Qiao
He‡
a,
Flurin D.
Eisner‡
b,
Drew
Pearce
b,
Thomas
Hodsden
a,
Elham
Rezasoltani
b,
Daniel
Medranda
b,
Zhuping
Fei
c,
Jenny
Nelson
*b and
Martin
Heeney
*a
aDepartment of Chemistry and Centre for Processable Electronics, Imperial College London, London W12 0BZ, UK. E-mail: m.heeney@imperial.ac.uk
bDepartment of Physics and Centre for Processable Electronics, Imperial College London, London SW7 2AZ, UK. E-mail: jenny.nelson@imperial.ac.uk
cInstitute of Molecular Plus, Tianjin Key Laboratory of Molecular Optoelectronic Science, Tianjin University, Tianjin 300072, P. R. China
First published on 17th November 2020
In this work, we designed and synthesized two novel perylene diimide (PDI) tetramers based on a tetrathienylethene core, named TTE-PDI4 and FTTE-PDI4, and investigated their application as non-fullerene acceptors for organic photovoltaics. The free rotation of PDIs and adjacent thiophene units renders TTE-PDI4 with a highly twisted molecular geometry. The ring fusion of TTE-PDI4 yields FTTE-PDI4, a more rigid molecule with increased intramolecular stacking. Interestingly, TTE-PDI4 and FTTE-PDI4 possess similar energy levels but very different UV-Vis absorptions, with the latter showing strong broad-band absorption with multiple sharp peaks in the 300–600 nm region. Through time-dependent density functional theory (TD-DFT) calculations, we show that this broad absorption spectrum in FTTE-PDI4 arises from the combination of multiple bright transitions in the visible region with a strong vibronic progression, tentatively assigned to the dominant CC stretching mode. TTE-PDI4, despite having a lower energy absorption onset, shows weaker absorption at long wavelengths. Due to its higher absorption as well as its increased rigidity, FTTE-PDI4 shows a higher photocurrent and hence a higher power conversion efficiency (PCE), of 6.6%, when blended with the polymer donor PFBDB-T than TTE-PDI4 based blends (PCE of 3.8%). The greater rigidity of FTTE-PDI4 is likely to contribute to the good fill factor of the blend devices. Potential for further improvement through reducing voltage losses is identified.
With regards to the various types of electron-withdrawing units to construct NFAs, perylene diimide (PDI) has been among the earliest studied and most promising.10–15 Pioneering studies on the PDI monomer and its derivatives found that the tendency of the coplanar PDI monomers to strongly aggregate resulted in micrometer-sized phase-separated domains in bulk-heterojunction blends, leading to low OPV device efficiencies.16,17 Disrupting molecular planarity by intramolecular twisting has been found to be a useful approach to suppress the formation of large aggregates and improve performance.18–21 By connecting two PDI monomers via a spacer or linking more PDIs with a central core, PDI dimers, trimers and tetramers have been prepared with a twisted 3D or quasi-3D geometry, resulting in improved OPV device performance. The spacers can be simply a single or double bond, an aromatic ring or a larger conjugated central core.22–27 Although the disruption of planarity is generally beneficial, excessive molecular twisting can be detrimental to charge transport and, consequently, high structural non-planarity of PDI-based acceptors can result in reduced device performance.27–29 To tackle this issue, ring fusion has been proposed in several studies to finely tune the intermolecular π–π stacking.30–34 Concomitantly, evidence also suggests that ring fusion blue-shifts absorption onset, enhances absorption strength, and up-shifts the lowest unoccupied molecular orbital (LUMO) energy level, which can be beneficial for overall OPV device performance.30,35 Through the employment of geometry twisting and aromatic ring fusion, currently the power conversion efficiency (PCE) of the best performing PDI-based NFAs have been enhanced to over 10%.35,36
We were particularly interested by the observations that ring fusion can lead a blue-shifted absorption, since wide band gap NFA absorbers with a strong absorption in the green to near-UV region are of interest for the front cell in tandem devices or for smart window applications.37,38 Ring fusion in the bay region of the PDI has typically been achieved by an oxidative ring closure with an electron rich heterocycle like thiophene or selenophene,30,35 and in order to prevent excessive aggregation it should be combined with the twisting approach. As such, we identified tetrathienylethene (TTE) as a potentially interesting central core for the attachment of four PDI arms, combining the requisite thienyl groups for ring fusion with a non-planar geometry. TTE has previously been primarily investigated in the field of aggregation-induced emission (AIE).39,40 Although the emissive properties are not the focus of this paper, we note that other AIE cores like tetraphenylethene (TPE) have been used for the attachment of PDI arms, although the resulting TPE-PDI4 was not able to undergo ring fusion with the PDI.25 Here we report the synthesis of two new non-fused and fused PDI tetramers (TTE-PDI4 and FTTE-PDI4) based on the tetrathienylethene core and investigate their performance in organic photovoltaic devices. For the non-fused TTE-PDI4, freely rotating single bonds between thiophene and the central ethene, as well as thiophene and the PDI, induces strong steric hindrance leading to a highly twisted molecular structure. In comparison, DFT calculations suggest the fused tetramer, FTTE-PDI4, displays a more ordered geometry with the interlocking of thiophenes and PDI units. The ring fusion leads to an up-shifted LUMO level, and an intense and extremely broad absorption spectrum at visible to near-UV wavelengths. These are found by time-dependent DFT to arise from multiple bright electronic transitions. These beneficial changes contribute to higher short-circuit current density (JSC) and improved power conversion efficiency from 3.8% (TTE-PDI4) to 6.6% (FTTE-PDI4) in solar cell devices, showing that careful design of the central core on PDI tetramers can lead to promising wide band gap acceptors.
