Open Access Article
Zefeng Liua,
Jingming Xin*a,
Xingpeng Liua,
Yutong Zhanga,
Mingzhi Duanb,
Yan Zhanga,
Zhenhui Xua,
Junzheng Yana,
Qiuju Liang*b and
Jiangang Liu
*a
aSchool of Electronics and Information, Northwestern Polytechnical University, Xi'an 710129, China. E-mail: jgliu@nwpu.edu.cn; xinjm@nwpu.edu.cn
bSchool of Microelectronics, Northwestern Polytechnical University, Xi'an 710129, China. E-mail: qiujuliang@nwpu.edu.cn
First published on 20th October 2025
Organic solar cells (OSCs) exhibit promising potential for low-cost photovoltaic applications but suffer from high energy losses (Eloss) that critically limit their open-circuit voltage and power conversion efficiency. To address this challenge, this review first outlines the mechanistic origins of Eloss, emphasizing exciton dissociation barriers, charge-transfer state energetics, and recombination pathways, supported by quantitative characterization methodologies. Building on this foundation, material design strategies are analyzed, where molecular synthesis targeting backbone, side-chain, and terminal group optimization alongside ternary blending collectively modulates energy-level alignments to minimize driving-force offsets. Concurrently, morphological control approaches are systematically evaluated, demonstrating that precise regulation of phase separation, crystallinity, and molecular orientation effectively suppresses recombination losses and enhances charge transport. By integrating these advances, this work establishes a unified framework for energy loss minimization, providing critical insights for developing high-performance OSCs.
Broader contextOrganic solar cells (OSCs) have garnered significant attentions as alternative photovoltaic technology due to inherent advantages. Substantial advances in organic semiconductors, morphology control, and interface engineering have elevated the power conversion efficiency (PCE) of OSCs to over 20%. Nevertheless, substantially higher energy loss compared to their inorganic counterparts consistently limits their performance, representing a critical challenge for further industrialization. This discrepancy is largely due to high energy loss (Eloss), which manifests primarily as a reduced open-circuit voltage (VOC) and poses a major obstacle to commercialization. Reducing energy loss is therefore essential for enhancing device performance. This review comprehensively examines recent strategies aimed at mitigating energy loss in high-efficiency OSCs. We begin by outlining the fundamental mechanisms underlying energy loss and summarizing common characterization techniques. We then discuss material-based strategies-including molecular design and the incorporation of a third component-that optimize energy levels to increase VOC and suppress energy loss. Furthermore, we explore morphology-centric approaches that reduce energy loss by enhancing charge transport and inhibiting bimolecular recombination, focusing on control of phase separation, crystallinity, and molecular orientation. By describing these developments, this review establishes design principles for minimizing energy loss and provides insightful perspectives toward achieving higher-performance OSCs. |
Substantial energy loss in OSCs is attributed to their complex photoconversion mechanism. Characterized by low dielectric constants and strong electron-lattice coupling, organic semiconductors generate bounded electron–hole pairs called excitons with substantial coulombic binding energies (≈0.5 eV)17,18 upon photoexcitation, instead of directly producing free charge carriers. The excitons diffuse to donor/acceptor (D/A) interfaces, undergoing conversion into charge-transfer (CT) states through energy-level offsets prior to dissociating into free carriers, a process that requires additional energy to overcome interfacial binding barriers.19,20 Crucially, sub-bandgap absorption is induced by CT states, promoting radiative recombination and establishing the interfacial species as the primary contributors to energy loss. A critical inverse correlation is observed between the CT state energy (ECT) and the magnitude of the D/A energy-level offsets: reduced offsets result in higher ECT values.21 Through strategic molecular engineering, ECT can be systematically elevated to reduce sub-bandgap radiative recombination loss and associated energy dissipation. The energetic difference between the highest occupied molecular orbital (HOMO) of the donor and the lowest unoccupied molecular orbital (LUMO) of the acceptor is enlarged by diminished energy-level offsets,22 thereby increasing VOC. The optical bandgap (Eg) tuning in material systems further reduces energy loss in the device.
Precise energy-level modulation23 and controlled morphology regulation24 are the two primary strategies for mitigating energy loss in OSCs. Energy-level modulation, implemented through molecular synthesis or ternary strategies, reduces the donor HOMO level or elevates the acceptor LUMO level,25,26 thereby enhancing VOC and minimizing energy loss. For example, Li et al.27 oxidized the dichloro terminal groups of BTP-eC9 to prepare BTP-eC9-4ClO, which increased the LUMO energy level of the acceptor and thus improved VOC. Compared with the PM6:BTP-eC9 system, the PM6:BTP-eC9-4ClO system exhibited a decrease in ΔE3 from 0.202 eV to 0.179 eV, a reduction in Eloss from 0.555 eV to 0.530 eV, an increase in VOC from 0.861 V to 0.891 V, and an improvement in PCE from 19.12% to 20.03%. Concurrently, morphology regulation optimizes active layer nanostructures to achieve tailored phase separation, enhanced crystallinity, and favorable molecular orientation.28,29 These structural features promote exciton dissociation and charge transport while suppressing bimolecular recombination. For instance, Zhang et al.30 introduced donor D18-Cl and acceptor AITC into the D18:N3 system. D18-Cl acted as a nucleating agent promoting D18 aggregation, while AITC facilitated N3 aggregation, collectively enhancing crystallinity. Grazing incidence wide-angle X-ray scattering (GIWAXS) revealed an increased crystal coherence length (CCL) in the out-of-plane (010) direction (27.7 Å to 29.2 Å) for the D18:D18-Cl:N3:AITC film compared to D18:N3, indicating improved crystallinity that promoted charge transfer and suppressed recombination. Consequently, the D18:D18-Cl:N3:AITC device showed reduced ΔE2 (0.060 eV to 0.050 eV), lower ΔE3 (0.226 eV to 0.216 eV), decreased Eloss (0.550 eV to 0.533 eV), increased VOC (0.846 V to 0.869 V), and improved PCE (18.3% to 20.1%).
Given the critical role of minimizing energy loss in improving OSC performance, this review provides a systematic analysis of its underlying mechanisms and characterization techniques. Subsequently, we summarize material design strategies focused on energy-level modulation to reduce energy loss, including synthetic approaches for optimizing molecular backbones, side chains, and terminal groups.31–33 Ternary strategies34 are also covered, highlighting methods to elevate acceptor LUMO levels or lower donor HOMO levels, thereby increasing device VOC and mitigating energy loss. Furthermore, we discuss morphology regulation as an effective pathway for energy loss reduction, analyzing control of phase separation, crystallinity, and molecular orientation to achieve optimal active layer structures35,36 that suppress bimolecular recombination and enhance device performance. Finally, we present recent advancements and emerging research directions to deepen the fundamental understanding of energy loss processes, providing critical insights for future studies aimed at minimizing energy loss and advancing OSC technologies.
