Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

nature energy

Article https://doi.org/10.1038/s41560-024-01557-z

Non-fullerene acceptor with asymmetric


structure and phenyl-substituted alkyl side
chain for 20.2% efficiency organic solar cells

Received: 28 June 2023 Yuanyuan Jiang1,2,4, Shaoming Sun1,2,4, Renjie Xu1,2,4, Feng Liu 1
,
Xiaodan Miao1,2, Guangliu Ran 3, Kerui Liu1,2, Yuanping Yi 1,
Accepted: 20 May 2024
Wenkai Zhang 3 & Xiaozhang Zhu 1,2
Published online: xx xx xxxx

Check for updates For organic solar cells (OSCs), bridging the gap with Shockley–Queisser
limit necessitates simultaneously reducing the energy loss for a high
open-circuit voltage, improving light utilization for enhanced short-circuit
current density and maintaining ideal nanomorphology with a high fill
factor through molecular design and device engineering. Here we design
and synthesize an asymmetric non-fullerene acceptor (Z8) featuring
tethered phenyl groups to establish an alloy acceptor in ternary OSCs. The
asymmetric structure minimizes non-radiative energy loss and charge
recombination owing to delocalized excitons. The phenyl-substituted alkyl
side chain impacts on the intermolecular interactions, improving the f­i l­m
n­an­omorphology with efficient exciton dissociation and reduced charge
recombination. We demonstrate OSCs with an efficiency of 20.2% (certified
19.8%) based on the D18:Z8:L8-BO ternary blend. Through theoretical
calculations, we examine the overall distribution of photon and carrier
losses and analyse the potential for improvement on open-circuit voltage,
short-circuit current density and fill factor, providing rational guidance for
further development of the OSC performance.

The power conversion efficiency (PCE) of organic solar cells (OSCs) simultaneous improvements in VOC, JSC and FF. NFA-based OSCs usually
is mainly determined by parameters: open-circuit voltage (VOC), show relatively large voltage loss in the range of 0.50–0.80 V, which is
short-circuit current density (JSC) and fill factor (FF), following far beyond the ideal value of 0.25–0.30 V given by the Shockley–Queis-
PCE = VOC × JSC × FF/Pin, where Pin is the incident light intensity. Over ser (S-Q) theory23, ascribed to the radiative and non-radiative energy
the past decades, great efforts have been consistently devoted to losses below bandgap. Recent studies indicate the A-DA’D-A-type of
improving the PCE of OSCs, including the innovation of new polymer Y-analogues can form the three-dimensional charge transport chan-
donors1–4, electron acceptors5–10 and device engineering11–17. The mile- nels with low energetic disorder24–26, enabling the hybridization of
stone achievements can be attributed to the rapid development of localized exciton (LE) and charge transfer (CT) states to mitigate the
novel fused-ring non-fullerene electron acceptors (NFAs) since 2015 non-radiative recombination loss27,28. Although voltage loss has long
(refs. 6,10,17–22). Currently, achieving highly efficient single-junction been considered as a limiting factor for OSCs, the current loss is an
OSCs with PCE exceeding 20% remains challenging, which necessitates equally critical problem that requires immediate attention29. The loss

Beijing National Laboratory for Molecular Sciences, CAS Key Laboratory for Organic Solids, Institute of Chemistry, Chinese Academy of Sciences,
1

Beijing, China. 2School of Chemical Sciences, University of Chinese Academy of Sciences, Beijing, China. 3Department of Physics and Applied Optics
Beijing Area Major Laboratory, Center for Advanced Quantum Studies, Beijing Normal University, Beijing, China. 4These authors contributed equally:
Yuanyuan Jiang, Shaoming Sun, Renjie Xu. e-mail: liufeng112@iccas.ac.cn; xzzhu@iccas.ac.cn

Nature Energy
Article https://doi.org/10.1038/s41560-024-01557-z

of incident photons is derived from insufficient internal quantum effi- Z8 and 2,2′-((2Z,2′Z)-((3,9-bis(2-butyloctyl)-12,13-bis(2-ethylhexyl)-
ciency (IQE), parasitic absorption of interlayers and electrodes, and the 12,13-dihydro-[1,2,5]thiadiazolo[3,4-e]thieno-[2′,3′: 4′,5′]
reflection from the device. For state-of-the-art OSCs, the charge genera- thieno[2′,3′:4,5]pyrrolo[3,2-g]thieno[2′,3′:4,5]thieno[3,2-b]
tion efficiency is mainly susceptible to the highest occupied molecular indole-2,10-diyl)bis(methaneyl-ylidene))bis(5,6-difluoro-3-oxo-
orbital (HOMO) energy offset (ΔEHOMO), and achieving efficient charge 2,3-dihydro-1H-indene-2,1-diylidene))dimalononitrile (L8-BO) are
generation with maximized JSC under minimal ΔEHOMO value has been located at 761, 746 and 731 nm in dilute chloroform solution, and 832,
considered as a practical approach to realizing high-performance 814 and 796 nm in neat films, respectively. Compared with L8-BO, Z7
OSCs30. FF is another key factor affecting the PCE of photovoltaics. and Z8 deliver red-shifted absorption spectrum, which can broaden
The non-ideal FF is, complicatedly, subject to the relatively low carrier the light-harvesting window for higher JSC in photovoltaic devices. The
mobility and affected by the blend nanomorphology, suffering inter- energy levels of the investigated materials were checked by ultravio-
facial and bulk charge recombination31,32. To simultaneously improve let photoelectron spectra and inverse photoemission spectroscopy
the VOC, JSC and FF of single-junction OSCs, constructing highly efficient measurements, and the relevant data are summarized in Supplemen-
ternary organic solar cells (TOSCs) by introducing a third component tary Table 1. The HOMO and the lowest unoccupied molecular orbital
into the binary host system is a promising strategy11,33–36. Blending (LUMO) energy levels are −5.60 and −3.96 eV for Z7, −5.68 and −3.91 eV
multi-component active materials produces a broader absorption for Z8, and −5.72 and −3.84 eV for L8-BO, respectively. For the blend
window, resulting in desired light harvesting. Moreover, the guest acceptors Z7:L8-BO and Z8:L8-BO, it is expected that they will exhibit
component can fine-tune the electronic properties and affect the nano- middling energy levels between these of L8-BO and Z-type acceptors,
morphology of the blend films, resulting in efficient charge generation considering the formation of alloy acceptors. From Supplementary
and transport, even with reduced voltage loss37–40. However, the PCE of Figs. 6 and 7, it can be observed that Z7:L8-BO (0.3:0.9) and Z8:L8-BO
TOSCs with all parameters simultaneously improved is rarely reported (0.3:0.9) deliver LUMO energy levels of −3.88 and −3.84 eV and HOMO
due to the trade-off among VOC, JSC and FF. energy levels of −5.72 and −5.72 eV, respectively. The variation of energy
In this Article, we design asymmetric NFA (Z8) featuring phenyl- levels agrees well with the density functional theory (DFT) calculations
substitution side chains that exhibits desirable optoelectronic proper- (Supplementary Fig. 4). Z-type acceptors exhibit more delocalized
ties, such as high photoluminescence quantum yield (PLQY) and delo- orbitals, which would facilitate exciton dissociation25.
calized excitons, increasing the probability to simultaneously achieve The molecular dynamics (MD) simulations of Z8 neat film
a low non-radiative energy loss and an efficient charge generation. The were conducted to reveal the effect of phenyl-substituted alkyl side
tethered phenyl groups in Z8 can participate in the intramolecular chain on intermolecular interactions (Supplementary Fig. 8). Sup-
and intermolecular interaction, affecting the acceptor:acceptor and plementary Figs. 9–11 show the neighbour atom distribution of the
donor:acceptor intermolecular interaction. The two NFAs Z8 and tethered phenyl groups and the possible π–π stacking dimers, con-
L8-BO can form an alloy acceptor, which helps to form a favourable taining the phenyl-mediated interactions. A relatively high ratio of
nanomorphology by blending with polymer donor D18, including suit- phenyl-mediated π–π stacking was obtained, and the tethered phe-
able phase separation, vertical component distribution and enhanced nyl groups trend to interact with the electron-withdrawing parts of
crystallinity. We achieve a high PCE of 20.2% in D18:Z8:L8-BO-based the NFAs. Moreover, the phenyl-alkyl groups are expected to restrict
single-junction OSCs. The photovoltaic parameters (VOC, JSC and molecular motion via phenyl-mediated intermolecular interactions,
FF) achieved maximal gain in ternary devices benefitting from the leading to a suppressed electron–phonon coupling and a reduced
improved charge extraction and reduced charge recombination, result- non-radiative energy loss43, as discussed below. The miscibility between
ing in efficient photon and carrier management. the active layer materials has a significant influence on the resulting
blend nanomorphology. Therefore, contact angle measurements were
Non-fullerene design and properties performed to investigate the miscibility between materials (Supple-
Aromatic additives have been widely utilized to fine-tune the nano- mentary Fig. 14), and the related data are presented in Supplementary
morphology of NFA-based OSCs, which can reduce the intermolec- Table 2. The Flory–Huggins interaction parameter (χ) for the donor–
ular absorption energy, leave relatively sufficient free volumes and acceptor pairs (D18:Z7, D18:Z8 and D18:L8-BO) give low χ values of
decrease nucleation sites, thus mitigating over-agglomeration of ~0.4 K, implying the good miscibility between donor and acceptors,
NFAs and improving molecular stacking with enhanced crystallin- which is beneficial to supplying sufficient donor–acceptor interfaces
ity41. Moreover, recent works also demonstrated that asymmetric for exciton dissociation. Additionally, the low χ values for Z7:L8-BO
NFAs present a promising potential in constructing high-performance (0.002 K) and Z8:L8-BO (0.004 K) generate great promise for forma-
TOSCs via modifying the nanomorphology of blend films42. Here, we tion of alloy acceptors.
designed NFAs 2,2′-((2Z,2′Z)-((12,13-bis(2-phenethyl-6-phenylhexyl)-
3,9-diundecyl-12,13-dihydro-[1,2,5]thiadiazolo[3,4-e]thieno[2′,3′:4′,5′]- Photovoltaic performance
thieno[2′,3′:4,5]pyrrolo[3,2-g]thieno[2′,3′:4,5]thieno[3,2-b]indole-2, To achieve complementary absorption and matched energy level, in
10-diyl)bis(methaneyl-ylidene))-bis(6,7-difluoro-3-oxo-2,3-dihydro- this work, poly[(2,6-(4,8-bis(5-(2-ethylhexyl-3-fluoro)thiophen-2-yl)-
1H-cyclopenta[b]naphthalene-2,1-diylidene))dimalononitrile (Z7) with benzo[1,2-b:4,5-b’]dithiophene))-alt-5,5′-(5,8-bis(4-(2-butyl-octyl)
symmetric structure and 2-((Z)-2-((10-(((Z)-1-(dicyanomethylene)- thiophen-2-yl)dithieno[3′,2′:3,4;2′,3′:5,6]benzo[1,2-c][1,2,5]thiadia-
5,6-difluoro-3-oxo-1,3-dihydro-2H-inden-2-ylidene)methyl)-12,13-bis zole)] (D18) was selected as polymer donor, and the corresponding
(2-phenethyl-6-phenylhexyl)-3,9-diundecyl-12,13-dihydro-[1,2,5] absorption spectrum and energy level diagram are shown in Fig. 1b,c.
thiadiazolo[3,4-e]thieno[2′,3′:4′,5′]thieno[2′,3′:4,5]pyrrolo[3,2-g] The detailed device optimizations, including D:A ratio and film
thieno[2′,3′:4,5]thieno[3,2-b]indol-2-yl)methylene)-6,7-difluoro-3-o thickness, are displayed in Supplementary Tables 3 and 4 and Sup-
xo-2,3-dihydro-1H-cyclopenta[b]-naphthalen-1-ylidene)malononitrile plementary Fig. 15. The optimal PCE values for D18:Z7-, D18:Z8- and
(Z8) with asymmetric terminal groups by introducing four aromatic D18:L8-BO-based binary solar cells are 17.6%, 18.6% and 18.6%, respec-
substituents (phenyl groups) on the inner side chains of the conjugated tively (Fig. 1d and Table 1). Compared with D18:L8-BO, the D18:Z7 and
backbone to finely modulate the aggregation behaviour. The molecular D18:Z8 binary OSCs show higher JSC of 26.5 mA cm−2 and 26.7 mA cm−2,
structures are shown in Fig. 1a. The detailed synthetic routes for NFAs respectively, attributed to a broader absorption range, as evidenced
Z7 and Z8 are depicted in Supplementary Scheme 1. by the external quantum efficiency (EQE) spectra (Fig. 1e). Because
The photoelectric properties of Z7 and Z8 are exhibited in Fig. 1b,c of the high-lying LUMO energy level in L8-BO, D18:L8-BO displays a
and Supplementary Figs. 1–7. The maximum absorption peaks for Z7, higher VOC of 0.92 V, surpassing the VOC values in D18:Z7 (0.88 V) and

