Elifnaz Sağlamkayaa,
Mohammad Saeed Shadabrooa,
Nurlan Tokmoldinab,
Tanner M. Melodyc,
Bowen Suna,
Obaid Alqahtanicd,
Acacia Pattersonc,
Brian A. Collinsc,
Dieter Nehera and
Safa Shoaee*ab
aInstitute of Physics and Astronomy, University of Potsdam, Karl-Liebknecht-Str. 24-25, 14476 Potsdam-Golm, Germany. E-mail: shoai@uni-potsdam.de
bPaul-Drude-Institut für Festkörperelektronik Leibniz-Institut im Forschungsverbund Berlin e.V., Hausvogteiplatz 5-7, 10117 Berlin, Germany
cDepartment of Physics and Astronomy, Washington State University, 100 Dairy Road, Pullman, WA 99164, USA
dDepartment of Physics, Prince Sattam bin Abdulaziz University, Alkharj, 11942, Kingdom of Saudi Arabia
First published on 1st August 2024
All-small molecule (ASMs) solar cells have great potential to actualize the commercialization of organic photovoltaics owing to their higher solubility, lesser batch-to-batch variety and simpler synthesis routes compared to the blend systems that utilize conjugated polymers. However, the efficiencies of the ASMs are slightly lacking behind the polymer: small molecule bulk-heterojunctions. To address this discrepancy, we compare an ASM blend ZR1:Y6 with a polymer:small molecule blend PM7:Y6, sharing the same non-fullerene acceptor (NFA). Our analyses reveal similar energetic offset between the exciton singlet state and charge transfer state (ΔES1–CT) in ZR1:Y6 and PM7:Y6. In comparison to the latter, surprisingly, the ZR1:Y6 has noticeably a stronger field-dependency of charge generation. Low charge carrier mobilities of ZR1:Y6 measured, using space charge limited current measurements, entail a viable explanation for suppressed charge dissociation. Less crystalline and more intermixed domains as observed in the ZR1:Y6 system compared to polymer:Y6 blends, makes it difficult for NFA to form a continuous pathway for electron transport, which reduces the charge carrier mobility.
New conceptsThis study investigates factors influencing free charge generation, decoupling the effect of the energetic landscape from morphology. We compare an all-small-molecule system, ZR1:Y6, with polymer-small-molecule systems PM7:Y6 and PM6:Y6. PM7:Y6 and ZR1:Y6 show similar energetic offsets, whereas PM6:Y6 has a higher offset compared to the other two. Despite the similar offsets, the field dependency of charge generation differs in ZR1:Y6. Remarkably, across all three systems, the activation energy for charge generation remains consistent, yet the dissociation rate coefficient is notably lower in the ZR1:Y6 blend. Our findings highlight the crucial role of structural order and π–π stacking in polymer blends, enhancing charge carrier mobility and reducing recombination losses. This morphological influence is starkly evident in the ZR1:Y6 blend, where, despite having lower energetic disorder, the system is still characterized by lower mobility and higher field dependency of charge generation due to less crystalline and more intermixed domains. Our work suggests that morphological engineering can adjust the minimum required donor/acceptor energetic offset for efficient charge generation. |
2,2′-((2Z,2′Z)-((12,13-bis(2-ethylhexyl)-3,9-diundecyl-12,13-dihydro-[1,2,5]thiadiazolo[3,4 e]thieno[2′′,3′′:4′,5′]thieno[2′,3′:4,5]pyrrolo[3,2-g]thieno[2′,3′:4,5]thieno[3,2-b]indole-2,10-diyl)bis(methanylylidene))bis(5,6-difluoro-3-oxo-2,3-dihydro-1H-indene-2,1-diylidene))dimalononitrile (Y6) is a ADA′DA (A = acceptor unit, D = donor unit) type NFA molecule that has a curved geometric conformation. The shape of the molecule enables a 3D network morphology in films where the D and A units of neighboring Y6 molecules stack on top of each other. Consequently, charge transport in Y6 is ambipolar and efficient. In addition, Y6 excitons are delocalized over more than one molecule which reduces the exciton binding energy (Eb).15–17 Owing to a smaller than typical Eb in Y6, bulk charge photogeneration, whilst not too efficient is observed in the single component Y6 devices.18,19 On the other hand, bulk heterojunction (BHJ) blends of Y6 with poly[(2,6-(4,8-bis(5-(2-ethylhexyl-3-fluoro)thiophen-2-yl)-benzo[1,2-b:4,5-b′]dithiophene))-alt-(5,5-(1′,3′-di-2-thienyl-5′,7′-bis(2-ethylhexyl)benzo[1′,2′-c:4′,5′-c′]dithiophene-4,8-dione)] (PM6) achieved near unity charge generation yields and field and temperature