Generally, Eloss is typically defined relative to the CT state (Fig. 1A). It comprises two components: energy loss during free charge generation (ΔE2) and energy loss from free charge recombination (ECT − qVOC). Furthermore, the recombination loss is further subdivided into radiative (ΔE1) and non-radiative (ΔE3) components. ΔE1 originates from the mismatch between solar radiation and absorbed omnidirectional blackbody radiation above Eg which is inevitable for any OSC once Eg is fixed. ΔE2 is closely related to ΔELE–CT (the energy difference between the localized excitonic state and the CT state). An increase in ΔELE–CT indicates a larger energy offset between the donor and acceptor, corresponding to a stronger driving force for exciton dissociation. In previous studies, it was generally believed that efficient exciton dissociation in OSCs requires a considerable driving force (>0.3 eV),44,45 leading to a large ΔE2. Currently, ΔE2 can be reduced by continuously decreasing the energy offset between the donor and acceptor materials. ΔE3 is caused by non-radiative recombination at the D/A interface in OSCs, with this energy loss being exclusively determined by the device's EQEEL (electroluminescence quantum efficiency of the solar cell upon carrier injection under dark conditions). The value of ΔE3 typically ranges from 0.20 eV to 0.30 eV,46,47 and it is difficult to keep ΔE3 at a low level. As consistent with the above, ΔE3 has become the main limiting factor affecting the total Eloss. Consequently, reduction of non-radiative energy loss is essential for enhancing the power conversion efficiency in OSCs.
ln(EQEEL), stems from non-radiative recombination. The value of ΔE3 typically falls within 0.20–0.30 eV (ref. 50 and 51) and proves difficult to minimize. Consequently, ΔE3 constitutes the primary limiting factor governing Eloss and further impacting the PCE of OSCs through maximizing VOC (or minimizing voltage loss to the greatest extent) without sacrificing JSC and FF.52,53
To precisely characterize energy loss, the absorption spectrum of the active layer first corresponds to the photon energy absorbed during transitions from the ground state (S0) to the first excited state (S1),54,55 as depicted in Fig. 1C. The long-wavelength absorption edge (absorption cutoff) corresponds to the Eg, representing the minimum energy required to excite an electron from the valence band maximum to the conduction band minimum. The electroluminescence (EL) spectrum arises from photon emission during exciton recombination.56,57 In OSCs, it predominantly corresponds to radiative recombination of CT charges or excitons within the active layer, where the EL peak energy consistently lies below the material's Eg due to energy loss. Significantly, the energy at which the absorption and EL spectra intersect simultaneously satisfies the minimum energy for exciton generation (absorption cutoff) and the maximum energy for radiative recombination (high-energy EL emission limit).58,59 This intersection energy therefore equates to the material's Eg, quantifying the energy difference between the ground-state and excited-state excitonic configurations. Furthermore, Eloss and ΔE1, ΔE2, and ΔE3 can be determined according to the SQ limit (Fig. 1D).
The Eg of the active layer was determined from the intersection point between the device's EL spectrum and the absorption spectrum of the active layer film. Device Eloss is quantified using the VOC and Eg. Through computational deconvolution according to the SQ framework, Eloss partitions into three discrete components, enabling precise determination of ΔE1, ΔE2 and ΔE3 values. Following this method, Min et al.60 calculated Eg based on the intersection of the EL spectrum and absorption spectrum (Eg = 1240/λ), obtaining Eg values of 1.457 eV, 1.466 eV, and 1.492 eV for the PM6:PYTL, PM6:PYTM, and PM6:PYTH blends (Fig. 2A). Subsequently, using the Eg values, the Eloss for each corresponding device was calculated, yielding loss values of 0.532 eV, 0.528 eV, and 0.544 eV, respectively. For PM6:PYTL, ΔE1 = 0.258 eV, ΔE2 = 0.045 eV, ΔE3 = 0.229 eV; for PM6:PYTM, ΔE1 = 0.259 eV, ΔE2 = 0.058 eV, and ΔE3 = 0.210 eV; for PM6:PYTH, ΔE1 = 0.260 eV, ΔE2 = 0.062 eV, and ΔE3 = 0.222 eV (Fig. 2B). This comprehensive approach combining electroluminescence and absorption spectroscopy thus provides a quantitative method for evaluating and understanding the energy loss mechanisms in the device. The hole and electron mobilities (μh/μe) of the PM6:PYTL, PM6:PYTM, and PM6:PYTH blends were determined by SCLC measurements to be 2.81 × 10−4/7.88 × 10−4 cm2 V−1 s−1, 6.12 × 10−4/8.91 × 10−4 cm2 V−1 s−1, and 0.249 × 10−4/0.592 × 10−4 cm2 V−1 s−1(Fig. 2C), respectively. Among them, the PM6:PYTM-based device exhibits the highest and most balanced charge carrier mobility, which is advantageous for efficient charge transport.