Nature Energy
Article https://doi.org/10.1038/s41560-024-01557-z

a b c
S
N N

C11H23 S S
C11H23

Normalized absorption intensity


CN D18 Z7 Z8 L8-BO
S
N N NC
1.0 D18 EVAC
NC S CN Z7
Z8
O Ph Ph O
0.8 L8-BO
F

F Ph Ph
–4.57 eV –4.85 eV –4.95 eV –4.94 eV
Z8 F
0.6
F

N
S
N
LUMO –2.88 eV
0.4
C11H23 S S –3.96 eV –3.91 eV –3.84 eV
C11H23
CN NC
S
N N
0.2
NC S CN EF
–0.68 eV
O Ph HOMO –5.25 eV –0.75 eV –0.73 eV –0.78 eV
Ph O
0 –5.60 eV –5.68 eV –5.72 eV
400 600 800 1,000
Ph Ph
Z7
F
F Wavelength (nm)
F
F

d e f
5 100
30 20
Current density (mA cm−2)

Current density (mA cm−2)

Power density (mW cm−2)


0
D18:Z7 80
−5 D18:Z8
D18:L8-BO
D18:Z7:L8-BO 60 20
EQE (%)

−10 D18:Z8:L8-BO D18:Z7


−15 D18:Z8 10
40 D18:L8-BO
D18:Z7:L8-BO 10
−20 D18:Z8:L8-BO
20 Certified result
−25 PCE = 19.8%

−30 0 0 0
0 0.5 1.0 400 600 800 1,000 0 0.2 0.4 0.6 0.8 1.0
Voltage (V) Wavelength (nm) Voltage (V)

g h VOC (V) D18:Z7


0.93 D18:Z8
0.30 ∆E1 ∆E2 ∆E3 D18:L8-BO
D18:Z7:L8-BO
0.25
Energy loss (eV)

0.90 D18:Z8:L8-BO
JSC 27.5 FF
0.20 (mA cm–2)
27.0
26.5 0.80
26.0
25.5 0.87
0.96 0.76
0.84

0.05
0.98
0.88

0 1.00
0.92
Pcoll Pdiss
BO
BO
O
Z8
Z7

-B
8:
8:

8-
8-
L8
D1
D1

:L
:L
8:

Z8
Z7
D1

8:

8:
D1

D1

Fig. 1 | Materials and device performance. a, Chemical structures of Z7 and Z8. energy level, respectively. d, Characteristic J–V curves. e, Corresponding EQE
Different colours contribute to enhancing the distinction between molecular curves based on D18:Z7, D18:Z8, D18:L8-BO, D18:Z7:L8-BO and D18:Z8:L8-BO
structures: grey, electron-donating units; blue/yellow, terminal electron- devices. The D:A weight ratios are fixed at 1:1.2 for binary devices and 1:0.3:0.9
withdrawing units; pink, phenyl groups. b, Normalized absorption intensity for for ternary devices. f, Certified J–V and power density curves based on the
pure D18, Z7, Z8 and L8-BO films. c, The energy level diagram of D18, Z7, Z8 and optimal D18:Z8:L8-BO device, obtained from the NIM. g, The detailed energy
L8-BO neat films. EVAC, vacuum energy; EF, Fermi energy level; LUMO, the LUMO loss for the binary and ternary OSCs. h, Parameter comparison for the binary
energy level; HOMO, the HOMO energy level. The values from top to bottom and ternary OSCs.
represent the work function, LUMO energy level, injection barrier and HOMO

Table 1 | Photovoltaic parameters

Active layer VOC (V) JSC (mA cm−2) FF (%) PCE (%) Eloss (eV) ∆E3 (eV)

D18:Z7 (1:1.2)
a
0.88 (0.87 ± 0.003) 26.5 (26.2 ± 0.2) 76.1 (76.3 ± 0.6) 17.6 (17.4 ± 0.1) 0.50 0.18
D18:Z8a (1:1.2) 0.89 (0.88 ± 0.003) 26.7 (26.6 ± 0.3) 78.4 (77.8 ± 0.8) 18.6 (18.2 ± 0.2) 0.51 0.19
D18:L8-BOa (1:1.2) 0.92 (0.92 ± 0.003) 25.9 (25.3 ± 0.3) 78.5 (78.9 ± 0.7) 18.6 (18.3 ± 0.2) 0.53 0.20
D18:Z7:L8-BOb (1:0.3:0.9) 0.91 (0.91 ± 0.004) 26.5 (26.1 ± 0.4) 78.3 (78.1 ± 0.8) 18.9 (18.6 ± 0.2) 0.52 0.19
D18:Z8:L8-BO (1:0.3:0.9) b
0.92 (0.92 ± 0.003) 27.2 (27.2 ± 0.3) 80.8 (79.4 ± 0.8) 20.2 (19.8 ± 0.3) 0.51 0.19
D18:Z8:L8-BOc (1:0.3:0.9) 0.90 27.0 81.0 19.8 – –
a
The average values with standard deviation were calculated from 30 devices. bThe average values with standard deviation were calculated from 50 devices. cThe certified PCE from NIM, China.
Summarized photovoltaic parameters of binary and ternary OSCs under AM 1.5G illumination, 100 mW cm–2.

D18:Z8 (0.89 V). Such a phenomenon inspired us to combine Z-type which achieved the high PCE of 20.2%, with a high VOC of 0.92 V, JSC of
acceptors with L8-BO to construct ternary devices, thus simultaneously 27.2 mA cm−2 and FF of 80.8%. The higher JSC and FF values of the ter-
achieving desired VOC and JSC. Inspiringly, an all-round enhancement of nary blends are correlated with the broadened light absorption, more
photovoltaic parameters was observed in D18:Z8:L8-BO (1:0.3:0.9), efficient charge generation, transport and collection properties as

Nature Energy
Article https://doi.org/10.1038/s41560-024-01557-z

discussed below. The maintained high VOC of 0.92 V in D18:Z8:L8-BO contributing to the reduction in energy loss and charge recombination.
can be ascribed to the reduced energy loss. However, with further addi- Given the fast non-radiative decay rate during the exciton relaxation
tion of Z8 into the D18:L8-BO host system, both VOC and FF exhibit a process, minimizing ∆E3 has been regarded as a key and challeng-
decreased trend, leading to inferior efficiency of 18.9%. Supplemen- ing issue in reducing the total Eloss. D18:Z7 and D18:Z8 binary devices
tary Fig. 16 shows that the Z8-based ternary exhibits a simultaneous exhibit smaller ∆E3 values of 0.18 and 0.19 eV than that of D18:L8-BO
reduction in Vloss, Jloss and FFloss, where Vloss, Jloss and FFloss are defined as (0.20 eV), which is consistent with the higher PLQY of Z7 and Z8 (Sup-
the difference between realistic and ideal parameters, providing an plementary Table 7). To our delight, D18:Z7:L8-BO and D18:Z8:L8-BO
explanation for the inspiring improvement in efficiency. It is notewor- ternary devices maintained high electroluminescence with suppressed
thy that the PCE of 20.2% is the highest ever reported for single-junction non-radiative recombination loss, leading to low ∆E3 values of 0.19 eV.
OSCs, and a certified PCE of 19.8% obtained from the National Institute From Supplementary Figs. 24–26, it can be observed that Z-type accep-
of Metrology (NIM), China, as depicted in Fig. 1f and Supplementary tors exhibit more evident delocalized singlet exciton (DSE) signals,
Fig. 17. Besides, the maximum power point tracking measurements were which are well maintained in their corresponding binary and ternary
performed for D18:Z8-, D18:L8-BO- and D18:Z8:L8-BO-based devices, blends. The DSE signals can offer an additional pathway (LE-DSE-CS)
as shown in Supplementary Fig. 18. After ~500 h of photo-ageing, the for exciton dissociation, where CS refers to charge-separated state,
D18:Z8:L8-BO-based device shows superior stability compared with which is beneficial to reducing CT-mediated recombination loss35.
the corresponding binary devices, retaining 83% of its original effi-
ciency, indicating that the ternary strategy can effectively improve the Charge dynamics
stability of OSCs. Supplementary Table 5 and Supplementary Fig. 19 To deeply understand the improved JSC and FF in ternary OSCs, it is
demonstrate the versatility of the ternary strategy in other OSC systems necessary to investigate the charge dynamics process, including exci-
(D18:Z8:BTP-eC9, D18:Z8:N3 and D18:Z8:Y6), all of which achieved ton generation, dissociation, charge transport and collection. In gen-
excellent photovoltaic performance. eral, NFA-based OSCs show ultrafast electron transfer because of the
Figure 1e illustrates the corresponding EQE response spectra for large LUMO energy offset (>0.3 eV) between donors and acceptors;
the binary and ternary OSCs. Compared with the D18:L8-BO-based thus, the determining factor for charge transfer efficiency lies in the
binary device, the D18:Z8:L8-BO-based ternary devices exhibit a hole transfer. Transient absorption (TA) spectroscopy was applied to
broader absorption range and enhanced EQE values above 80% investigate the exciton dissociation process in the binary and ternary
across 600–800 nm, contributing to the increased JSC. Because of systems, in which only the acceptor can be excited with a low-energy
the unique bulk-heterojunction (BHJ) construction in OSCs, the pump beam at 800 nm. The positive signals at 590 nm can be assigned
photon-to-electron conversion efficiency is subject to exciton disso- to the ground state bleaching (GSB) of D18, where no GSB signals were
ciation and charge collection. Thus, the charge dissociation probability observed from the neat acceptor films (Fig. 2a and Supplementary
(Pdiss) and charge collection probability (Pcoll) for all OSCs were calcu- Fig. 27). Therefore, the hole transfer process can be assessed by fitting
lated from the photo-induced current density against effective voltage the donor GSB (590 nm) traces with a double exponential function,
plots (Supplementary Fig. 20). The D18:Z8:L8-BO-based devices exhibit as displayed in Fig. 2c and Supplementary Table 8. For binary blends,
the highest Pdiss of 91.4% and Pcoll of 99.6% (Fig. 1h), enabling a high the Z-type acceptors exhibit faster exciton diffusion and exciton dis-
charge generation probability and consequently yielding the best JSC sociation compared with those of L8-BO. Additionally, the incorpo-
and FF. In contrast to relevant binary cells, the Z8-based ternary OSCs ration of Z-type acceptors can accelerate the hole transfer process.
simultaneously improved VOC, JSC and FF, producing a high PCE over To quantitatively evaluate the hole transfer efficiency for binary and
20%, underlining the significance of the ternary strategy. ternary blends27, the photoluminescence quenching experiments were
performed (Supplementary Fig. 30). The highest quenching efficiency
Energy loss analysis of 97.3% in the D18:Z8:L8-BO ternary blend indicates superior charge
To gain insight into the high VOC obtained in Z8-based ternary OSCs, transfer probability. These phenomena demonstrated that the intro-
a detailed energy loss (Eloss) analysis was performed (Fig. 1g), and duction of Z8 can effectively mitigate the geminate recombination
the relevant data are summarized in Supplementary Table 6. The and yield efficient charge generation, allowing the increase in JSC value.
Eloss (Eloss = Eg − qVOC) for D18:Z7, D18:Z8 and D18:L8-BO binary OSCs Transient photocurrent (TPC), transient photovoltage (TPV) and
is calculated to be 0.50, 0.51 and 0.53 eV, respectively, and the values charge extraction (CE) measurements were conducted to assess the
for D18:Z7:L8-BO and D18:Z8:L8-BO are 0.52 and 0.51 eV, respectively, charge recombination kinetics in the binary and ternary cells45. From
where two TOSCs exhibited lower Eloss than that of D18:L8-BO-based the TPC results (Supplementary Fig. 31), Z8-based ternary solar cell
binary OSCs34,44. According to the detailed-balance theory, Eloss can exhibits rapid rise and decay processes and possesses a shorter charge
be divided into three parts: Eloss = q(∆VSQ + ∆Vr + ∆Vnr) = ∆E1 + ∆E2 + ∆E3, extraction time, suggesting rapid charge generation and extraction
where ∆E1 is the radiative loss above the bandgap derived from the with fewer charge traps. Figure 2d plots the charge density against VOC,
S-Q limit and ∆E2 and ∆E3 are the radiative and non-radiative energy and the largest charge carrier density (nCE) in D18:Z8:L8-BO implies
losses below the bandgap, respectively. the most efficient charge generation. Moreover, the ternary OSCs
∆E1 is inevitable for any kind of photovoltaics, determined yield a longer carrier lifetime (τ) than those of the corresponding
only by the bandgap. The ∆E1 values for the investigated devices binary devices (Fig. 2e), which is related to reduced charge recombi-
are 0.26–0.27 eV. The binary OSCs exhibit the same ∆E2 values of nation. The charge recombination can be evaluated using the equation
0.06 eV, whereas Z8-based TOSCs present a reduced ∆E2 value 1 1
k= ∝ , where k is the recombination coefficient that can
of 0.05 eV, which can be ascribed to the decreased energetic disorder RτnCE τnCE