independent charge dissociation, despite of the small D/A energy offset.20–22 Morphological studies on the most efficient polymer:Y6 blends revealed the polymers do not disturb the 3D networks in Y6 domains and both the donor and the Y6 aggregates adapt a face-on alignment to the substrate.23–25 Spatial confinement of the opposite charges in the donor and Y6 prevents the fast recombination of the free charges.19 Moreover, polymer chains of the donor materials provide conductive channels between different Y6 domains, enabling high charge carrier mobilities.26 Favorable morphologies in the polymer:Y6 blends enable fast charge dissociation rates, concomitantly suppressing free charge recombination.27 Obtaining equally optimal morphologies with efficient percolation pathways in ASM blends is more challenging. ASM blends can suffer from large crystalline domains due to over aggregation of one or both of the components.1,12,28
Herein we show a systematic comparison between an ASM blend based on 5,5′-[[4,8-bis[5-(2-butyloctyl)-2-thienyl]dithieno[2,3-d:2′,3′-d′]benzo[1,2-b:4,5-b′]dithiophene-2,6-diyl]bis([3′-octyl[2,2′-bithiophene]-5′′,5-diyl)methylidyne ]]bis[3-hexyl-2-thioxo-4-thiazolidinone] (ZR1) with polymer-small molecule blends based on: poly[(2,6-(4,8-bis(5-(2-ethylhexyl-3-chloro)thiophen-2-yl)-benzo[1,2-b:4,5-b′]dithiophene))-alt-(5,5-(1′,3′-di-2-thienyl)-5′,7′-bis(2-ethylhexyl)benzo[1′,2′-c:4′,5′-c′]dithiophene-4,8-dione)] (PM7) and PM6, all using the same Y6 NFA. As expected, ZR1:Y6 exhibits low energetic disorder compared to the polymer:Y6 blends. Despite the low disorder, low electron mobility is measured for ZR1:Y6. From temperature dependent electroluminescence (T-EL) measurements, the ΔES1−CT energy offset is estimated to be similar for both PM7:Y6 and ZR1:Y6 blends, which sits moderately smaller than the PM6:Y6 offset. The largest free charge recombination coefficient is measured for ZR1:Y6, followed by PM7:Y6, while PM6:Y6 exhibits the slowest recombination. A slower dissociation rate constant (kd) is estimated for the ASM ZR1:Y6 blend with a slight bias-dependency in charge generation, whereas PM7:Y6 demonstrates bias-independent charge generation despite having similar energetics. Furthermore, same temperature dependency of kd for all three blends hints at a similar activation energy for CT dissociation. Our morphological investigation shows suppressed π–π stacking of the Y6 aggregates as well as more intermixed domains in the ZR1:Y6 blend. Therefore, we assign the lower kd of the ASM system to the suppressed mobility and subsequently to its unfavorable molecular orientation and low π–π stacking. Our study exhibits a correlation between the charge carrier mobility, charge recombination and efficient charge generation and addresses the morphological roots of challenges in the ZR1:Y6 system.
VOC [V] | JSC [mA cm−2] | Jint [mA cm−2] | FF [%] | PCE [%] | Charge generation efficiency (%) | |
---|---|---|---|---|---|---|
ZR1:Y6 | 0.883 ± 0.003 | 22.6 ± 1.0 | 21.5 | 68.6 ± 0.7 | 14.2 ± 0.4 | 90 |
PM7:Y6 | 0.869 ± 0.002 | 25.6 ± 0.2 | 24.4 | 71.8 ± 0.5 | 16.3 ± 0.2 | 95 |
PM6:Y6 | 0.838 ± 0.006 | 24.9 ± 0.5 | 24.7 | 74.6 ± 1.6 | 15.8 ± 0.3 | 99 |
The bias-dependence of charge generation is measured with time-delayed collection field (TDCF) plotted in Fig. 1a–c (right axis), shown together with the J–V (left axis) curves extending to a large negative bias range. A large reverse bias of −6 V is selected to ensure saturation of the current. The J–V curve of PM6:Y6 aligns with the TDCF collected charge as a function of bias, and shows no bias-dependency (Fig. 1a). For PM7:Y6 both TDCF and J–V data show the same bias-dependency in the negative biases, (Fig. 1b) although in the positive range, J–V curve displays a higher bias-dependency. ZR1:Y6 exhibits the highest field dependency of charge generation (Fig. 1c and Fig. S3, ESI†) and the lowest FF. Charge generation efficiencies of blends calculated from the ratio of collected charge at JSC to that at −6V, are tabulated as a measure of the bias-dependency. We select JSC to calculate the charge generation efficiency rather than a value close to VOC because recombination of generated charges with dark injected charges decreases the collected charge in forward biases.31 The results are presented in Table 1.