![]() | ||
| Fig. 2 (A) Normalized absorption and EL spectra of PM6:PYTL, PM6:PYTM and PM6:PYTH blends; (B) semi-logarithmic plots display the normalized electroluminescence (EL), the measured external quantum efficiency (EQE), and the EQE derived from Fourier-transform photocurrent spectroscopy (EQEFTPS) as functions of photon energy for devices based on PM6:PYTL, PM6:PYTM, and PM6:PYTH; (C) the dark J–V characteristics of PM6:PYTL, PM6:PYTM, and PM6:PYTH based electron-only devices. The red lines represent the best fitting using the SCLC model.60 Copyright 2020 Elsevier. | ||
![]() | ||
| Fig. 3 (A) Chemical structure and energy level of PBD-Cl; (B) energy loss of devices based on PBD-ClL:N3, PBD-ClM:N3, and PBD-ClH:N3.69 Copyright 2024 WILEY-VCH. (C) Structure diagram of PFBCNTx and energy level of donor–acceptor materials; (D) energy loss of devices based on PBDB-TF and PFBCNTx.72 Copyright 2021 The Royal Society of Chemistry. (E) Chemical structures and energy levels of Qx-1 and Qx-2; (F) energy loss of binary OSC devices based on Qx-1 and Qx-2.73 Copyright 2022 Springer Nature. | ||
Molecular backbone modification through the introduction of electron-withdrawing groups enables alteration of molecular electron cloud distribution74,75 which can achieve optimized donor–acceptor alignment. This facilitates interfacial charge transfer while reducing charge transport energy loss. For example, Wei et al.73 synthesized acceptors Qx-1 and Qx-2 by fusing Qx derivatives into Y-series backbones (Fig. 3E). Replacing Y6's BTZ core with the more electron-withdrawing Qx core weakened intramolecular charge transfer (ICT), elevating the acceptor LUMO level from −4.05 eV (Y6) to −3.86 eV. According to space charge limited current (SCLC) measurements, the hole and electron mobilities (μh and μe) of the PM6:Qx-1 blend film were determined to be 3.64 × 10−4 cm2 V−1 s−1 and 1.88 × 10−4 cm2 V−1 s−1, respectively. In comparison, the PM6:Y6 blend film exhibited μh and μe values of 3.36 × 10−4 cm2 V−1 s−1 and 1.33 × 10−4 cm2 V−1 s−1, respectively. The PM6:Qx-1 blend demonstrates a more balanced charge carrier mobility compared to PM6:Y6, which is beneficial for enhanced charge transport properties. Consequently, the PM6:Qx-1 device exhibited reduced ΔE2 (0.066 eV to 0.031 eV), lower ΔE3 (0.233 eV to 0.214 eV), decreased Eloss (0.561 eV to 0.508 eV), increased VOC (0.859 V to 0.911 V), and higher PCE (16.6% to 17.9%) versus PM6:Y6 (Fig. 3F). Alex K.-Y. Jen et al.76 introduced selenium (Se) atoms into PYT-2S to prepare PYT-1S1Se and PYT-2Se (Fig. 4A). Increased Se content lowered acceptor LUMO levels. PYT-1S1Se exhibited a higher LUMO level (−3.75 eV) than PYT-2Se (−3.79 eV). The PM6:PYT-1S1Se device thus showed reduced ΔE2 (0.042 eV to 0.037 eV), lower Eloss (0.510 eV to 0.506 eV), increased VOC (0.908 V to 0.926 V), and improved PCE (15.5% to 16.3%) relative to PM6:PYT-2Se (Fig. 4B). Notably, Sun et al.77 further demonstrated this approach by introducing T-F units into PTQ10, yielding PTQ13-5 (Fig. 4C). The strong electron-withdrawing nature of fluorine atoms decreased the donor HOMO level from −5.51 eV to −5.54 eV. SCLC measurements were further conducted on PTQ13-5:K4 and PTQ10:K4 devices, revealing hole and electron mobilities (μh and μe) of 6.89 × 10−4/7.13 × 10−4 cm2 V−1 s−1 and 5.98 × 10−4/6.54 × 10−4 cm2 V−1 s−1, respectively. The corresponding μh/μe ratios were calculated to be 0.96 and 0.91. The PTQ13-5:K4-based device exhibits higher μh and μe values along with a more balanced charge transport ratio, leading to improved charge carrier extraction and overall device performance. The PTQ13-5:K4 device consequently demonstrated reduced ΔE3 (0.207 eV to 0.189 eV), lower Eloss (0.514 eV to 0.506 eV), increased VOC (0.888 V to 0.891 V), and higher PCE (17.61% to 18.42%) versus PTQ10:K4 (Fig. 4D).
![]() | ||
| Fig. 4 (A) Chemical structures and energy levels of PYT-1S1Se and PYT-2Se; (B) energy loss of devices based on PM6:PYT-1S1Se and PM6:PYT-2Se.76 Copyright 2021 Springer Nature. (C) Energy levels of PTQ10 and PTQ13-5; (D) energy loss of devices based on PTQ10:K4 and PTQ10:K4.77 Copyright 2025 American Chemical Society. (E) Chemical structures and energy levels of S-Alkyne-YF and T-Alkyne-Y; (F) energy loss of devices based on PM6:S-Alkyne-YF and PM6:T-Alkyne-Y.80 Copyright 2025 WILEY-VCH. (G) Chemical structures of PL1, PL2, PL3, and BTP-eC9-4F; (H) energy levels of PL1, PL2, PL3, and BTP-eC9-4F.81 Copyright 2022 WILEY-VCH. | ||
Optimization of molecular backbone geometry enhances structural regularity and enables precise energy level alignment in donor–acceptor systems.78,79 This modification promotes improved orbital overlap between adjacent molecules, facilitating the formation of continuous charge transport pathways. Consequently, carrier recombination during transport is suppressed, leading to minimized energy loss. For example, Li et al.80 developed alkyne-linked oligomeric electron acceptors (S-Alkyne-YF and T-Alkyne-YF) derived from M3 (Fig. 4E). Alkyne linkages, with their high bond dissociation energies and planar geometries, boost molecular planarity and promote ordered aggregation in thin films. Compared to S-Alkyne-YF, T-Alkyne-YF shows a higher peak (I0–0/I0–1) ratio, indicating stronger molecular aggregation and better packing order. It also exhibits a slightly raised LUMO energy level (−3.82 eV to −3.79 eV), widening the energy level offset for increased VOC and reduced energy loss. As shown in Fig. 4F, the PM6:T-Alkyne-YF device (vs. PM6:S-Alkyne-YF) shows lower ΔE3 (0.218 eV to 0.185 eV) and Eloss (0.536 eV to 0.516 eV), higher VOC (0.894 V to 0.943 V), and improved PCE (13.96% to 17.90%). Bo et al.81 synthesized polymer donor PL1 via random copolymerization of PM6 and D18 units (Fig. 4G). This reduced the HOMO level from −5.51 eV to −5.55 eV (Fig. 4H), reducing the energy level offset. SCLC measurements revealed hole and electron mobilities (μh and μe) of 8.87 × 10−4/7.45 × 10−4 cm2 V−1 s−1 for the PL1:BTP-eC9-4F device and 6.97 × 10−4/6.04 × 10−4 cm2 V−1 s−1 for the D18:BTP-eC9-4F device. The higher and more balanced charge carrier mobilities observed in the PL1-based blend contribute to enhanced charge transport properties. The PL1:BTP-eC9-4F device thus showed reduced ΔE3 (0.260 eV to 0.185 eV), lower Eloss (0.583 eV to 0.535 eV), increased VOC (0.827 V to 0.876 V), and improved PCE (16.18% to 18.14%) relative to D18:BTP-eC9-4F.