of band tail states. In Supplementary Fig. 21b–f, D18:Z8:L8-BO exhibits reflect the overall recombination occurring in blend, and R is
the lowest Urbach energy (EU) value of 22.8 meV, implying that the constant, representing the recombination order. Compared with binary
introduction of Z8 into D18:L8-BO can effectively reduce energetic dis- OSCs, ternary OSCs deliver lower k values under varied charge densities
order, not only allowing the decrease in Eloss but also helping to reduce (Fig. 2f), implying suppressed charge recombination in ternary devices.
trap density with restricted charge recombination24. The impedance It is widely accepted that the FF of OSCs is determined by the
and capacitance–voltage measurements were performed to evaluate competing process of charge extraction and charge recombination
the trap density (Nt) in ternary and binary OSCs. As shown in Supple- at a low internal electric field. The recombination mechanism can be
mentary Fig. 22, ternary OSCs deliver reduced Nt and D18:Z8:L8-BO- revealed by the relationship between VOC and light intensity (Plight)46.
based ternary cells exhibit the lowest Nt value of 6.33 × 1021 m−3, As shown in Supplementary Fig. 20b, all the investigated binary and

Nature Energy
Article https://doi.org/10.1038/s41560-024-01557-z

a b c
–3 –3
∆T/T(10 ) ∆T/T(10 )

3 Z8
Z8 1.0 3 D18:Z8:L8-BO 1.0
10 10
3

Time delay (ps)

Time delay (ps)


10
2 0 10
2
Donor 0

∆T/T (10–3)
CS
GSB ESA GSB Acceptor 2
1 −1.0 1 ESA −1.0
10 10 Donor GSB
1 D18:Z7
D18:Z8
D18:L8-BO
Acceptor D18:Z7:L8-BO
1 1 GSB D18:Z8:L8-BO
0
0 0
600 700 800 900 600 700 800 900 −1 0 1 10 100 1,000
Wavelength (nm) Wavelength (nm) Time delay (ps)

d e f
8
D18:Z7
Charge density (×1016 cm−3)

D18:Z7 D18:Z7
6 D18:Z8 D18:Z8 D18:Z8
D18:L8-BO D18:L8-BO D18:L8-BO

Carrier lifetime (µs)


D18:Z7:L8-BO 10 D18:Z7:L8-BO D18:Z7:L8-BO
4 D18:Z8:L8-BO D18:Z8:L8-BO D18:Z8:L8-BO

k (m3 s−1)
−17
10
2

−18
10
0.65 0.70 0.75 0.80 0.85 0.90 0.95 1 2 3 4 5 6 1 2 3 4 5 6
16 −3
VOC (V) Charge density (×10 cm ) Charge density (×1016 cm−3)

g 0 –15
h i
10 10 1.5 80
Ldiff
kex (×107)/krec (×107) (s−1)

1.0
Ldr
0.5 60 900
kL/γ (m3 s−1)

Ldiff (nm)

Ldr (nm)
γpre

10
–1
0.5 40 600
1.0
γpre 10
–16 1.5 kex 20 300
kL
γ 2.0 krec
–2
10 2.5 0 0
BO

BO
BO

BO

BO

BO
O

O
Z8

Z8

Z8
Z7

Z7

Z7
-B

-B

-B
8:

8:

8:
8:

8:

8:
8-

8-
8-

8-

8-

8-
L8

L8

L8
D1

D1

D1
D1

D1

D1
:L
:L

:L

:L

:L
:L
8:

8:

8:
Z8

Z8
Z7

Z7

Z8

Z7
D1

D1

D1
8:

8:

8:
8:

8:

8:
D1

D1

D1
D1

D1

D1
Fig. 2 | Charge dynamics analysis. a,b, Femtosecond TA spectra of Z8 neat reduction factor (γpre), Langevin recombination coefficient (kL) and bimolecular
film (a) and D18:Z8:L8-BO blend film (b). ESA, excited-state absorption; recombination coefficient (γ) of the devices. h, Bimolecular recombination
T, transmittance. c, The trace kinetic of donor GSB signal probed at 590 nm. rate (krec) and charge extraction rate (kex) for the binary and ternary OSCs. i, The
d, Charge density versus VOC. e,f, Carrier lifetime (e) and recombination rate exciton diffusion length (Ldiff ) and charge drift length (Ldr) of the devices.
coefficient (f) as a function of the charge density of the devices. g, The Langevin

ternary cells are mainly subjected to bimolecular recombination. in D18:Z8:L8-BO-based OSCs, it can be concluded that the incorpora-
Therefore, the FF–θ analysis model was introduced to quantitatively tion of Z8 is favourable for reducing charge recombination, both in
analyse the improvement of FF in the optimal TOSCs47. The ratio of geminate and non-geminate recombination, as well as elevating charge
krec transport and extraction.
extraction to recombination, θ, can be defined as θ ∝ , where krec
kex

and kex are the bimolecular recombination and charge extraction rate, Morphology analysis
respectively. krec is proportional to the bimolecular recombination The aggregation behaviour of polymer donor and NFAs in the blend
constant (γ) that obeys the equation of γ = γprekL, in which γpre and kL are film is of vital importance for achieving high photovoltaic performance
the Langevin reduction factor and the classic Langevin recombination in OSCs, and the nanomorphology of the neat and blend films was
constant, respectively (the related data are displayed in Supplemen- systematically investigated. The surface morphology for different
tary Table 9)48. As illustrated in Fig. 2g, D18:Z8:LB-BO cells have small binary and ternary blends was evaluated by atomic force microscopy
γpre and γ values of 0.026 and 2.31 × 10−17 m3 s−1, respectively, leading (AFM) as displayed in Fig. 3a and Supplementary Figs. 38 and 39. The
to minimal carrier loss. Figure 2h displays the kex and krec variations in smaller roughness (Rq) value in the D18:Z8:L8-BO-based blend indicates
ternary and binary OSCs. All the devices give larger kex values of ~107 s−1 a more uniform surface achieved by adding Z8 as the third component,
than that of krec (~104 s−1), guaranteeing efficient charge collection. promoting the formation of superior ohmic contact with reduced series
D18:Z8:L8-BO-based TOSCs exhibit the highest kex value of 1.33 × 107 s−1 resistance. Transmission electron microscopy (TEM) measurement
and the lowest krec value of 1.45 × 104 s−1, allowing faster charge extrac- was performed to study the bulk texture (Supplementary Fig. 40).
tion with suppressed charge recombination, thus leading to a desired JSC A relatively obvious fibre structure is visible in the D18:Z8 and D18:Z7
and FF gain. The FF-versus-θ plot is provided in Supplementary Fig. 37, blend films, but it is absent in the D18:L8-BO blend. For ternary blends,
and a reduced θ is accompanied by the improvement of FF. In addition, the D18:Z8:L8-BO possesses a relatively high-quality fibril texture,
Z8-based TOSCs show superior charge transport properties with the indicating that the addition of Z8 can better enhance the crystallinity.
highest Ldiff (69.9 nm) and Ldr (1070 nm) values (Fig. 2i), contributing The fibril networks can be clearly observed by using the photo-induced
to decreasing recombination loss49. Combining the higher JSC and FF force microscopy measurements (Supplementary Fig. 42). Both donors

Nature Energy
Article https://doi.org/10.1038/s41560-024-01557-z

a b
D18:Z8
2.0 D18:Z8 D18:L8-BO D18:Z8:L8-BO
Rq = 1.13 nm

Intensity (a.u.)
1.6 3.5

qz (Å−1)
3.0
1.0 2.5
0.6 2.0
0
0.5 0 –0.5 –1.0 –1.5 –2.0 0.5 0 –0.5 –1.0 –1.5 –2.0 0.5 0 –0.5 –1.0 –1.5 –2.0
4.0 nm –4.0 nm 8.0° –8.0°
qxy (Å−1) qxy (Å−1) qxy (Å−1)
D18:L8-BO
Rq = 1.17 nm
c 10
9 d 19
(010) (010)
8
(001) 18
10 D18:Z7

Crystal coherence length (Å)


7 17
10
D18:Z8 16
6
10

Intensity (a.u.)
88 (100)
4.0 nm –4.0 nm 7.0° –7.0° 5
10 D18:L8-BO 84
D18:Z8:L8-BO 4
10 80
Rq = 0.96 nm
3
D18:Z7:L8-BO 76
10
80 (001)
2
10 70

1
D18:Z8:L8-BO 60
10
(100) 50
10
0
40
1 2

Z8

BO

BO
Z7
3.3 nm –3.3 nm 10.0° –10.0°

-B
8:
q (Å−1)

8:

8-

8-
L8
D1
D1

:L

:L
8:

Z7

Z8
D1

8:

8:
D1

D1
Fig. 3 | Morphology characterizations. a, AFM (3 × 3 μm) height (left) and based on the blend films (the solid lines represent the OOP direction and the
phase (right) images of the optimized D18:Z8, D18:L8-BO and D18:Z8:L8-BO. dashed lines represent the IP direction). d, Crystal coherence length (CCL) of
b, Two-dimensional GIWAXS patterns for the D18:Z8, D18:L8-BO and D18:Z8: (010), (100) and (001) diffraction peaks in the binary and ternary blend films.
l8-BO blends. c, The in-plane (IP) and OOP line cut profiles of the 2D GIWAXS

and acceptors can form a fibril structure, and they entangled together The amorphous intermixing phase (ζ) sizes for D18:Z7, D18:Z8 and
to construct an interpenetrating bicontinuous nanostructure, facili- D18:L8-BO blends are 30.0, 25.5 and 37.1 nm, respectively. Adding
tating suitable phase separation and efficient charge transport. This Z8 into D18:L8-BO blend produces a smaller intermixing domain
favourable nanotexture can reduce both geminate and non-geminate size (31.8 nm), while an increased ζ value of 45.1 nm is observed in
recombination and, thus, yielding higher JSC and FF. D18:Z7:L8-BO-based ternary blends, which may lead to the more charge
The crystalline behaviours of different films were assessed by recombination. Moreover, the incorporation of Z8 into D18:L8-BO
the grazing-incidence wide-angle X-ray scattering (GIWAXS). Supple- blends can improve the acceptor domain size (26.7 nm), thus enhanc-
mentary Figs. 43–45 and Fig. 3b,d display 2D GIWAXS patterns for neat ing electron transport property. The time-of-flight secondary ion
and blend films, and the detailed data are summarized in Supplemen- mass spectrometry (TOF-SIMS) was applied to investigate the vertical
tary Tables 10 and 11. All the acceptors exhibit a preferential face-on component distribution in blend films. As displayed in Supplementary
orientation, in which strong π–π stacking diffraction (010) peaks in Fig. 48, Z8-based ternary film can maintain a favourable vertical dis-
the out-of-plane (OOP) direction are clearly seen, benefitting to the tribution, while the addition of Z7 into D18:L8-BO blends is unable to
formation of vertical charge transport50. L8-BO neat film exhibits a retain the desired component variation along the vertical direction.
broad π–π stacking peak with a shoulder peak located at 1.51 Å−1 (OOP), Therefore, we deduced that the incorporation of Z8 into D18:L8-BO
suggesting the presence of amorphous content7, while Z7 and Z8 neat blends can reduce the amorphous intermixing phase, enlarge the
films exhibit stronger diffraction intensity and narrower π–π stacking acceptor domain and form a desired vertical distribution, leading to
peaks, implying less disordered states and parasitic crystallinity. Thus, decreased charge recombination, improved charge transport and
the incorporation of Z-type acceptors into D18:L8-BO contributes to collection, thus realizing higher JSC and FF.
enhancing crystallinity as evidenced by the stronger (010) and (100) dif- To elucidate the correlation among molecular structure, blend
fraction intensity and enlarged CCLs (Fig. 3d). Besides, the (001) peak at morphology and device performance, a deep insight into the inter-
~0.5 Å−1 can be assigned to the characteristic peak of D18 (Supplemen- molecular interactions was further studied by using MD simulations.
tary Fig. 44). The larger CCL of the (001) peak in Z-type acceptor-based Supplementary Fig. 49 shows the detailed simulation processes and
binary and ternary blends indicates the favourable crystallinity of the related data are summarized in Supplementary Tables 14–16. As
D18 (Fig. 3d). The enhanced crystallinity in D18:Z8:L8-BO contributes illustrated in Supplementary Fig. 53, the electron-withdrawing units
to improving charge transport property, leading to higher and more (dithieno[3′,2′:3,4;2′,3′:5,6]benzo[1,2-c][1,2,5]thiadiazole) of polymer
balanced charge carrier mobilities (Supplementary Figs. 34–36 and donor D18 trend to interact with the 2-(5,6-difluoro-3-oxo-2,3-dihy-
Supplementary Table 12). dro-1H-inden-1-ylidene)malononitrile (IC-2F)/2-(6,7-difluoro-3-oxo-
The grazing-incidence small-angle X-ray scattering (GISAXS) meas- 2,3-dihydro-1H-cyclopenta[b]naphthalen-1-ylidene)malononitrile
urements were performed to quantitatively analyse the phase separa- (ITN-2F) terminals of NFAs. The respective acceptor–acceptor (A–A)
tion of blend films, and the relevant data are shown in Supplementary dimers and donor–acceptor (D–A) pairs were extracted from the corre-
Figs. 46–47 and Supplementary Table 13. Compared with D18:L8-BO sponding MD simulation, and the relevant exciton binding energies (Eb)
film, Z-type blends exhibit a larger values of acceptor domain size (2Rg), of excited and CT states were calculated, as displayed in Supplementary
which can be ascribed to the higher crystallinity of Z-type acceptor. Figs. 54–55 and Supplementary Table 14. For a given type of A–A dimer,

Nature Energy
Article https://doi.org/10.1038/s41560-024-01557-z

a c e ηSQ
100 50 1.0
2.0 ∆E1 Reflectance 0.9
∆E2 Parasitic absorption
55
Si (ref. )
0.8

v*f (VOC × FF/VOC,SQ × FFSQ)


Current density (mA cm–2)
80 ∆E3 IQE loss 40 Perovskite (ref. )
54
0.7
1.6
qVOC
Light percentage (%)

ITO JSC 0.8


JSC,SQ D18:Z8:L8-BO (this work) 0.6
PEDOT 58
0.5

Energy (eV)
PBQx-TF:eC9-2Cl:F-BTA-3 (ref. )
60 qVOC,SQ 30
1.2
24
Simulated EQE PM6:Y11 (ref. ) 11
D18:PM6:L8-BO (ref. )
0.4
PDINN 0.6 57
PM6:IT4F (ref. )
56
0.3
40 Ag 20 PTB7-Th:ATT-9 (ref. )
0.8 0.2
Reflectance 24 0.1
23 PM6:PCBM (ref. )
Transmittance 0.4 PTB7:PCBM (ref. )
<0.1
20 0.4 10
Test EQE
23
P3HT:PCBM (ref. )
0 0 0 0.2
400 600 800 1,000 Eg JSQ 0.4 0.6 0.8 1.0
Wavelength (nm) j/(JSC/JSC,SQ)
b d Eg (eV) f S-Q limit
PCE (%) at 1.43 eV
2
|E| 2.1 1.6 1.2 1.0
33.7 ∆E3
3.10 33.1%
9.5 6.4 ∆E2
Ag FFloss
300 Jloss (par. abs. & R)
2.71 0.5
PDINN 18.6 15.5 27.6
12.5 12.5 IQEloss 27.3%
25.9%
0.4 23.7%
2.03
Position (nm)

21.6

PCE (%)
200 21.0%
∆Vnr (V)

D18:Z8:L8-BO 20.2%
0.3
15.5 24.6
PEDOT:PSS 1.35 15.5
0.2 27.3
100 ITO
0.68 9.46
0.1 21.6 30.7 27.6
Photon loss Carrier loss
0 0 3.40
400 600 800 1,000 600 800 1,000 1,200
Wavelength (nm) Wavelength (nm)
Fig. 4 | Theoretical evaluation. a,b, Light percentage distribution (a) and optical represents the ideal PCE at the bandgap and ∆Vnr of D18:Z8:L8-BO. e, Fraction
simulation on electric field intensity distribution (b) across the D18:Z8:L8-BO of S-Q limit (at their corresponding bandgap) for photon loss (j = JSC/JSC,SQ) and
device with a conventional structure of ITO/PEDOT:PSS/D18:Z8:L8-BO/PDINN/ carrier loss (v*f = VOC/VOC,SQ × FF/FFSQ) achieved by typical and state-of-the-
Ag. Simulated EQE assumes 100% IQE. c, The detailed voltage loss and current art OSCs11,23,24,54,55,68–70. f, Detailed various factors contributing to PCE loss in
loss for the D18:Z8:L8-BO-based device, VOC,SQ and JSC,SQ, are the maximum D18:Z8:L8-BO-based OSC on approaching the S-Q limit efficiency. The arrow
voltage and current density according to the S-Q limit. d, The detailed balanced indicates the direction for performance improvement. Jloss(par.abs.&R) represents the
limit of PCE under different bandgap wavelengths and ∆Vnr. The red star photon loss caused by parasitic absorption and device reflection.

the Z-type acceptors show the relatively smaller Eb values compared FFs exist in the relevant binary OSCs. A strong D–A interaction between
with L8-BO, generating a higher probability of direct photogeneration host donor and guest acceptor and a good miscibility between the guest
in acceptor domain35,51. The MD simulation of ternary blends is depicted and host acceptors are necessary to form favourable nanomorphology,
in Supplementary Fig. 56, from which more comprehensive insights thus enhancing the optoelectronic process. All these factors contribute
can be gained and the detailed insights are provided in Supplementary to achieving an overall improvement in OSC performance.
Figs. 57–59 and Supplementary Tables 15 and 16. Supplementary Fig. 57
reveals that the tethered phenyl groups on the Z-type acceptor can Theoretical evaluation
interact with polymer donor, thereby enhancing the D–A intermolecu- A theoretic calculation was performed to evaluate the photovoltaic
lar interaction. The average intermolecular binding energies (ΔE) are performance reliability of Z8-based TOSCs. For photovoltaic materi-
−3.903, −4.076 and −3.499 kJ mol−1 for D18:Z7, D18:Z8 and D18:L8-BO, als with a given bandgap, the realistic efficiency is inferior to the S-Q
respectively, indicating a stronger D–A interaction in D18:Z8 pairs. limit, which is mainly caused by the non-ideal charge carrier collec-
Supplementary Fig. 58 visualizes the interacting atoms between the tion and recombination and then leads to current and voltage losses.
host donor and the host/guest acceptors in two ternary blends. The Polman et al. proposed two characteristic parameters52, j (j = JSC/JSC,SQ)
D18:Z8:L8-BO-based ternary blend exhibits a higher number of D–A and v*f (v*f = VOC/VOC,SQ × FF/FFSQ), to evaluate photovoltaic systems,
(guest) pairs in the region with a large number of interacting atoms which is also performed in this study. Optical simulation based on the
compared to that of the D18:Z7:L8-BO-based blend, implying a higher transfer matrix method was carried out to evaluate the current loss
probability to forming D18:Z8 pairs with stronger intermolecular in D18:Z8:L8-BO ternary OSC53. Figure 4a,b illustrates the light per-
interaction, which contributes to facilitating charge transfer process. centage distribution in different layers and electric field distribution
Compared with the D18:L8-BO-based binary blend, the incorporation across positions in the Z8-based ternary OSCs. Light absorbed by the
of Z-type acceptors in ternary leads to an enhanced ratio of D–A pairs active layer accounts for a current density of 28.2 mA cm−2, which cor-
(Supplementary Fig. 59), implying more D/A interfaces and thereby responds to the theoretical EQE if all absorbed photons are transferred
promoting exciton dissociation. into photocurrent (IQE 100%). The experimental integrated JSC value
Considering the superior photoluminescence properties and the from the EQE test is 26.3 mA cm−2, indicating that the average IQE of our
more delocalized exciton in asymmetric NFA Z8, the formation of D18/ OSC is 93%, which is consistent with the IQE test result (Supplementary
Z8 interfaces is expected to minimize non-radiative energy loss and Fig. 63). According to the S-Q model (Fig. 4c), for Z8-based ternary OSC
geminate charge recombination. Based on the experimental results with a bandgap of 1.43 eV (867 nm), the maximum theoretical current
and MD simulations, constructing a highly efficient TOSC necessitates density (JSC,SQ) can reach 31.8 mA cm−2. A high j value of over 85% sug-
a guest acceptor with high PLQY and delocalized exciton as well as high gests the highest light management among the state-of-the-art OSCs.