TTE-PDI4 and FTTE-PDI4 were purified by a combination of column chromatography followed by preparative recycling gel-permeation chromatography (GPC) to remove non-tetrameric by-products (see ESI†). The identities of TTE-PDI4 and FTTE-PDI4 were confirmed by MALDI-TOF mass spectrometry (Fig. S25 and S27, ESI†). Room temperature 1H NMR spectra were broad and poorly resolved, similar to other PDI tetramers,15,34,35 as a result of the restricted rotation of the PDI subunits and presence of multiple conformers. Heating FTTE-PDI4 solutions at 40 °C, 80 °C and 120 °C resulted in improved resolution (Fig. S24, ESI†), particularly of the protons on the outer shell of the molecule which were clearly resolved as sets of doublets. The inner protons nearer the TTE core were less well resolved, possibly due to more restricted movement.
As such, it is difficult to completely rule out the presence of some incompletely fused material, although the significant change in optical properties (vide infra) suggests that this would only be a minor impurity if present.
Both compounds are readily soluble in common organic solvents such as chloroform and chlorobenzene at room temperature. The structures of the two PDI tetramers were characterized and confirmed by MALDI-TOF MS, 1H NMR, and 13C NMR. Both TTE-PDI4 and FTTE-PDI4 exhibited good thermal stability, with thermal decomposition temperatures (Td) of 320 °C and 375 °C, respectively, measured by thermogravimetric analysis (TGA, at 5% weight loss, Fig. S1, ESI†). In addition, in the range of 20 °C to 280 °C, no obvious crystallization or melting transitions could be observed for both PDI tetramers, indicating their weak crystallization tendency (Fig. S2, ESI†).
In order to investigate the effect of the different structural geometries of the two molecules on their absorption profile, we additionally carried out time-dependent DFT (TD-DFT) calculations. Fig. 1B and C thus show the electronic and vibrational structure calculations, respectively. The results firstly show that we expect TTE-PDI4 to have a smaller band gap than FTTE-PDI4 by around 0.1 eV due to a weak low energy transition that appears in the extinction spectrum as a long and shallow tail. Secondly, FTTE-PDI4 shows strong absorption at low wavelengths due to several bright electronic transitions in the visible, in agreement with previous studies,42 while lacking the weak low energy transition that creates the absorption tail in TTE-PDI4. Thirdly, vibrational structure calculations show that there is a strong vibrational band at 1600 cm−1 that is present in both molecules. The mode is assigned to the CC bond stretch within the PDI units, as shown in Fig. S5 (ESI†). When a mode of this frequency is used to model the vibronic progression of the extinction spectra of the two molecules (Fig. 1D) a spectrum with multiple peaks across the visible results. We note that broad absorption features at low wavelengths have often been reported in the absorption spectra of PDI-based molecules, in particular in fused PDIs;35,43 our calculations suggest that these features may result from the combination of multiple close lying electronic transitions with a strong vibronic progression.
λ max,sol (nm) | α max,sol (M−1 cm−1) | λ max,film (nm) | α max,film (cm−1) | E g,opt (eV) | E g,electro (eV) | LUMOc (eV) | HOMOc (eV) | |
---|---|---|---|---|---|---|---|---|
a Determined by the absorption onset of the films. b Electrochemical band gap, determined by (LUMO–HOMO). c Estimated from the reduction and oxidation onsets of the CV curves. | ||||||||
TTE-PDI4 | 525 | 1.19 × 105 | 530 | 3.55 × 104 | 1.90 | 2.20 | −3.74 | −5.94 |
FTTE-PDI4 | 429 | 1.41 × 105 | 424 | 3.77 × 104 | 2.00 | 2.29 | −3.68 | −5.97 |
In addition, the maximum extinction coefficients of FTTE-PDI4 in solution and in film were calculated to be 1.41 × 105 M−1 cm−1 and 3.77 × 104 cm−1, respectively, higher than those of TTE-PDI4 (1.19 × 105 M−1 cm−1 and 3.55 × 104 cm−1). The minor differences in the absorption spectra between solution and film further suggest that intermolecular packing is weak for both molecules. This is further confirmed by absence of any change in solution absorption measurements as a function of concentration (Fig. S10, ESI†). This additionally suggests that the breadth of the absorption of FTTE-PDI4, at over 50% of maximum intensity from 300 to 600 nm, must be due to the multiple bright electronic states of the molecules identified by TD-DFT rather than due to intermolecular interactions.