![]() | ||
| Fig. 5 (A) Chemical structures of PBDTSi-TZ and PBDTCl-TZ; (B) energy loss of devices based on PBDTSi-TZ:IT-4F and PBDTCl-TZ:IT-4F.86 Copyright 2020 WILEY-VCH. (C) Chemical structures of PE62 and F-Qx3b; (D) energy levels of PE62 and F-Qx3b.87 Copyright 2021 Elsevier. (E) Chemical structure of Y6-R and energy level of its device; (F) energy loss of devices based on PM6:Y6 and PM6:Y6-R.91 Copyright 2025 American Chemical Society. (G) Chemical structures and energy levels of PEH-F and THE-F; (H) energy loss of devices based on PTQ11:PEH-F and PTQ11:THE-F.92 Copyright 2025 Royal Society of Chemistry. | ||
Tailoring alkyl side chains effectively optimizes material energy levels and modulates donor–acceptor offsets.88–90 This molecular engineering approach facilitates elevated VOC, mitigated energy loss, and enhanced photovoltaic device performance. For instance, an asymmetric small-molecule acceptor, Y6-R, was developed by Zhou et al.91 (Fig. 5E). Its LUMO energy level increased from −4.09 eV to −3.97 eV compared to Y6. Relative to the PM6:Y6 device (Fig. 5F), the PM6:Y6-R counterpart demonstrated a reduction in ΔE3 (0.236 eV to 0.217 eV), lower Eloss (0.559 eV to 0.544 eV), higher VOC (0.839 V to 0.863 V), and an improved PCE (16.84% to 18.62%). Similarly, Sun et al.92 designed and synthesized two acceptors, PEH-F and TEH-F, with different conjugated outer side chains (Fig. 5G). The LUMO energy level of the acceptor increased from −3.92 eV to −3.85 eV. Compared with the PTQ11:TEH-F device (Fig. 5H), the PTQ11:PHE-F-based device showed a decrease in ΔE3 from 0.219 eV to 0.182 eV, an increase in VOC from 0.909 V to 0.936 V, and an improvement in PCE from 17.40% to 19.73%. Chen et al.93 employed an asymmetric design to develop acceptor BTP-BO-TBO, featuring alkyl/thienyl hybrid side chains. Its LUMO level increased from −3.99 eV to −3.91 eV compared to BTP-DTBO, enlarging the energy level offset and boosting VOC. The PM6:BTP-BO-TBO device showed reductions in ΔE2 (0.055 eV to 0.027 eV), ΔE3 (0.233 eV to 0.198 eV), and Eloss (0.550 eV to 0.485 eV), alongside increased VOC (0.878 V to 0.907 V) and PCE (17.22% to 19.76%).
![]() | ||
| Fig. 6 (A) Chemical structure and energy level of BPT-FT; (B) energy loss of devices based on PM6:BPT-4F and PM6:BPT-FT.96 Copyright 2025 WILEY-VCH. (C) Chemical structure and energy level of BTP-C9-ICX; (D) energy loss of devices based on PM6:Y5 and PM6:BTP-C9-ICT.97 Copyright 2022 Royal Society of Chemistry. (E) Chemical structure and energy level of CH17; (F) π–π stacking distances between molecular layers and different molecular packing modes of CH17 and Y6.98 Copyright 2022 Springer Nature. | ||
Strategic incorporation of a third component with strong D/A interactions facilitates the formation of an alloy phase103,104 that enables precise energy-level and ECT modulation reducing Eloss. For instance, the wide-bandgap donor polymer D18-Cl (higher-lying HOMO) was incorporated into the PM6:L8-BO blend (Fig. 7A) by Huang et al.105 An alloy phase formed with PM6, lowering the donor HOMO level (−5.57 eV vs. −5.53 eV). Consequently, VOC was increased and ECT was elevated (Fig. 7B), reducing the Eg − ECT offset. The ternary system exhibited decreased ΔE2 (0.07 eV vs. 0.08 eV), ΔE3 (0.207 eV vs. 0.227 eV), and Eloss (0.546 eV vs. 0.570 eV), alongside increased VOC (0.91 V vs. 0.87 V) and PCE (19.22%). Similarly, An et al.106 introduced the small molecule L8-BO as the third component into the B1:BO-4Cl system (Fig. 7C). L8-BO forms an alloy phase with BO-4Cl, increasing the LUMO energy level of the acceptor phase. According to −kT/q
ln(EQEEL), the ΔE3 of the ternary system decreases from 0.287 eV to 0.274 eV, and Eloss reduces from 0.541 eV to 0.530 eV as shown in Fig. 7D. Concurrently, VOC increases from 0.822 V to 0.841 V, with the device's PCE reaching 17.10%. Song et al.107 introduced BTP-BIO as the third component into the PM6:Y6 system (Fig. 7E). BTP-BIO forms an alloy phase with Y6, causing the acceptor LUMO energy level to increase from −4.10 eV to −4.01 eV. For the ternary device, ΔE3 decreases from 0.269 eV to 0.239 eV, Eloss reduces from 0.584 eV to 0.557 eV, VOC increases from 0.842 V to 0.848 V, and the device's PCE reaches 17.10%.
![]() | ||
| Fig. 7 (A) Energy level of the PM6:D18-Cl:L8-BO ternary system; (B) sEQE and EL spectra of PM6:L8-BO and PM6:D18-Cl:L8-BO devices.105 Copyright 2022 WILEY-VCH. (C) Chemical structure and energy level diagram of L8-BO; (D) EQEEL values of binary and ternary devices based on B1:BO-4Cl:L8-BO.106 Copyright 2023 American Chemical Society. (E) Chemical structure and energy level of Y6 and BTP-PIO.107 Copyright 2024 American Chemical Society. (F) Chemical structure and energy level of ZY-4Cl and BTP-eC9.110 Copyright 2022 Royal Society of Chemistry. | ||
The cascade energy-level structure can also be constructed by the appropriate incorporation of a third component with tailored energy levels,108,109 which is able to enhance charge-transfer efficiency through optimizing the driving force while suppressing bimolecular recombination to collectively reduce energy loss. For example, Sun et al.110 introduced ZY-4Cl as the third component into the PM6:BTP-eC9 system (Fig. 7F), forming a cascaded energy level structure where charge transfer occurs between the donor and the two acceptors, suppressing bimolecular recombination. Meanwhile, it increases the LUMO energy level of the acceptor, enlarging the energy level offset for enhanced VOC. Compared with the PM6:BTP-eC9 device, the PM6:ZY-4Cl:BTP-eC9-based device exhibits decreased ΔE2 from 0.084 eV to 0.079 eV, reduced ΔE3 from 0.214 eV to 0.198 eV, and decreased Eloss from 0.532 eV to 0.520 eV, with the VOC increasing from 0.844 V to 0.863 V and PCE improving from 17.72% to 18.69%. Similarly, Peng et al.111 introduced IDIC as the third component into the PM6:TOBDT system, where the ternary blend forms a cascade energy level structure, promoting charge transfer while lowering the HOMO energy level of the donor and increasing the VOC. Compared with the PM6:TOBDT device, the PM6:IDIC:TOBDT-based device shows a decrease in ΔE3 from 0.249 eV to 0.235 eV, with the VOC increasing from 0.88 V to 0.91 V and PCE improving from 11.0% to 13.7%.