Nature Energy
Article https://doi.org/10.1038/s41560-024-01557-z

The 15% loss of incident photons comes from three parts: (1) IQE loss, absorbance expansion to the subgap region often occurs due to the
(2) device reflection loss and (3) parasitic absorption of interlayers and existence of CT states and energetic disorder in realistic non-ideal
electrodes, as displayed in Fig. 4a,c. systems, and for high-performance OSCs, the CT states are very close
Since we attributed the efficiency loss from both the voltage and to the bandgap and subgap absorption is also low, resulting in a small
current sides, based on our best-performance TOSC, attention needs to ΔE2 of around 0.05 eV. If ΔE2 is totally overcome in future material sys-
be paid to several possible approaches to realize a higher PCE towards tems, a PCE of 27.3% can be achieved. ∆Vnr counts for a large part of PCE
the theoretical S-Q limit for the future development of OSCs. The VOC loss; however, OSCs are developing rapidly with reduced ∆Vnr. A PCE
of our ternary cell is 0.92 V, which accounts for 79% of the theoretical of 33.1% can be expected if no non-radiative recombination loss exists
limit of VOC,SQ (1.16 V). Considering that the radiative loss (∆E1 and ∆E2) is in a solar cell, which is close to the ultimate S-Q limit. The final gap is
inevitable for a solar cell at a certain bandgap due to mechanisms such due to the difference in bandgap where PCESQ has two local maxima of
as dissipative process, carrier cooling and entropy-related étendue 33.7% (1.34 eV, 930 nm) and 33.5% (1.15 eV, 1,080 nm).
expansion, non-radiative recombination loss (∆E3 or q∆Vnr) becomes
the most influential and the most promising factor to overcome in Conclusions
voltage loss, since multiple ways to suppress ∆Vnr are focused by many In this work, alloyed acceptors consisting of two NFAs Z8 and L8-BO in
works recently. Here, we simulated the theoretical FF and PCE limit on ternary OSCs contributed to the reduced photon and carrier losses, and
the basis of a detailed-balance model applying different ∆Vnr values in a high PCE of 20.2% (certified PCE of 19.8%) is achieved in single-junction
Fig. 4d and Supplementary Fig. 64. Our ternary OSC exhibits ∆Vnr of OSCs. The tethered phenyl groups in the asymmetric Z-type accep-
0.19 V at the bandgap wavelength of 867 nm, in which the theoretical tor can not only endow NFA with high PLQY and delocalized exciton
limits of FF and PCE are 88.1% and 27.3%, respectively, leaving a remark- but also contribute to facilitate charge transfer and minimize charge
able improvement room for further development. recombination in ternary blend to simultaneously achieve ideal VOC, JSC
Figure 4e summarizes the light management (j) and carrier man- and FF. The overall distribution of photon and carrier losses for OSCs is
agement (v*f) ability towards the detailed-balance S-Q limit of the presented through detailed theoretical calculation, and the potential
D18:Z8:L8-BO ternary solar cell along with some typical OSCs, one for improvement on each device parameter is also analysed, which
of the state-of-art perovskite54 and single-crystal silicon (c-Si)55 pho- indicates that the material and device revolution is essential to fur-
tovoltaics. The classic fullerene-based OSCs suffer from a high rate ther minimize photon and carrier losses. These findings highlight that
of non-radiative recombination, thus undergoing very large voltage both photon loss and carrier loss should gain equal attention from the
loss and can hardly exceed 40% of its PCE limit. As the development of community, which could push the photovoltaic performance of OSCs
non-fullerene acceptors and the iteration of new polymer donors, the towards the efficiency limit.
PCE of OSCs has grown rapidly and is gradually closing the gap to the
theoretical limit, especially in the direction of carrier management. In Methods
this work, a high level of carrier management was maintained, and the Materials
remarkable increase in light management was simultaneously realized All reactions involving air or moisture-sensitive compounds were
in Z8-based ternary OSCs. The D18:Z8:L8-BO ternary solar cell performs carried out in a dry reaction vessel under a positive pressure of nitro-
71.6% in carrier management (79.3% for v and 90.3% for f), and 85.5% in gen. Unless stated otherwise, starting materials were obtained from
light management, which is one of the top values among OSCs. Such an Adamas, Energy Chemical or J&K and were used without further puri-
achievement points out that the process towards S-Q limited efficiency fication. Anhydrous tetrahydrofuran (THF) and toluene were distilled
needs the simoultaneous management of light and carrier. Compared over Na/benzophenone before use.
with the c-Si counterpart with over 97% of j values, OSCs still lag behind Compound 2 was synthesized using iridium-catalysed
in device engineering and optical modulation to reduce reflection and 4-phenylbutan-1-ol 1 as the raw material, followed by bromina-
parasitic absorption to a minimum. Compared with perovskite coun- tion with an excess amount of PBr3 to afford the key intermediate
terparts with 89% of v*f, OSCs perform at almost the same level in FF ((3-bromomethyl)heptane-1,7-diyl)dibenzene 3. Compound 5 was
but have an intrinsic disadvantage in ∆Vnr due to the narrower bandgap. synthesized by N-alkylation of compound 4, which was prepared by
However, the difference is narrowing as the active layer materials are double intramolecular Cadogan reductive cyclization of the dini-
developing rapidly. More and more high-performance OSCs with low trobenzothiadiazole precursor based on reported procedures, with
∆Vnr (even less than 0.18 V) have been reported. Future development ((3-bromomethyl)heptane-1,7-diyl)dibenzene 3 using potassium
of OSCs needs to focus on both voltage and current, including further carbonate as the base in 69% yield. The dialdehyde compound 6 was
material revolution and device engineering, to achieve higher PCE. On efficiently obtained from compound 5 using the Vilsmeier–Haack reac-
the one hand, device engineering, including modification on optical tion, followed by Knoevenagel condensation with 2-(6,7-difluoro-3-oxo-
modulation layer and interlayers/electrodes, contributes to minimiz- 2,3-dihydro-1H-cyclopenta[b]naphthalen-1-ylidene)malononitrile
ing reflection and parasitic absorption. On the other hand, developing (ITN-2F), and the target molecule Z7 was obtained in 80% yield as a black
new materials with excellent mobility and emission properties offers solid. The asymmetric NFA Z8 was efficiently synthesized by a one-pot
the benefit of wiping out carrier recombination loss. Thus, to advance reaction in 35% yield, where the 2-(5,6-difluoro-3-oxo-2,3-dihydro-1H-i
the performance of OSCs, device engineering and material innovation nden-1-ylidene)malononitrile (IC-2F) and ITN-2F were sequentially
should be efficiently united to eliminate light and carrier losses. reacted with dialdehyde compound 6.
Figure 4f gives the potential improvement of PCE in different Compound 2: an oven-dried Schlenk flask was loaded with
factors based on existing D18:Z8:L8-BO ternary solar cells towards 4-phenylbutan-1-ol 1 (900 mg, 6 mmol), [Cp*IrCl2]2 (96 mg, 0.12 mmol)
the S-Q limit. By overcoming the IQEloss through possible material and t-BuOK (269 mg, 2.4 mmol) in anhydrous 1,4-dioxane (10 ml) under
optimization, the PCE would increase to 21.0%. By carefully designing N2 atmosphere. The reaction solution was stirred and refluxed for 24 h
the device structure and applying optical modulations to reduce reflec- under dark. The resulting mixture was extracted with dichloromethane
tion and parasitic absorption, which is done well in c-Si photovoltaics, (DCM) and purified by column chromatography to give 250 mg of com-
the PCE could reach 23.7% to its maximum JSC,SQ value. FF is affected by pound 2 in 30% yield. 1H NMR (400 MHz, CDCl3): δ 7.34–7.23 (m, 4H),
both materials and devices, and reaching an ideal FFSQ would increase 7.18 (m, 6H), 3.58 (d, J = 4.5, 2H), 2.63 (m, 4H), 1.61 (m, 5H), 1.39 (m, 4H).
the PCE to 25.9%. Considering the voltage loss part, the classical S-Q Compound 3: compound 2 (282 mg, 1 mmol) and PBr3 (1080 mg,
theory assumes full absorption above the bandgap and zero below 4 mmol) were dissolved in dichloromethane (10 ml), and the mixture
the bandgap, so ΔE2 does not exist in the original S-Q model. However, was stirred at room temperature for 6 h. The resulting solution was

