Mater., 2019, 9, 1901443 CrossRef .
  142. T. Kim, R. Younts, W. Lee, S. Lee, K. Gundogdu and B. J. Kim, J. Mater. Chem. A, 2017, 5, 22170–22179 RSC .
  143. P. Cheng and X. Zhan, Chem. Soc. Rev., 2016, 45, 2544–2582 RSC .
  144. N. Kim, W. J. Potscavage, A. Sundaramoothi, C. Henderson, B. Kippelen and S. Graham, Sol. Energy Mater. Sol. Cells, 2012, 101, 140–146 CrossRef CAS .
  145. Y. Lin, S. Dong, Z. Li, W. Zheng, J. Yang, A. Liu, W. Cai, F. Liu, Y. Jiang, T. P. Russell, F. Huang, E. Wang and L. Hou, Nano Energy, 2018, 46, 428–435 CrossRef CAS .
  146. Y. Lin, C. Cai, Y. Zhang, W. Zheng, J. Yang, E. Wang and L. Hou, J. Mater. Chem. A, 2017, 5, 4093–4102 RSC .
  147. C. Cai, Y. Zhang, R. Song, Z. Peng, L. Xia, M. Wu, K. Xiong, B. Wang, Y. Lin, X. Xu, Q. Liang, H. Wu, E. Wang and L. Hou, Sol. Energy Mater. Sol. Cells, 2017, 161, 52–61 CrossRef CAS .
  148. C. N. Hoth, S. A. Choulis, P. Schilinsky and C. J. Brabec, Adv. Mater., 2007, 19, 3973–3978 CrossRef CAS .
  149. F. C. Krebs, M. Jørgensen, K. Norrman, O. Hagemann, J. Alstrup, T. D. Nielsen, J. Fyenbo, K. Larsen and J. Kristensen, Sol. Energy Mater. Sol. Cells, 2009, 93, 422–441 CrossRef CAS .
  150. F. C. Krebs, Sol. Energy Mater. Sol. Cells, 2009, 93, 484–490 CrossRef CAS .
  151. J. M. Ding, A. de la Fuente Vornbrock, C. Ting and V. Subramanian, Sol. Energy Mater. Sol. Cells, 2009, 93, 459–464 CrossRef CAS .
  152. D. Zielke, A. C. Hübler, U. Hahn, N. Brandt, M. Bartzsch, U. Fügmann, T. Fischer, J. Veres and S. Ogier, Appl. Phys. Lett., 2005, 87, 123508 CrossRef .

This journal is © The Royal Society of Chemistry 2021
Click here to see how this site uses Cookies. View our privacy policy here.