In binary organic systems, intermolecular interactions are modulated through solid additives, enabling active layers with optimized domain size.54,118 This controlled morphology enhances charge transport while suppressing bimolecular recombination, reducing Eloss. For example, DIMCH was added to PM6:L8-BO binary system by Yang et al.119 DIMCH exhibits concentrated electrostatic potential (Fig. 8A), demonstrating strong attraction with PM6 and repulsion with L8-BO. Consequently, PM6 crystallinity was enhanced while L8-BO aggregation was suppressed (Fig. 8B), balancing donor–acceptor crystallinity and reducing phase separation. According to Fig. 8C, grazing incidence small-angle X-ray scattering (GISAXS) analysis revealed smaller domain sizes (17.2 nm vs. 23.2 nm with DIO). The optimized morphology suppressed non-radiative recombination, reducing ΔE2 (0.082 eV vs. 0.087 eV) and ΔE3 (0.209 eV vs. 0.229 eV) (Fig. 8D). Ultimately, VOC increased (0.906 V vs. 0.866 V) with an improved PCE of 18.8%. Zhang et al.120 added 1,8,9-trihydroxyanthracene (TOHA) as a solid additive to the D18-Cl:N3 binary blend system. According to time-of-flight secondary ion mass spectrometry (ToF-SIMS) measurements, phase separation was characterized by comparing changes in the content of characteristic elements. Compared with films without additives, TOHA inhibits the diffusion of D18-Cl and N3, resulting in a uniform phase distribution in the film, thereby improving the phase separation structure of the active layer and reducing energy loss (Fig. 8E). For a device with TOHA addition, ΔE3 decreases from 0.152 eV to 0.148 eV, and Eloss reduces from 0.613 eV to 0.608 eV. Meanwhile, VOC increases from 0.854 V to 0.862 V, leading to a device PCE of 17.91%.
![]() | ||
| Fig. 8 (A) GIWAXS patterns of PM6 or L8-BO with addition, and pole figure of PM6 and L8-BO; (B) vertical distribution maps of donors and acceptors in BHJ films with addition; (C) GISAXS patterns of BHJ films with DIO or DIMCH addition; (D) energy loss of devices with addition.119 Copyright 2024 WILEY-VCH. (E) Energy loss of devices treated with different additives.120 Copyright 2023 WILEY-VCH. | ||
Introducing polymers as the third component into the system can improve film-forming kinetic behavior and regulate phase separation.121,122 For example, the polymer acceptor PYIT, which exhibits a narrower bandgap than Y6, was introduced into the PM6:Y6 binary system by Han et al.,123 forming an optimized ternary system. PYIT crystallizes before Y6, shortening Y6 crystallization time (Fig. 9A). A blue-shifted absorption spectrum was observed in Y6:PYIT blend films relative to Y6 films, indicating suppressed aggregation and reduced acceptor domain size. Absorption spectroscopy (Fig. 9B) revealed a blue shift of the absorption maximum to 810 nm compared to 815 nm in the PM6:Y6 blend, indicating reduced Y6 aggregation. During film formation, a bicontinuous network was formed between PYIT and Y6, promoting exciton dissociation and suppressing bimolecular recombination. Consequently, device Eloss was reduced (0.539 eV vs. 0.559 eV) and VOC increased (0.864 V vs. 0.844 V) (Fig. 9C). Identically, the ternary system achieved 17.05% PCE versus 15.40% for PM6:Y6. PC71BM was incorporated by Hou et al.124 into the IDIC:DRTB-T binary blend system. Strong acceptor miscibility inhibited IDIC aggregation, reducing phase separation from 81.4 nm to 15.5 nm. This enhanced exciton dissociation, reduced energy loss, and improved JSC, FF, and PCE. Similarly, domain size within the D18:AQx-2F system was precisely modulated through eC9 introduction, optimizing exciton dissociation and charge transport, as demonstrated by Zhu et al.17 (Fig. 9D), increasing the domain size from 13.4 nm to 20.0 nm, promoting charge transport, suppressing bimolecular recombination, and reducing energy loss (Fig. 9E). Compared with the D18:AQx-2F device, the ΔE3 of the D18:AQx-2F:eC9-based device decreases from 0.194 eV to 0.180 eV, Eloss reduces from 0.533 eV to 0.510 eV, and PCE improves from 19.2% to 20.6%.
![]() | ||
| Fig. 9 (A) Absorption peak position of Y6 as a function of time; (B) absorption spectra of binary and ternary films; (C) energy loss of PM6, Y6, and PYIT binary and ternary devices.123 Copyright 2024 Royal Society of Chemistry. (D) The GISAXS patterns of D18:AQx-2F, D18:eC9, and D18:AQx-2F:eC9 blend films; (E) energy loss of D18, AQx-2F, and eC9 binary and ternary devices.17 Copyright 2025 WILEY-VCH. (F) Film formation process schematic diagram of the LBL device PBQx-TF/DT-Y6:ZY-4C; (G) contact angle images of DT-Y6, ZY-4Cl, and DT-Y6:ZY-4Cl; (H) energy loss of PBQx-TF device.126 Copyright 2024 Royal Society of Chemistry. (I) Schematic diagram of active layer formation for PPHJ structures with different electrostatic potentials; (J) energy loss of devices with different systems.127 Copyright 2024 WILEY-VCH. | ||
Vertical phase separation occurs within the active layer, where the donor concentrates near the anode and the acceptor enriches near the cathode,11,125 which suppresses bimolecular recombination during carrier transport and reduces overall device Eloss. For example, Huang et al.126 incorporated ZY-4Cl with a higher LUMO energy level, into the PBQx-TF/DT-Y6 system. Contact angle measurements during spin-coating (Fig. 9F) showed that ZY-4Cl and DT-Y6 exhibit similar γ values (21.24 mN m−1 and 21.8 mN m−1, respectively; Fig. 9G), indicating good miscibility between the donor and acceptor materials. However, the γ value of the ZY-4Cl/DT-Y6 mixture (21.44 mN m−1) became significantly higher than that of the donor PBQx-TF (18.8 mN m−1). This increases the surface energy difference and promotes an optimal vertical phase separation structure. Consequently, bimolecular recombination is suppressed, enhancing carrier lifetime and charge mobility. As a result, the devices incorporating PBQx-TF/DT-Y6:ZY-4Cl show increased VOC (0.881 V vs. 0.840 V), reduced Eloss (0.55 eV vs. 0.601 eV), and a higher PCE of 19.46% compared to the binary counterpart (Fig. 9H). Similarly, Chen et al.127 introduced BTP-BO as the third component into the PBDB-TF/BTP-BO-4CI binary system to optimize the vertical phase separation morphology of the active layer in the binary system. This increases the degree of vertical phase separation in the active layer (Fig. 9I), effectively reducing Eloss. Energy loss results in Fig. 9J show that the ΔE3 of the ternary system decreases from 0.323 eV to 0.219 eV, while the VOC increases from 0.748 V to 0.866 V, with the PCE of the ternary device reaching as high as 19.09%.