Nature Energy
Article https://doi.org/10.1038/s41560-024-01557-z

extracted with DCM and purified on a silica-gel column chromatogra- pyridine (0.1 ml) was added in chloroform (5 ml). The reaction was
phy to give 310 mg compound 3 in 90% yield. 1H NMR (400 MHz, CDCl3): stirred at 65 °C overnight. After cooling to room temperature, the
δ 7.32–7.25 (m, 4H), 7.23–7.15 (m, 6H), 3.54–3.42 (m, 2H), 2.70–2.51 solvent was removed. The residue was purified by column chroma-
(m, 4H), 1.79–1.56 (m, 5H), 1.50–1.24 (m, 4H). 13C NMR (100 MHz, CDCl3): tography to yield a dark-blue solid (63.2 mg, 35%). 1H NMR (400 MHz,
δ 142.49, 142.02, 128.39, 128.38, 128.31, 128.27, 125.85, 125.67, 77.32, CDCl3) δ 9.23 (s, 1H), 9.14 (s, 1H), 9.09 (s, 1H), 8.56 (dd, J = 9.6, 6.4 Hz,
77.00, 76.68, 39.11, 38.85, 35.78, 34.38, 32.84, 32.30, 31.47, 26.11. 1H), 8.25 (d, J = 2.1 Hz, 1H), 7.79 (dd, J = 9.2, 7.7 Hz, 1H), 7.75–7.60 (m,
Compound 5: the cyclized intermediate 4 (373 mg, 0.5 mmol) and 2H), 7.23–7.15 (m, 4H), 7.15–7.08 (m, 2H), 7.07–6.96 (m, 4H), 6.96–6.83
compound 3 (688 mg, 2 mmol) were mixed with KI (42 mg, 0.25 mmol) (m, 6H), 6.69–6.61 (m, 2H), 6.61–6.54 (m, 2H), 4.79–4.53 (m, 4H), 3.22
and K2CO3 (414 mg, 3 mmol) and then dissolved in DMF (10 ml) under (d, J = 8.2 Hz, 4H), 2.55–2.20 (m, 8H), 2.11 (s, 2H), 1.89 (t, J = 7.8 Hz, 4H),
N2 atmosphere. The reaction mixture was stirred at 80 °C overnight. 1.35–1.19 (m, 48H), 0.87 (m, 6H). 13C NMR (101 MHz, CDCl3) δ 187.79,
The resulting solution was extracted with DCM and washed with water 186.12, 160.26, 158.85, 153.92, 153.80, 147.45, 145.33, 145.18, 142.17, 142.14,
followed by brine. The crude mixture was purified by silica-gel column 142.10, 142.07, 140.87, 140.82, 137.81, 136.12, 135.75, 135.37, 135.21, 134.06,
chromatography, and 440 mg compound 3 was obtained in 69% yield. 133.72, 133.44, 133.26, 132.51, 130.94, 130.76, 128.30, 128.28, 128.26,
1
H NMR (400 MHz, CDCl3): δ 7.19 (t, J = 7.3, 4H), 7.12 (t, J = 7.1, 2H), 7.01 128.23, 128.19, 128.09, 128.07, 128.04, 128.02, 127.87, 127.80, 125.77,
(s, 2H), 6.98–6.90 (m, 10H), 6.65–6.55 (m, 4H), 4.68–4.44 (m, 4H), 2.83 122.30, 120.08, 115.72, 115.47, 114.98, 114.57, 113.76, 68.81, 67.26, 55.71,
(t, J = 7.6, 4H), 2.38–2.18 (m, 8H), 2.16–2.04 (m, 2H), 1.91–1.84 (m, 4H), 38.14, 35.68, 35.65, 35.57, 35.54, 32.01, 31.94, 31.61, 31.48, 31.45, 31.33,
1.51–0.99 (m, 46H), 0.98–0.78 (m, 12H). 13C NMR (100 MHz, CDCl3): δ 31.27, 30.42, 30.25, 29.92, 29.85, 29.70, 29.66, 29.57, 29.49, 29.38, 25.45,
147.58, 142.29, 142.18, 141.29, 137.18, 136.87, 131.46, 128.23, 128.16, 128.05, 25.26, 22.72, 14.16. MS (MALDI-TOF) calculated for C110H104F4N8O2S5 [M]+
127.84, 125.62, 125.56, 123.53, 122.88, 119.38, 111.75, 54.94, 38.16, 35.47, 1,805.68, found 1,805.68.
32.23, 31.93, 31.74, 31.29, 30.10, 29.70, 29.64, 29.56, 29.47, 29.36, 28.87,
25.00, 22.69, 14.11. MS (MALDI-TOF) calculated for C80H98N4S5 [M]+ Structural and optoelectronic characterization
1
1,274.64, found 1,274.64. H and 13C spectra were measured with Bruker Avance 400 spectrom-
Compound 6: to a solution of compound 5 (255 mg, 0.2 mmol) eters. Chemical shifts for hydrogens are reported in parts per million
in dry THF was dropped n-BuLi/n-hexane (0.38 ml, 0.6 mmol, 1.6 M) (ppm, scale) downfield from tetramethylsilane and are referenced to
at −78 °C, followed by stirring for 1 h. Then, N,N-dimethylformamide the residual protons in the NMR solvent (CDCl3: 7.26).13C NMR spectra
was added and reacted for another 1 h. After warming to room tem- were recorded at 100 MHz. Chemical shifts for carbons are reported in
perature, the reaction mixture was extracted with DCM and purified ppm (scale) downfield from tetramethylsilane and are referenced to the
by silica-gel column chromatography to give 210 mg of compound 6 carbon resonance of the solvent (CDCl3: 77.2). The data are presented as
in 79% yield. 1H NMR (400 MHz, CDCl3): δ 10.15 (s, 2H), 7.20 (t, J = 7.4, follows: chemical shift, multiplicity (s, singlet; d, doublet; t, triplet; m,
4H), 7.13 (t, J = 7.3, 2H), 6.96 (d, J = 7.3, 4H), 6.90 (m, 6H), 6.56 (m, 4H), multiplet and/or multiple resonances; br, broad), coupling constant in
4.58–4.43 (m, 4H), 3.29–3.09 (m, 4H), 2.43–2.13 (m, 8H), 2.10–1.84 (m, hertz (Hz) and integration. MALDI measurements were performed with
7H), 1.53–0.96 (m, 36H), 0.96–0.77 (m, 12H). 13C NMR (100 MHz, CDCl3): a MALDI-FT 9.4 T, Bruker solariX, or MALDI-TOF MS Bruker Autoflex
δ 181.65, 147.41, 146.71, 143.22, 141.93, 140.78, 137.08, 136.86, 136.83, III. UV–visible was recorded with Jasco V-570 spectrometers. Cyclic
132.59, 129.44, 128.23, 128.18, 128.02, 127.75, 127.35, 127.33, 125.69, voltammetry was performed with a CHI620D potentiostat. All meas-
112.55, 112.53, 55.18, 37.97, 35.42, 31.89, 31.52, 31.27, 30.34, 30.04, 29.68, urements were carried out in a one-compartment cell under a nitrogen
29.63, 29.59, 29.52, 29.36, 29.32, 28.16, 27.90, 25.00, 22.67, 14.10. MS atmosphere, equipped with a glassy-carbon electrode, a platinum
(MALDI-TOF) calculated for C82H98N4O2S5 [M]+ 1,330.63, found 1,330.63. counter-electrode and an Ag/Ag+ reference electrode with a scan rate of
100 mV s−1. The supporting electrolyte was a 0.1 mol l−1 dichlorometh-
Z7 ane solution of tetrabutylammonium perchlorate. All potentials were
Compound 6 (247.8 mg, 0.186 mmol, 1 equiv.), 2-(5,6-difluoro-3-oxo- corrected against Fc/Fc+. Single crystals were grown through slow
2,3-dihydro-1H-inden-1-ylidene)malononitrile (208 mg, 0.4 mol, diffusion of ethanol to its chloroform solution. Low-energy inverse
4 equiv.) and pyridine (0.1 ml) were added in chloroform (5 ml). photoemission spectroscopy (LEIPS) measurement was performed
The reaction was stirred at 65 °C overnight. After cooling to room on a customized ULVAC-PHI LEIPS instrument with Bremsstrahlung
temperature, the solvent was removed and the residue was purified isochromatic mode. UV photoelectron spectroscopy measurement
by column chromatography to yield a dark-blue solid (262.4 mg, 76%). was performed by AXIS ULTRA DLD instrument of Kratos company.
1
H NMR (400 MHz, CDCl3) δ 9.19 (s, 2H), 9.05 (s, 2H), 8.25 (d, J = 6.9 Hz, The UV light source used is non-monochromatic He I, and the energy of
2H), 7.72 (m, 9.8 Hz, 4H), 7.20 (m, 4H), 7.13 (d, J = 2.6 Hz, 2H), 7.07 He I light source is 21.22 eV. The basic vacuum of the analysis chamber
(d, J = 7.4 Hz, 2H), 7.03 (d, J = 7.4 Hz, 2H), 6.96–6.85 (m, 6H), 6.63 is 3.0 × 10−8 Torr, and the bias voltage applied during the test is −9 V.
(dd, 4H), 4.86–4.57 (m, 4H), 3.18 (m, 4H), 2.63–2.25 (m, 8H), 2.13
(s, 2H), 1.86 (m, 4H), 1.32 (m, 48H), 0.88 (m, 7.3 Hz, 6H). 13C NMR Device fabrication and characterization
(101 MHz, CDCl3) δ 187.78, 160.21, 153.89, 147.46, 145.35, 142.23, 142.18, The device was equipped with a conventional construction of glass/ITO/
140.92, 140.86, 137.86, 136.15, 135.69, 135.36, 134.08, 133.78, 133.44, PEDOT:PSS/D:A/PDINN/Ag, where ITO is indium tin oxide, PEDOT:PSS
133.33, 132.51, 132.43, 130.94, 128.32, 128.30, 128.25, 128.09, 128.03, is poly(3,4-ethylenedioxythiophene):polystyrene sulfonate and PDINN
127.90, 127.82, 125.82, 125.78, 125.50, 123.32, 122.28, 116.54, 116.37, 116.05, is N,N′-bis{3-[3-(dimethylamino)propylamino]propyl}perylene-3,4,9,
115.88, 115.73, 115.48, 113.81, 77.35, 77.04, 76.72, 67.26, 55.74, 38.69, 38.24, 10-tetracarboxylic diimide. The weight ratio of D18:acceptor is 1:1.2,
35.72, 35.68, 32.09, 31.94, 31.66, 31.55, 31.50, 31.25, 30.51, 30.32, 29.92, and chloroform was applied as a solvent. The D18:acceptor blend
29.84, 29.71, 29.66, 29.58, 29.50, 29.38, 27.90, 25.75, 25.51, 22.72, 22.67, solution with a total concentration of 10 mg ml−1 was heated at 100 °C
19.21, 14.16, 11.45. MS (MALDI-TOF) calculated for C114H106F4N8O2S5 [M]+ till completely dissolved. Then the blend solution was spin-coated on
1,855.70, found 1,855.70. PEDOT:PSS, and then the blend was heated at 120 °C for 10 min. The
PDINN in ethyl alcohol (1 mg ml−1) was spin-coated on the active layer
Z8 by 3,000 rpm. The Ag (100 nm) was deposed onto the substrates by
Compound 6 (133.2 mg, 0.1 mmol, 1 equiv.), 2-(5,6-difluoro-3-oxo- vacuum evaporation. Shadow masks were used to define the OSC
2,3-dihydro-1H-inden-1-ylidene)malononitrile (23.1 mg, 0.1 mol, active area (0.04 cm2) of the devices. The current density–voltage (J–V)
1 equiv.), 2-(6,7-difluoro-3-oxo-2,3-dihydro-1H-cyclopenta[b] characteristics of unencapsulated photovoltaic devices were measured
naphthalen-1-ylidene)malononitrile (28.1 mg, 0.1 mol, 1 equiv.) and under N2 using a Keysight B2912A Precision Source/Measure unit.