The electrochemical properties of these two compounds were investigated by cyclic voltammetry of thin films, with the LUMO/HOMO levels of TTE-PDI4 and FTTE-PDI4 (Fig. 2C) estimated to be similar at −3.74/−5.94 eV and −3.68/5.97 eV, respectively from the onset of the reduction and oxidation peaks. The slight upshift of the LUMO level upon ring fusion is in agreement with the DFT calculations and with other studies on fused PDIs.30,34,35
Reasonably high efficiencies (∼5%) were achieved with PTB7-Th, PBDB-T and PFBDB-T as the donor (Table S1, ESI†), with devices with PFBDB-T showing the highest efficiency; we therefore selected this polymer to optimize fabrication conditions with both TTE-PDI4 and FTTE-PDI4 blends. The optimized photovoltaic parameters and typical J–V characteristics of the PFBDB-T:TTE-PDI4 and PFBDB-T:FTTE-PDI4 devices are presented in Fig. 3A and Table 2. The best performance with PFBDB-T:FTTE-PDI4 devices was achieved with a 1:2 blend ratio, 2% 1-chloronaphthalene (CN) as solvent additive and annealing at 140 °C, showing an average PCE of 6.4% (best 6.6%) with a JSC of 10.9 mA cm−2, an open-circuit voltage (VOC) of 1.0 V, and a FF of 0.59. The relatively large acceptor fraction required for maximum device performance is in agreement with weak intermolecular packing of the molecules (evidenced by the absorption data above) and the low crystallinity (evidenced by featureless DSC spectra). Non-crystalline acceptor molecules generally need to be added in greater excess than crystalline acceptors to achieve percolating networks and phase-pure domains for charge separation. Interestingly, a high PCE was also maintained in as-cast cells without additive and thermal annealing, with such devices showing a lower JSC but higher VOC and FF of 1.02 V and 0.63, respectively. We note, also, that the highest FF (0.64) is achieved in a 1:1.5 blend ratio, which is amongst the highest reported for PDI-based acceptors. The PFBDB-T:TTE-PDI4 based active layers were also systematically optimized under various conditions, with a donor:acceptor ratio of 1:1.5, no additive and thermal annealing at 140 °C leading to the optimized average PCE of the TTE-PDI4 based OPVs of 3.8%, with inferior current density and FF to the FTTE-PDI4 based OPV devices. A similar VOC was observed for TTE-PDI4 and FTTE-PDI4 based OPVs, in spite of the slightly higher-lying LUMO levels for the latter. Moreover, external quantum efficiencies (EQEs) of the optimal devices were measured to verify the JSC values obtained from the J–V measurement (Fig. 3B). The JSC values integrated from EQE are 7.4 and 11.9 mA cm−2, both in less than 10% deviation from JSC measured from the TTE-PDI4 and FTTE-PDI4 based OPV devices. The PFBDB-T:FTTE-PDI4 devices exhibit stronger photo-response in the spectral range between 350 and 600 nm, with a higher maximum EQE of 72% than the PBDB-T:TTE-PDI4 devices (EQEmax = 56%), in agreement with the stronger absorption of FTTE-PDI4 compared to TTE-PDI4 in that region. This indicates FTTE-PDI4 contributes more to the charge generation and JSC than TTE-PDI4 as electron acceptor in their corresponding blends.