Additives exhibiting strong intermolecular interactions with donor/acceptor materials regulate material crystal growth,132,133 achieving optimized morphology and reduced energy loss. For example, TMB was introduced into the PBDB-TF:BTP-eC9 binary system by Hou et al.114 Opposing electrostatic potentials between TMB and BTP-eC9 promoted strong π–π interactions, which inhibited excessive aggregation while optimizing π–π stacking (Fig. 10A). In the meantime it increased the coherence length (26.96 Å vs. 23.96 Å) and enhanced JSC. TMB volatilization prior to film solidification yielded optimized crystallinity (Fig. 10B). Bimolecular recombination was suppressed by the resulting fibrous network, reducing ΔE2 (0.296 eV vs. 0.310 eV), ΔE3 (0.207 eV vs. 0.228 eV), and Eloss (0.541 eV vs. 0.563 eV) versus the DIO-treated device. Concurrently, VOC increased (0.854 V vs. 0.837 V) with 18.61% PCE. Li et al.134 introduced TTBB and TTBI into the PM6/L8-BO binary system, where TTBB enhanced intermolecular interactions with PM6, promoting crystal growth and improving crystallinity. TTBI enabled tighter molecular packing and more ordered arrangement of L8-BO, resulting in a highly crystalline morphological structure in the active layer. Compared with the PM6/L8-BO(DIO) device, the PM6(TTBB)/L8-BO(TTBI) devices showed a decrease in ΔE3 from 0.24 eV to 0.23 eV, a reduction in Eloss from 0.56 eV to 0.54 eV, and an increase in VOC from 0.881 V to 0.898 V, with the device PCE reaching 19.4%. Furthermore, 1,2-dimethylnaphthalene (DMN) was added to Y6 (o-XY) solution by Xie et al.135 Dual-phase DMN-induced nucleation generated high-crystallinity phase separation in PM6/Y6 (Fig. 10C). This yielded decreased Eloss (0.515 eV vs. 0.589 eV), increased VOC (0.849 V vs. 0.764 V), and improved PCE (16.15% vs. 12.76%).
![]() | ||
| Fig. 10 (A) Schematic diagram of the mechanism of action of the TMB additive during the film formation process; (B) 2D GIWAXS patterns and their scattering curves of PBDB-TF:eC9 films with DIO or TMB addition.114 Copyright 2023 Royal Society of Chemistry. (C) AFM height images and phase images of PM6/Y6(o-XY) and PM6/Y6(o-XY + DMN) films.135 Copyright 2019 Royal Society of Chemistry. (D) GIWAXS patterns and the scattering curves of PM6:Y6 and PM6:BTTzR:Y6 films.140 Copyright 2021 Elsevier. (E) GIWAXS patterns and the scattering curves of PM6:L8-BO and PM6:L8-BO:DY-TF films.141 Copyright 2023 WILEY-VCH. | ||
Post-treatment strategies optimize active layer morphology by regulating crystallization behavior.71,136 During spin-coating, rapid solvent evaporation traps the active layer morphology in a kinetically quenched, thermodynamically metastable state. This metastability is exacerbated by structural similarities between donors and acceptors, which promote poorly crystalline interpenetrating phases. Post-treatment can enhance molecular mobility, enabling redistribution into thermodynamically favorable configurations. Consequently, bimolecular recombination is suppressed and energy loss reduced. For example, solvent vapor annealing (SVA) was applied to the BTTzR:Y6 device by Li et al.137 GIWAXS analysis revealed increased coherence lengths (CCL(100): 7.14 nm vs. 5.49 nm; CCL(010): 2.61 nm vs. 1.87 nm) and reduced π–π stacking distances (3.50 Å vs. 3.53 Å). The crystalline bicontinuous network reduced ΔE3 to 0.18 eV and Eloss to 0.51 eV, improving PCE from 0.1% to 13.9%. Similarly, sequential TA/SVA treatment was employed on the BSCl:IDIC-4Cl binary blend system by Wei et al.,138 enhancing OOP crystallinity (CCL(010): 13.8 nm vs. 13.5 nm), yielding reduced ΔE3 (0.373 eV vs. 0.394 eV) and Eloss (0.721 eV vs. 0.745 eV), and in the meantime, increasing VOC (0.865 V vs. 0.845 V) and 13.03% PCE, which is a 23% improvement compared to those of the original device. Ge et al.121 optimized the active layer morphology of OSCs based on the BT-2F:N3 system through CS2 SVA, improving the crystallinity of the active layer. Compared with the TA-treated device, the ΔE3 of the CS2 SVA-treated device decreased from 0.244 eV to 0.226 eV, Eloss reduced from 0.529 eV to 0.521 eV, VOC increased from 0.829 V to 0.845 V, and the PCE reached 15.39%, which was significantly higher than the 14.66% of the TA-treated device.
Introducing a third component as nuclei to promote heterogeneous nucleation of molecules in the active layer is an effective approach to improve crystallization behavior.81,139 For example, Bo et al.139 introduced LA15 as the third component into the D18:L8-BO binary blend system. Its branched alkyl side chains improved molecular planarity and intermolecular interactions, yielding high crystallinity. Acting as nuclei, LA15 induced crystallization of donor polymer D18 while increasing L8-BO coherence length (CCL(010): 20.94 Å vs. 16.58 Å) via GIWAXS. The ternary film formed an interconnected network, establishing efficient electron transport channels that suppressed recombination. Consequently, EQEEL increased (0.063% vs. 0.036%), VOC increased (0.92 V vs. 0.90 V), and ΔE3 decreased (0.18 eV vs. 0.20 eV), achieving 19.13% PCE. Similarly, the highly crystalline small-molecule donor BTTzR was introduced into the PM6:Y6 binary system by Li et al.,140 enhancing active layer crystallinity (Fig. 10D). The ternary device exhibited reduced ΔE3 (0.19 eV vs. 0.23 eV), lower Eloss (0.53 eV vs. 0.57 eV), increased VOC (0.87 V vs. 0.83 V), and 17.7% PCE versus 15.8%. Furthermore, Huang et al.141 introduced DY-FT into the PM6:L8-BO system; DY-FT has a similar main chain structure and crystal structure to L8-BO, increasing the crystallinity of the acceptor phase (Fig. 10E), which suppresses bimolecular recombination and reduces energy loss. Relative to the binary device, the ternary device showed a decrease in ΔE3 from 0.247 eV to 0.228 eV, a reduction in Eloss from 0.591 eV to 0.565 eV, an increase in VOC from 0.884 V to 0.905 V, and an improvement in PCE from 18.44% to 19.13%.