Nature Energy
Article https://doi.org/10.1038/s41560-024-01557-z

The illumination of AM 1.5G (100 mW cm−2) was achieved by a XES- fitted with the restrained electrostatic potential method. Moreover,
70S1 solar simulator (SAN-EI Electric Co., Ltd., AAA grade, the missing torsion potentials were reparametrized according to DFT
70 mm × 70 mm photo-beam size). The EQE was determined using calculations. The simulations were carried out with three-dimensional
certified incident photon to current conversion efficiency equipment periodic boundary conditions; the leap-frog integrator with a time
from Enlitech. step of 1.0 fs was selected. A spherical cut-off of 1.2 nm for the sum-
mation of van der Waals interactions and short-range Coulomb inter-
Morphological characterization actions and the particle-mesh Ewald method for solving long-range
AFM images of the thin films were obtained on a NanoscopeIIIa AFM Coulomb interactions was used. These thin films were built and imi-
(Digital Instruments) operating platform in tapping mode. TEM obser- tated using the following procedure: (1) randomly placing molecules
vation was performed on JEOL 2200 FS at 160 kV accelerating voltage. in a large box to generate an initial geometry (D18:L8-BO, D18:Z7 and
The samples for electron microscopy were prepared by dissolving D18:Z8 thin films were obtained by adding 100 D18 molecules and 315
the PEDOT:PSS layer using water and transferring the floating active L8-BO, 251 Z7 and 258 Z8 molecules, respectively; D18:Z7:L8-BO and
layer to the TEM grids. The GIWAXS characterization of the thin films D18:Z8:L8-BO thin films were obtained by adding 100 D18 molecules,
was performed at the Advanced Light Source on the beamline of 7.3.3. 237 L8-BO molecules and extra 63 Z7 or 65 Z8 molecules according
Samples were prepared under device conditions on the Si/PEDOT:PSS to the experimental D:A weight ratio); (2) 5 ns of simulation at 600 K
substrates. GISAXS measurements were carried out with a Xeuss 2.0 and 100 bar to quickly bring molecules close together; (3) 10 ns of
SAXS laboratory beamline using a Cu X-ray source (8.05 keV, 1.54 Å) simulation at 600 K and 1 bar, then cooling down to 300 K in 3 ns; (4)
and a Pilatus3R 300K detector. The samples were prepared on sili- 50 ns of equilibration at 300 K and 1 bar. The velocity rescaling thermal
con wafer, and the incidence angle was 0.2°. The data were fitted by thermostat63 and the Berendsen barostat64 under the NPT ensemble
the Debye–Anderson–Brumberger model and a fractal-like network were applied to control the temperature and pressure, respectively.
model. TOF-SIMS measurement was conducted using a TOF-SIMS However, for the final 10 ns of equilibration, the Nosé–Hoover thermo-
5–100 instrument (ION-TOF GmbH), where a 1 keV Cs+ ion beam was stat65,66 and Parrinello–Rahman barostat67 were used to obtain better
used for erosion and a 30 keV Bi13+ pulsed primary ion beam was used equilibrium conformations.
for the analysis. The samples were prepared by spin-coating the active
layers on the PEDOT:PSS layer. The corresponding blend samples were Reporting summary
thermally annealed for 10 min. Further information on research design is available in the Nature
Portfolio Reporting Summary linked to this article.
FTPS-EQE and EQEEL measurements
Fourier transform photocurrent spectroscopy (FTPS)-EQE was meas- Data availability
ured using an integrated system (PECT-600, Enlitech), where the The authors declare all data supporting the findings of this study are
photocurrent was amplified and modulated by a lock-in instrument. available within the manuscript and Supplementary Information.
Electroluminescent spectroscopy (EQEEL) measurements were per- Source data are provided with this paper.
formed by applying external voltage/current sources through the
devices (ELCT-3010, Enlitech). EQEEL measurements were performed References
for all devices according to the optimal device preparation conditions. 1. Yao, H. et al. Molecular design of benzodithiophene-based
organic photovoltaic materials. Chem. Rev. 116, 7397–7457 (2016).
TPC, TPV and CE measurements 2. Guo, X. et al. Polymer solar cells with enhanced fill factors. Nat.
The TPC, TPV and CE measurements were performed on an all-in-one Photon. 7, 825–833 (2013).
characterization platform Paios developed by Fluxim AG. 3. Liang, Y. et al. For the bright future—bulk heterojunction polymer
solar cells with power conversion efficiency of 7.4%. Adv. Mater.
PLQY measurements 22, E135–E138 (2010).
PLQY measurements were conducted utilizing an FLS980 Fluorimeter, 4. Vohra, V. et al. Efficient inverted polymer solar cells employing
equipped with a LabSpheres integrating sphere and a near-infrared favourable molecular orientation. Nat. Photon. 9, 403–408 (2015).
photomultiplier tube (NIR-PMT, model R5509-72). Subsequent data 5. Yuan, J. et al. Enabling low voltage losses and high photocurrent in
analysis and calculation were carried out using a specialized PLQY fullerene-free organic photovoltaics. Nat. Commun. 10, 570 (2019).
software package. 6. Yuan, J. et al. Single-junction organic solar cell with over 15%
efficiency using fused-ring acceptor with electron-deficient core.
Quantum chemical calculations Joule 3, 1140–1151 (2019).
The geometry optimization of the molecules is performed by DFT cal- 7. Li, C. et al. Non-fullerene acceptors with branched side chains
culations at the b3lyp/6-311g(d,p) level. All the alkyl groups are replaced and improved molecular packing to exceed 18% efficiency in
by methyl groups to reduce the computational costs. The calculations organic solar cells. Nat. Energy 6, 605–613 (2021).
were conducted using the polarizable continuum model, taking the 8. Liang, Y. et al. Organic solar cells using oligomer acceptors for
dielectric constant for chloroform (ε = 4.7113) as reference. All the improved stability and efficiency. Nat. Energy 7, 1180–1190 (2022).
DFT calculations were carried out using the Gaussian 16 package56. 9. Baran, D. et al. Reducing the efficiency–stability–cost gap of
VMD (version 1.9.3) was used to plot the images of molecular orbitals organic photovoltaics with highly efficient and stable small
(isosurface 0.02) and dipole moments57. Multiwfn software is used for molecule acceptor ternary solar cells. Nat. Mater. 16, 363–369
the calculation of the orbital delocalization index and the analysis of (2017).
dipole moment of molecular fragments58. 10. Lin, Y. et al. An electron acceptor challenging fullerenes for
efficient polymer solar cells. Adv. Mater. 27, 1170–1174 (2015).
MD simulations 11. Zhu, L. et al. Single-junction organic solar cells with over 19%
All atomistic MD simulations were performed using the Gromacs–2019.3 efficiency enabled by a refined double-fibril network morphology.
software package59. For all the studied systems, the atom types and the Nat. Mater. 21, 656–663 (2022).
intra- and intermolecular interaction parameters were taken from 12. Fu, J. et al. 19.31% binary organic solar cell and low non-radiative
the general AMBER force field (GAFF)60. The atomic partial charges recombination enabled by non-monotonic intermediate state
were obtained by DFT calculations at the B3LYP/cc-PVTZ level61,62, and transition. Nat. Commun. 14, 1760 (2023).

Nature Energy
Article https://doi.org/10.1038/s41560-024-01557-z

13. Wang, D. H. et al. Sequential processing: control of nanomorpho­ 37. Gasparini, N., Salleo, A., McCulloch, I. & Baran, D. The role of the
logy in bulk heterojunction solar cells. Nano Lett. 11, 3163–3168 third component in ternary organic solar cells. Nat. Rev. Mater. 4,
(2011). 229–242 (2019).
14. He, Z. et al. Enhanced power-conversion efficiency in polymer 38. Günther, M. et al. Models and mechanisms of ternary organic
solar cells using an inverted device structure. Nat. Photon. 6, solar cells. Nat. Rev. Mater. 8, 456–471 (2023).
591–595 (2012). 39. Xu, X., Li, Y. & Peng, Q. Ternary blend organic solar cells:
15. Jiang, Y. et al. An alcohol-dispersed conducting polymer complex understanding the morphology from recent progress. Adv. Mater.
for fully printable organic solar cells with improved stability. Nat. 34, 2107476 (2022).
Energy 7, 352–359 (2022). 40. Zhou, Z. et al. High-efficiency small-molecule ternary solar cells
16. Meng, L. et al. Organic and solution-processed tandem solar cells with a hierarchical morphology enabled by synergizing fullerene
with 17.3% efficiency. Science 361, eaat2612 (2018). and non-fullerene acceptors. Nat. Energy 3, 952–959 (2018).
17. Zheng, Z. et al. Tandem organic solar cell with 20.2% efficiency. 41. Yu, R. et al. Design and application of volatilizable solid additives
Joule 6, 171–184 (2021). in non-fullerene organic solar cells. Nat. Commun. 9, 4645 (2018).
18. Holliday, S. et al. A rhodanine flanked nonfullerene acceptor for 42. Zhan, L. et al. Desired open-circuit voltage increase enables
solution-processed organic photovoltaics. J. Am. Chem. Soc. 137, efficiencies approaching 19% in symmetric-asymmetric molecule
898–904 (2015). ternary organic photovoltaics. Joule 6, 662–675 (2022).
19. Yang, Y. The original design principles of the Y series nonfullerene 43. Jiang, Y. et al. Suppressing electron–phonon coupling in
acceptors, from Y1 to Y6. ACS Nano 15, 18679–18682 (2021). organic photovoltaics for high-efficiency power conversion.
20. Yuan, J., Zou, Y. & Jun Yuan The history and development of Y6. Nat. Commun. 14, 5079 (2023).
Org. Electron. 102, 106436 (2022). 44. Vandewal, K., Tvingstedt, K., Gadisa, A., Inganas, O. & Manca, J. V.
21. Cui, Y., Xu, Y. & Hou, J. High efficiency and more functions bring On the origin of the open-circuit voltage of polymer-fullerene
a bright future for organic photovoltaic cells. Sci. Bull. 67, solar cells. Nat. Mater. 8, 904–909 (2009).
1300–1303 (2022). 45. Maurano, A. et al. Transient optoelectronic analysis of charge
22. Yi, J., Zhang, G., Yu, H. & He, Y. Advantages, challenges and carrier losses in a selenophene/fullerene blend solar cell. J. Phys.
molecular design of different material types used in organic solar Chem. C 115, 5947–5957 (2011).
cells. Nat. Rev. Mater. 9, 46–62 (2024). 46. Koster, L. J. A., Kemerink, M., Wienk, M. M., Maturová, K. &
23. Yao, J. et al. Quantifying losses in open-circuit voltage in solution- Janssen, R. A. J. Quantifying bimolecular recombination losses in
processable solar cells. Phys. Rev. Appl. 4, 014020 (2015). organic bulk heterojunction solar cells. Adv. Mater. 23, 1670–1674
24. Liu, S. et al. High-efficiency organic solar cells with low (2011).
non-radiative recombination loss and low energetic disorder. Nat. 47. Bartesaghi, D. et al. Competition between recombination and
Photon. 14, 300–305 (2020). extraction of free charges determines the fill factor of organic
25. Zhang, G. et al. Delocalization of exciton and electron solar cells. Nat. Commun. 6, 7083 (2015).
wavefunction in non-fullerene acceptor molecules enables 48. Wetzelaer, G. A. H., Kaap, N. J. V., der, Koster, L. J. A. &
efficient organic solar cells. Nat. Commun. 11, 3943 (2020). Blom, P. W. M. Quantifying bimolecular recombination in organic
26. Zhu, X., Zhang, G., Zhang, J., Yip, H.-L. & Hu, B. Self-stimulated solar cells in steady state. Adv. Energy Mater. 3, 1130–1134 (2013).
dissociation in non-fullerene organic bulk-heterojunction solar 49. Neher, D., Kniepert, J., Elimelech, A. & Koster, L. J. A. A new
cells. Joule 4, 2443–2457 (2020). figure of merit for organic solar cells with transport-limited
27. Chen, X.-K. et al. A unified description of non-radiative voltage photocurrents. Sci. Rep. 6, 24861 (2016).
losses in organic solar cells. Nat. Energy 6, 799–806 (2021). 50. Rivnay, J., Mannsfeld, S. C. B., Miller, C. E., Salleo, A. & Toney, M. F.
28. Eisner, F. D. et al. Hybridization of local exciton and Quantitative determination of organic semiconductor microstruc­
charge-transfer states reduces nonradiative voltage losses in ture from the molecular to device scale. Chem. Rev. 112, 5488–5519
organic solar cells. J. Am. Chem. Soc. 141, 6362–6374 (2019). (2012).
29. Almora, O. et al. Device performance of emerging photovoltaic 51. Wang, R. et al. Charge separation from an intra-moiety
materials (version 2). Adv. Energy Mater. 11, 2102526 (2021). intermediate state in the high-performance PM6:Y6 organic
30. Classen, A. et al. The role of exciton lifetime for charge generation photovoltaic blend. J. Am. Chem. Soc. 142, 12751–12759 (2020).
in organic solar cells at negligible energy-level offsets. Nat. 52. Polman, A., Knight, M., Garnett, E. C., Ehrler, B. & Sinke, W. C.
Energy 5, 711–719 (2020). Photovoltaic materials: present efficiencies and future challenges.
31. Li, S. et al. Refined molecular microstructure and optimized carrier Science 352, aad4424 (2016).
management of multicomponent organic photovoltaics toward 53. Katsidis, C. C. & Siapkas, D. I. General transfer-matrix method for
19.3% certified efficiency. Energy Environ. Sci. 16, 2262–2273 (2023). optical multilayer systems with coherent, partially coherent, and
32. Proctor, C. M., Kim, C., Neher, D. & Nguyen, T. Nongeminate incoherent interference. Appl. Opt. 41, 3978 (2002).
recombination and charge transport limitations in 54. Peng, W. et al. Reducing nonradiative recombination in perovskite
diketopyrrolopyrrole‐based solution‐processed small molecule solar cells with a porous insulator contact. Science 379, 683–690
solar cells. Adv. Funct. Mater. 23, 3584–3594 (2013). (2023).
33. Lu, L., Kelly, M. A., You, W. & Yu, L. Status and prospects for ternary 55. Haase, F. et al. Laser contact openings for local poly-Si-metal
organic photovoltaics. Nat. Photon. 9, 491–500 (2015). contacts enabling 26.1%-efficient POLO-IBC solar cells. Sol.
34. Zuo, L. et al. Dilution effect for highly efficient Energy Mater. Sol. Cells 186, 184–193 (2018).
multiple-component organic solar cells. Nat. Nanotechnol. 17, 56. Frisch, M. J. et al. Gaussian 16 Rev. A.03 (Gaussian, Inc., 2016).
53–60 (2022). 57. Humphrey, W., Dalke, A. & Schulten, K. VMD: visual molecular
35. Jiang, K. et al. Suppressed recombination loss in organic dynamics. J. Mol. Graph. 14, 33–38 (1996).
photovoltaics adopting a planar–mixed heterojunction 58. Lu, T. & Chen, F. Multiwfn: a multifunctional wavefunction
architecture. Nat. Energy 7, 1076–1086 (2022). analyzer. J. Comput. Chem. 33, 580–592 (2012).
36. Chen, T. et al. Compromising charge generation and recombi­ 59. Abraham, M. J. et al. GROMACS: high performance molecular
nation of organic photovoltaics with mixed diluent strategy for simulations through multi-level parallelism from laptops to
certified 19.4% efficiency. Adv. Mater. 35, 2300400 (2023). supercomputers. SoftwareX 1–2, 19–25 (2015).