Acceptor | V OC (V) | J SC (mA cm−2) | FF (%) | PCEa (%) | Calc. JSC (mA cm−2) | V OC,rad (V) | ΔVOC,nrad (V) |
---|---|---|---|---|---|---|---|
a The values in parentheses are the average PCEs. The average values are collected from >12 independent devices. | |||||||
TTE-PDI4 | 0.99(0.99) | 7.5(7.3) | 51(50) | 3.8(3.6) | 8.2 | 1.35 | 0.36 |
FTTE-PDI4 | 1.0(1.0) | 10.9(10.9) | 60(59) | 6.6(6.4) | 11.9 | 1.38 | 0.38 |
Using the measured EQE edge (Fig. 3C), we then calculated the photovoltaic band gap energy (EPVg), using the method proposed by Rau,51 for both blends to be 1.84 eV. These values suggested high voltage losses (EPVg/q − VOC) of 0.85 and 0.84 V for PFBDB-T:TTE-PDI4 and PFBDB-T:FTTE-PDI4 devices, respectively, where q is electronic charge. In order to better understand the origin of these high losses, we therefore performed a voltage-loss analysis of the optimized blends by employing electroluminescence (EL) spectra to extend the measured sub-bandgap EQE spectra (Fig. 3C), using the reciprocity relation between light absorption and light emission,52,53 to calculate the open-circuit voltage in the radiative limit (VOC,rad). VOC,rad is the ideal VOC that a device can achieve when there is only radiative recombination, with the difference between VOC,rad and the measured VOC of the device then being defined as the voltage loss due to non-radiative recombination, ΔVOC,nrad. The VOC,rad and ΔVOC,nrad values for the optimized devices are shown in Table 2 and we also note that the maximum ideal VOC in the Shockley–Queisser limit (VOC,sq) achievable by a device with the measured band gap of 1.84 eV is 1.54 V.
We firstly draw attention to the fact, then, that the difference between VOC,sq and VOC,rad is large for both blends, which suggests that losses due a broad absorption edge are large; this is expected given the large offset between the LUMOs of the donor (−3.29 eV) and the acceptors (−3.74 eV and −3.68 eV), and suggests that the VOC could be significantly improved by using a donor with a lower lying LUMO level. Secondly, we note that both blends show similarly high ΔVOC,nrad of 0.36 V and 0.38 V for PFBDB-T:TTE-PDI4 and PFBDB-T:FTTE-PDI4, respectively, with the higher value for the FTTE-PDI4 blend meaning that the VOC is not larger than that of the TTE-PDI4 blend despite the larger optical gap. An earlier study identified several molecular factors that could reduce the nonradiative voltage loss in solar cells,54 among which increased molecular rigidity, expressed as reduced reorganization energy, is one. Interestingly, whilst the higher molecular rigidity leading to the smaller reorganization energy of FTTE-PDI4 appears to facilitate charge transport in the blend (manifested through the higher FF), non-radiative recombination does not appear to be suppressed. It is therefore likely that the voltage losses are influenced also by microstructural properties of the blend films and not only by molecular parameters.
To confirm this hypothesis, atomic-force-microscopy (AFM) images were measured for both blends, as shown in Fig. 4. Both films display nanoscale phase separation with fine and small domain size, suggesting their twisted molecular geometry lead to the formation of favorable morphology for OPV devices. The root-mean-square (RMS) roughness are 3.0 nm and 7.6 nm for the PFBDB-T:TTE-PDI4 and PFBDB-T:FTTBPDI4 blends, respectively. As expected, the PFBDB-T:TTE-PDI4 blend shows strong intermixing and a smoother surface morphology, with fiber-like aggregates observed in the film, whilst the PFBDB-T:FTTEPDI4 blend shows more aggregated domains of roughly 25–30 nm in length, possibly due to the larger excess of acceptor in the optimized blend compared to in PFBDB-T:TTE-PDI4. This is likely to lead to a larger degree of disorder in the blend, which is consistent with the slightly larger ΔVOC,nrad observed in this blend, although further measurements would be required to confirm this.
Fig. 4 AFM height images (2 μm × 2 μm) of the PFBDB-T:TTE-PDI4 (left) and PFBDB-T:FTTBPDI4 (right) blend films. The topographical height difference (Δz) are 21 nm and 51 nm, respectively. |
The molecular packing behaviour of compounds TTE-PDI4, FTTE-PDI4, PFBDB-T and their blend films were also investigated by X-ray diffraction on drop-cast films. As shown in Fig. S12 (ESI†), a broad π–π stacking peak was observed for neat PFBDB-T film, with a maxima at 2θ = 24.26°. After annealing at 160 °C for 10 min, the film not only exhibited a π–π stacking at a similar position, but also weak diffraction peaks at 2θ = 12.72°, 14.14°, 18.60° and 21.68°, suggestive of enhanced order. The enhanced order after thermal annealing is in agreement with our previous GIWAXS results on thin-films.6 TTE-PDI4 and FTTE-PDI4 exhibited no diffraction peaks, indicating their low crystallinity in thin films, which was also consistent with DSC data. The optimal blend films of PFBDB-T:TTE-PDI4 and PFBDB-T:FTTE-PDI4 also showed no diffraction peaks from their drop cast blends after annealing at 140 °C for 10 min, suggesting blending may inhibit polymer crystallisation somewhat, with the caveat that any crystallites present may not be readily observed with the lab based diffractometer used in these measurements.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d0tc04110f |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2020 |