Additive molecules thermodynamically optimize active layer morphology through distinct donor–acceptor interactions,31,144 inducing molecular reorientation to reduce energy loss and enhance performance. For instance, DBOF was added to the D18:Y6 binary system by Peng et al.145 Strong π–π interactions with Y6 promoted acceptor rearrangement during volatilization, optimizing vertical donor–acceptor distribution. This yielded the narrowest out-of-plane (010) peak of Y6 films, indicating enhanced face-on orientation and ordered packing (Fig. 11A). The resulting nanoscale interpenetrating network reduced ΔE3 (0.230 eV vs. 0.250 eV) and Eloss (0.533 eV vs. 0.553 eV), while increasing VOC (0.875 V vs. 0.869 V) and achieving 18.60% PCE. Zhan et al.146 first utilized CsPb3 perovskite quantum dots (PQDs) as additives. In PTB7-Th:FOIC films treated with PQDs, molecules preferentially adopt a face-on orientation (Fig. 11B), which facilitates charge transport. Finally, for OSCs based on the PTB7-Th:FOIC system treated with PQDs, ΔE2 decreases from 0.120 eV to 0.010 eV, ΔE3 reduces from 0.311 eV to 0.295 eV, and Eloss decreases from 0.607 eV to 0.580 eV. Meanwhile, VOC increases from 0.711 V to 0.740 V, thus improving the PCE of the PTB7-Th:FOIC-based device from 11.6% to 13.2%. Furthermore, the 3D star-shaped Star-A was introduced into the PM6:Y6 binary blend system by Dae Hwan Lee et al.147 Star-A disrupted intrinsic molecular orientations (edge-on PM6 and face-on Y6), inducing preferential face-on alignment as shown in Fig. 11C. This enhanced exciton dissociation and reduced energy loss, yielding decreased ΔE3 (0.224 eV vs. 0.240 eV), lower Eloss (0.563 eV vs. 0.581 eV), increased VOC (0.856 V vs. 0.835 V), and improved PCE (17.2% vs. 15.6%).
![]() | ||
| Fig. 11 (A) 2D GIWAXS patterns, their scattering curves, and azimuthal dependence curves around the (010) peak in the 2D GIWAXS patterns of D18:Y6 films without an additive and with DBOF addition.145 Copyright 2024 Elsevier. (B) 2D GIWAXS patterns and their scattering curves of PTB7-Th:FolC and PTB7-Th:FolC:PQDs films.146 Copyright 2020 WILEY-VCH. (C) 2D GIWAXS patterns, FWHM, crystalline coherence length plots of the (010) peak, and azimuthal scan plots of the (100) peak in GIWAXS modes of PM6:Y6 films without an additive and with Star-A addition.147 Copyright 2022 Elsevier. | ||
Thermal annealing enables precise regulation of molecular packing within the active layer,148–150 which mitigates energy loss and consequently enhances device performance. For example, Zhang et al.151 adopted a rapid thermal annealing (RTA) strategy (Fig. 12A), to optimize the PM6:L8-BO-X blend film, resulting in favorable face-on orientations. GIWAXS analysis revealed enhanced face-on orientations versus thermal annealing (TA), attributed to slower solvent evaporation enabling molecular reorientation. The minimized in-plane (010) peak indicated superior face-on stacking. Optimized D/A morphology with increased interfacial contact promotes exciton dissociation. Consequently, the RTA-treated device (Fig. 12B) exhibited reduced ΔE2 (0.070 eV vs. 0.076 eV), ΔE3 (0.215 eV vs. 0.227 eV), and Eloss (0.549 eV vs. 0.567 eV), alongside increased VOC (0.896 V vs. 0.877 V) and 19.91% PCE versus 18.89%. Zhong et al.152 employed a fluorinated solvent vapor annealing (FSVA) strategy to optimize the PF7:L8BO blend film, leading most of the molecules in the active layer to adopt a face-on orientation. Compared with the TA-treated device, the FSVA-treated device shows a decrease in ΔE3 from 0.228 eV to 0.225 eV, a reduction in Eloss from 0.554 eV to 0.550 eV, and an improvement in device PCE from 18.68% to 19.09%. Furthermore, SVA was implemented on the PBDS-T:BTP-eC9 binary system by René A. J. Janssen et al.153 GIWAXS showed decreased in-plane (010) and increased out-of-plane (010) peaks, confirming the dominant face-on orientation versus unannealed films (Fig. 12C). The SVA-treated device achieved 16.4% PCE (vs. 14.7%) with reduced energy loss, with a ΔE1 of 0.270 eV, ΔE3 of 0.035 eV, ΔE3 of 0.226 eV, and Eloss of 0.531 eV.
![]() | ||
| Fig. 12 (A) Schematic diagrams of the thermal annealing (TA) process and RTA process, and GIWAXS patterns of PM6:L8-BO-X films treated with TA or RTA; (B) energy loss of the PM6:L8-BO-X device treated with TA or RTA.151 Copyright 2024 WILEY-VCH. (C) 2D GIWAXS patterns and their scattering curves of active layers without annealing or treatment with SVA.152 Copyright 2022 Elsevier. (D) GIWAXS patterns and their scattering curves of DRCN5T:ITIC-Th and DRCN5T:ITIC-Th:N2200 films.155 Copyright 2022 WILEY-VCH. (E) 2D GIWAXS patterns of PBDB-T:HF-PCIC and PBDB-T:HF-PCIC:IEICO-4F films.157 Copyright 2018 WILEY-VCH. (F) 2D GIWAXS patterns of PBDB-T:NNBDT and PBDB-T:NNBDT:FDNCTF films.158 Copyright 2018 WILEY-VCH. | ||
Introducing a third component induces the transformation of donors or acceptors from the edge-on to the face-on orientation.154,155 For example, the highly crystalline N2200 with a face-on orientation was incorporated into the DRCN5T:ITIC-Th binary system by Liu et al.156 High lattice matching (93%) with DRCN5T and limited miscibility enabled N2200 to act as nucleation sites. Elevated temperature enhanced DRCN5T diffusion sufficiently to overcome heterogeneous nucleation barriers, inducing the face-on DRCN5T orientation (Fig. 12D). The uniform face-on ternary active layer reduced Eloss (0.74 eV vs. 0.79 eV), increased VOC (0.99 V vs. 0.94 V), and achieved 54% relative PCE enhancement. Chen et al.157 introduced IEICO-4F into the PBDB-T:HF-PCIC system. The absorption peak in the in-plane (010) direction decreases, while that in the out-of-plane (010) direction increases, indicating that the molecular orientation in the ternary blend film is mainly face-on (Fig. 12E), which promotes exciton dissociation and reduces energy loss. Compared with the binary device, the ternary device exhibits a decrease in ΔE2 from 0.13 eV to 0.04 eV, a reduction in ΔE3 from 0.38 eV to 0.28 eV, a decrease in Eloss from 0.80 eV to 0.59 eV, and an improvement in device PCE from 8.82% to 11.20%. Chen et al.158 introduced FDNCTF into the PBDB-T:NNBDT binary system. In the binary blend film, NNBDT is mainly oriented in a face-on manner. FDNCTF has a similar chemical structure and molecular orientation to NNBDT, leading to more molecules in the ternary film transforming into a face-on orientation159 (Fig. 12F), further reducing energy loss. Compared with the binary device, the ternary device shows a decrease in Eloss from 0.55 eV to 0.54 eV, an increase in VOC from 0.880 V to 0.887 V, and an improvement in device PCE from 11.7% to 12.8%.