Nature Energy
Article https://doi.org/10.1038/s41560-024-01557-z

60. Wang, J., Wolf, R. M., Caldwell, J. W., Kollman, P. A. & Case, D. A. microscopy measurement, and Y. Ouyang and C. Zhang for their
Development and testing of a general amber force field. J. Comput. help in TA measurements, and X. Chen, Q. Luo and C. Ma for their
Chem. 25, 1157–1174 (2004). help in MMP measurements and Y. Fu and X. Lu for their help in
61. Fox, T. & Kollman, P. A. Application of the RESP methodology in GISAXS measurements.
the parametrization of organic solvents. J. Phys. Chem. B 102,
8070–8079 (1998). Author contributions
62. Bayly, C. I., Cieplak, P., Cornell, W. & Kollman, P. A. A well-behaved X.Z. and F.L. conceived the study. Y.J. and F.L. fabricated the solar cells
electrostatic potential based method using charge restraints and performed the related measurements. S.S. performed theoretical
for deriving atomic charges: the RESP model. J. Phys. Chem. 97, calculations. R.X. and K.L. synthesized the NFAs Z7 and Z8. G.R. and
10269–10280 (1993). W.Z. performed the TA spectra analysis. X.Z. supervised the project.
63. Bussi, G., Donadio, D. & Parrinello, M. Canonical sampling through X.M. and Y.Y. performed MD simulation. The manuscript was mainly
velocity rescaling. J. Chem. Phys. 126, 014101 (2007). written by Y.J., S.S., F.L. and X.Z. All the authors contributed to the data
64. Berendsen, H. J. C., Postma, J. P. M., van Gunsteren, W. F., analysis and discussion.
DiNola, A. & Haak, J. R. Molecular dynamics with coupling to an
external bath. J. Chem. Phys. 81, 3684–3690 (1984). Competing interests
65. Nosé, S. A unified formulation of the constant temperature The authors declare that they have no competing interests.
molecular dynamics methods. J. Chem. Phys. 81, 511–519 (1984).
66. Hoover, W. G. Canonical dynamics: equilibrium phase–space Additional information
distributions. Phys. Rev. A Gen. Phys. 31, 1695–1697 (1985). Supplementary information The online version
67. Parrinello, M. & Rahman, A. Polymorphic transitions in single contains supplementary material available at
crystals: a new molecular dynamics method. J. Appl. Phys. 52, https://doi.org/10.1038/s41560-024-01557-z.
7182–7190 (1981).
68. Liu, W. et al. Theory-guided material design enabling high- Correspondence and requests for materials should be addressed to
performance multifunctional semitransparent organic photo­ Feng Liu or Xiaozhang Zhu.
voltaics without optical modulations. Adv. Mater. 34, 2200337
(2022). Peer review information Nature Energy thanks Yanming Sun,
69. Li, W. et al. A high-efficiency organic solar cell enabled by Hin-Lap Yip and the other, anonymous, reviewer(s) for their
the strong intramolecular electron push–pull effect of the contribution to the peer review of this work.
nonfullerene acceptor. Adv. Mater. 30, 17071170 (2018).
70. Cui, Y. et al. Single-junction organic photovoltaic cell with 19% Reprints and permissions information is available at
efficiency. Adv. Mater. 33, 2102420 (2021). www.nature.com/reprints.

Acknowledgements Publisher’s note Springer Nature remains neutral with regard to


This work was supported by the Beijing Natural Science jurisdictional claims in published maps and institutional affiliations.
Foundation (Z230019), the National Key R&D Program of China
(2019YFA0705900), the Strategic Priority Research Program of the Springer Nature or its licensor (e.g. a society or other partner) holds
Chinese Academy of Sciences (XDB0520202), the International exclusive rights to this article under a publishing agreement with
Partnership Program of the Chinese Academy of Sciences the author(s) or other rightsholder(s); author self-archiving of the
(027GJHZ2022036GC) and National Natural Science Foundation of accepted manuscript version of this article is solely governed by the
China (52225305, 22175187 and 22375119). The authors thank C. Zhu, terms of such publishing agreement and applicable law.
L. Meng and Y. Li for their help with the TPC, TPV and CE measurements,
and M. Xie and Z. Wei for their help with the FTPS-EQE and EQEEL © The Author(s), under exclusive licence to Springer Nature Limited
measurements, and X. Pu for his help with photo-induced force 2024

Nature Energy
nature research | reporting checklist for manuscripts on photovoltaics research
Corresponding author(s): Xiaozhang Zhu
Initial submission Revised version Final submission

Reporting Checklist For Solar Cell Manuscripts


In order to improve the reproducibility and transparency of manuscripts related to photovoltaic cells, we request that authors
consider the following points in the preparation of their manuscript. Please supply a response to the checklist alongside your
submitted manuscript, and ensure that the relevant responses are also provided in the main manuscript, methods section or
supplementary information as appropriate. The completed checklist will be shared with reviewers.

` Solar Cell Data


1. Dimensions
State the dimensions of the tested solar cells. The dimensions of the tested solar cells is 3.159 mm^2.
2. Current-voltage characterization
Does the manuscript supply current density - voltage The current-voltage (J-V) plots in forward is supplied. Both forward and backward
(J-V) plots in both forward and backward direction? scans were conducted, which yielded identical results.

Describe the voltage scan conditions for your devices. direction: forward; dwell times: 20 ms; speed: 0.015 V/s

Describe the test environment for your devices. We tested the devices at room temperature under N2 in the glove box.

If a preconditioning protocol was used before the


characterization, describe it.

Does the manuscript state whether the stability of the J-V The long-term stability of D18:Z8:L8-BO-, D18:L8-BO-, D18:Z8-based devices under
characteristic has been checked with a) stabilization of maximum power point (MPP) tracking were performed with continuous
photocurrent at maximum power point or b) maximum illumination in N2-filled chamber, as shown in Supplementary Figure 18.
power point data?
Please see ref. 7 for further details.

3. Hysteresis
Did you observe hysteresis or any other unusual No hysteresis or other unusual behaviour was observed during the characterization
behaviour during the characterization of the solar cells? of the solar cells.
4. Efficiency
Does the manuscript provides external quantum External Quantum Efficiency (EQE) data for the devices are provided in Figure 1e
efficiency (EQE) or incident photons to current efficiency and calculated short-circuit current densities are comparable to that derived from
(IPCE) data for the devices? J-V plots.

Was the integrated response under the standard The integrated response under the standard reference spectrum is compared to
reference spectrum compared to the response measured the response measured under the simulator.
under the simulator?
For tandem solar cells, does the manuscript provide the Describe where this information can be found in the text OR state why this
bias illumination intensity and the bias voltage used for characterization has not been performed/is not applicable.
each subcell?

5. Calibration
State the light source and the reference cell or sensor Light source: solar simulator (XES-70S1, 7 × 7 cm^2 beam size; SAN-EI Electric Co.
used for the characterization. Ltd.) coupled with AM 1.5G solar spectrum filters; Reference cell: SRC-1000-TC-QZ,
2 × 2 cm^2; VLSI Standards Inc.
The reference cell was calibrated and certified.
Was a spectral mismatch calculation between the The reference cell was calibrated and certified before the test.
June 2017

reference cell and the devices under test performed?

6. Mask/aperture
State the size of the mask/aperture used during testing. A mask of 3.159 mm^2 was used in NIM.

Does the measured short-circuit current density of the No, a JSC of 27.0 mA cm-2 was obtained using a mask of 3.159 mm^2.
devices vary with the mask/aperture area?

1
7. Performance certification
Name any independent certification laboratories that The performance of the devices is confirmed from National Institute of Metrology,

nature research | reporting checklist for manuscripts on photovoltaics research


have confirmed the photovoltaic performance of your China (NIM). The detailed data is provided in Figure 1f and Supplementary
devices. information.

You have provided a copy of any certificate(s) in the Supplementary Information.

8. Statistics
How many solar cells were tested in the study? We tested more than 30 devices in this study.

Does the manuscript include a statistical analysis of the We have a statistical analysis of the performance included in the manuscript. Table
performance? 1 shows the best and average efficiencies with standard deviation.

9. Long-term stability analysis


Clearly describe the type of stability analysis performed The long-term stability of optimal D18:Z8:L8-BO devices under maximum power
and the bias conditions used during this characterization. point (MPP) tracking was performed with continuous illumination in N2-filled
chamber, as shown in Supplementary Figure 18.

` Further reading
1. Shrotriya, V. et al. Accurate measurement and characterization of organic solar cells. Adv. Funct. Mater. 16, 2016–2023 (2006).
2. Dennler, G. et al. The value of values. Mat. Today 10, 56 (2007).
3. Cravino, A., Schilinsky, P. & Brabec, C. J. Characterization of organic solar cells: the importance of device layout. Adv. Funct. Mater. 17, 3906–3910 (2007).
4. Reese, M. O. et al. Consensus stability testing protocols for organic photovoltaic materials and devices. Sol. Energ. Mat. Sol. C 95, 1253–1267 (2011).
5. Snaith H. .J. The perils of solar cell efficiency measurements. Nat. Photon. 6, 337–340 (2012).
6. Luber, E. J. & Buriak, J. M. Reporting performance in organic photovoltaic devices. ACS Nano 7, 4708–4714 (2013).
7. Snaith, H. J. et al. Anomalous hysteresis in perovskite solar cells. J. Phys. Chem. Lett. 5, 1511–1515 (2014).
8. Grätzel M. The light and shade of perovskite solar cells. Nat. Mat. 13, 838–842 (2014).
9. Zimmermann E. et al. Erroneous efficiency reports harm organic solar cell research. Nat. Photon. 8, 669–672 (2014).
10. Beard M.C., Luther J.M. & Nozik A.J. The promise and challenge of nanostructured solar cells. Nat. Nanotech. 9, 951–954 (2014).
11. Timmreck, R. et al. Characterization of tandem organic solar cells. Nat. Photon. 9, 478–479 (2015).

A number of international committees develop industry standards on the characterization of photovoltaic technologies (for example ASTM-E44 and
IEC-TC 82), which can provide guidance for academic research.

June 2017

You might also like