In large-area devices, the typically thicker active layers result in longer charge transport paths.166,167 This increased distance inherently increases the probability of charge carrier recombination, leading to greater energy loss. For example, Ge et al.168 developed a chlorinated analogue, eC9-2Cl, based on the BTP-eC9 acceptor, where the incorporation of chlorine atoms enhances the molecular dipole moment, promoting tighter and more ordered intermolecular packing. To further facilitate large-area module fabrication, the active-layer thickness was reduced, shortening charge transport pathways and suppressing bimolecular recombination, thereby lowering energy loss. Compared to the PCE10:BTP-eC9 large-area module (0.09 cm2), the PCE10:eC9-2Cl-based module exhibits reduced energy loss, with ΔE2 decreasing from 0.103 eV to 0.080 eV, ΔE3 from 0.311 eV to 0.264 eV, and total Eloss from 0.677 eV to 0.628 eV. The optimized module delivers a VOC of 5.318 V, a JSC of 2.60 mA cm−2, an FF of 68.15%, and a PCE of 9.42%.
Imperfect interfacial contact in large-area modules impedes charge extraction, leading to localized recombination and increased non-radiative energy loss.169 For example, Zheng et al.170 developed a self-assembled monolayer (SAM), termed 4BCP, which features an asymmetric even-carbon alkyl chain that imparts high solubility and an enhanced dipole moment. This molecular design enables uniform and dense coverage on indium tin oxide (ITO) electrodes, effectively minimizing contact defects at the interface with the active layer. The optimized contact facilitates efficient charge collection by the electrode, suppresses localized recombination, and reduces energy loss. As a result, the PM6:BTP-eC9-based device incorporating 4BCP achieved a PCE of 19.7%. Moreover, a large-area device (1.10 cm2) with the same configuration attained a PCE of 17.2%, with a VOC of 0.876 V, a JSC of 26.4 mA cm−2, and an FF of 74.4%.
Enhancing device stability can be achieved by suppressing the growth of phase-separated domains, as excessive domain size impedes charge transport and promotes recombination, which are primary drivers of performance degradation.173 For instance, Li et al.27 developed an asymmetric acceptor, BTP-eC9-4ClO, based on the BTP-eC9 structure. GISAXS measurements revealed a reduction in the 2Rg from 31.2 nm for the PM6:BTP-eC9 blend to 21.1 nm for the PM6:BTP-eC9-4ClO blend, indicating finer phase separation in the latter. This suppressed phase aggregation facilitates efficient charge transport, reduces bimolecular recombination, and lowers energy loss, while simultaneously enhancing device stability. Compared to the PM6:BTP-eC9-based device, the PM6:BTP-eC9-4ClO-based device exhibited a decrease in ΔE3 from 0.202 eV to 0.179 eV and a reduction in total Eloss from 0.555 eV to 0.530 eV, accompanied by an improvement in VOC from 0.861 V to 0.891 V and an increase in PCE from 19.12% to 20.03%. The optimized device also demonstrated exceptional thermal stability, with a T80 lifetime of 7834 hours under heating at 80 °C in a nitrogen atmosphere-significantly exceeding the 395 hours observed for the control device.
Suppressing the detrimental molecular reorientation enhances operational stability by preserving efficient charge transport pathways and minimizing degradation from recombination.174 For example, Chen et al.175 introduced a gradient thermal annealing strategy to regulate the molecular orientation and crystallinity of the polymer donor PM6. This process promoted a transition toward face-on molecular stacking in the bulk-heterojunction blend, facilitating charge transport and reducing energy loss while enhancing device stability. In the PM6:BO-4Cl system, the optimized crystallization behavior of PM6 suppressed its excessive diffusion into the acceptor phase, thereby preserving a well-interpenetrated donor–acceptor network. Compared to the control device, the gradient-annealed device exhibited a reduction in ΔE2 from 0.061 eV to 0.056 eV, a decrease in ΔE3 from 0.219 eV to 0.212 eV, and a lowering of total Eloss from 0.541 eV to 0.528 eV, accompanied by an increase in VOC from 0.847 V to 0.857 V and an improvement in PCE from 17.25% to 18.22%. Furthermore, the treated device demonstrated superior operational stability, retaining 91.6% of its initial PCE after 300 hours of continuous illumination under a nitrogen atmosphere, outperforming the control device which maintained 86.1%.
Current optimization strategies predominantly address energy level modulation and morphological control as separate domains, whereas the reciprocal interaction mechanisms between these two regulatory dimensions remain inadequately elucidated. For example, subtle modifications of molecular energy levels may affect crystallization processes through altered intermolecular interactions, while morphological evolution may in turn influence frontier orbital energetics, as exemplified by π–π stacking modulation of HOMO and LUMO energy levels. Future research must establish a dynamic “energy level-morphology-performance” correlation model integrated with in situ characterization techniques to elucidate real-time coupled evolution of electronic and structural properties during material synthesis and film formation, enabling precise synergistic control of multiple parameters.
Despite being achieved primarily through elaborate molecular modifications or complex post-treatments, the long-term stability and compatibility with large-scale manufacturing processes of existing high-performance systems remain significant limitations. On one hand, it is necessary to develop donor–acceptor materials with intrinsic high stability; for instance, strategies such as rigid backbone design and side-chain cross-linking can be employed to suppress molecular aggregation-induced degradation. On the other hand, exploration of “self-assembling” active layer systems should be conducted, utilizing specific intermolecular interactions (e.g., hydrogen bonds and halogen bonds) to achieve ideal morphologies without complex post-treatments, thus balancing efficiency and fabrication cost. Meanwhile, stabilization of ternary and multicomponent systems requires rigorous characterization of long-term phase separation dynamics among constituents, along with the development of thermodynamically stable alloy-like phases.
Furthermore, a significant gap persists between the practical energy loss in OSCs and the theoretical limit, with ΔE3 accounting for over 30% of total loss representing a major obstacle toward reaching the S–Q limit. Current understanding remains largely macroscopic. Future efforts should integrate quantum chemical calculations with single-molecule spectroscopy to establish a comprehensive model linking “molecular packing–defects–non-radiative decay”, and employ techniques such as transient absorption spectroscopy to elucidate energy dissipation pathways. Concurrently, tackling challenges including defect modulation and energy level alignment will be essential to reduce non-radiative loss below 100 meV, thereby enabling the synergy of high VOC and high JSC.
To achieve higher PCE, future research should focus on synergistically minimizing energy loss while enhancing photon harvesting and charge collection capabilities. Key priorities include developing novel narrow-bandgap materials to broaden spectral response, employing multilayer architectures or photon management techniques to improve light absorption, and incorporating dipole-interfacial layers to optimize energy-level alignment and suppress non-radiative recombination. Advanced in situ characterization methods are needed to resolve film formation dynamics and morphological evolution, enabling precise control over active layer aggregation. Furthermore, integrating large-area fabrication processes with stability studies is essential to maintain high performance under real-world operating conditions. A systematic approach combining material design, structural optimization, and process innovation will be crucial to push the PCE of organic solar cells toward the theoretical limit of 25% and beyond, laying a solid foundation for their commercial application.
| This journal is © The Royal Society of Chemistry 2025 |