From the convolution of the EQEPV and blackbody radiation spectrum, we calculate the same radiative VOC limit (VOCrad) of 1.09 V for all systems, with the EQEPV onsets aligning with the Y6 absorption. Moving on to the emission properties, we are able to calculate the non-radiative VOC losses (ΔVOCnrad) of the devices from the electroluminescence quantum yield (ELQY) (Fig. 1e) with qΔVOCnrad = −kBTln(ELQY), where q is the elementary charge, kB is the Boltzmann constant and T the temperature in Kelvin. The largest ΔVOCnrad is calculated for the PM6:Y6 (0.28 V), followed by the PM7:Y6 (0.25 V) and the lowest value of 0.24 V is estimated for the ASM ZR1:Y6 system. ΔVOCnrad is affected both from energetic disorder and the D/A energy offsets.32,33 In order to study the D/A energetics, we consider the energy offset between the singlet S1 and the CT states. For the blends with sufficiently small ΔES1−CT, electroluminescence is almost entirely dominated by the decay of the reformed of S1 exciton via transfer of the charges from CT back to S1.34 Temperature dependence of the ELQY reveals the energy barrier from the CT to S1 state, denoted as ΔES1−CT.21,30 The estimated ΔES1−CT values for PM7:Y6 and ZR1:Y6 are the same (64 and 61 meV respectively) (Fig. 1f). The literature value for PM6:Y6 (114 meV)21 is larger than both PM7:Y6 and ZR1:Y6, a finding consistent with the lower value of VOC observed in PM6:Y6. The larger ΔVOCnrad in PM6:Y6 may originate from a correspondingly higher barrier of ΔES1−CT resulting in a lower exciton reformation efficiency of 0.06% compared to PM7:Y6 and ZR1:Y6 (0.2% and 0.37%, respectively) at room temperature. (Table S1, ESI†) Reformation efficiency of the blend can be estimated by relating the ELQY of the blend to the photoluminescence quantum efficiency (PLQY) of the neat acceptor (photoluminescence spectra of the films are shown in Fig. S5, ESI†).21
ΔVOCnrad has been suggested to exhibit a correlation with the CT energetic disorder which can be related to LUMO disorder (σLUMO) and HOMO disorder (σHOMO) .8 Thus we estimate the static energetic disorder (σS) of the HOMO/LUMO levels and the carrier mobility from the transport properties. Charge carrier mobilities of electrons (μe) and holes (μh) are determined via space charge limited current (SCLC) measurements conducted on electron-only and hole-only devices of PM7:Y6 and ZR1:Y6. The PM6:Y6 mobility is taken from our previous study with the same device structure.35 Ideally, a quadratic relationship is formed between the current and voltage in the space charge limited region. However, slightly higher slopes than 2 are observed due to the field dependent mobility in presence of traps. Murgatroyd–Gill equation (eqn (1)) is used in order to account for the field effects and estimate the mobilities in zero-field. We use Gaussian disorder model (GDM) (eqn (2)) to calculate σS from the temperature dependence of μe and μh. (SCLC curves and temperature dependent mobility of ZR1:Y6 and PM7:Y6 are shown in Fig. S6, ESI†)
(1) |
(2) |
The HOMO energetic disorder (σHOMO) are in the order of PM6:Y6 (78 meV) > PM7:Y6 (70 meV) > ZR1:Y6 (59 meV) and all blends exhibit similar LUMO energetic disorder. ZR1:Y6 blend exhibited surprisingly low electron mobility (μe) compared to the two polymer:Y6 blends despite showing similar LUMO disorder (σLUMO). In a similar vein, the hole mobility (μh) of ZR1:Y6, albeit higher than μe, is still lower than the polymer:Y6 blends. This is in contrast with the expectation of enhanced charge transport with suppressed energetic disorder. The SCLC μ and σs are given in Table 2.
μe [cm2 V−1 s−1] | μh [cm2 V−1 s−1] | σLUMO [meV] | σHOMO [meV] | |
---|---|---|---|---|
ZR1:Y6 | 2.7 × 10−5 | 2.8 × 10−4 | 63 | 59 |
PM7:Y6 | 1.3 × 10−3 | 3 × 10−4 | 60 | 70 |
PM6:Y6 | 5.4 × 10−4 | 5.5 × 10−4 | 60 | 78 |
Tenet holds that free charge recombination processes that determine the device VOC take place from the relaxed density of state (DOS).37,38 For Gaussian DOS, the Fermi energies of these charges can be estimated with Gauss-Fermi integral. In the absence of surface recombination, VOC of a solar cell is equal to the quasi-Fermi level splitting (QFSL) of the electrons and holes (qVOC = EF,e − EF,h). Photovoltaic band-gap (Eg) equals the difference between the LUMO energy of the acceptor and the HOMO energy of the donor. There is an energy difference between qVOC and Eg, extent of which depends on σS and charge carrier density (n). At high enough temperatures, Fermi level of relaxed holes (EF,h) and electrons (EF,e) can be determined by the Boltzmann distribution. Hence, temperature dependent VOC is analytically described as:38,39
(3) |
Fig. 2 The temperature dependent photoinduced absorption (PIA) spectra of (a) ZR1:Y6, and (b) PM7:Y6 films on glass with the same active layer thickness with the devices are shown. (c) Calculated charge carrier densities from PIA spectra.41 (d) Experimental VOC as a function of the temperature (scatters) shown with the Gauss fittings of the degenerate and non-degenerate Gauss-Fermi approximation expression (lines) of PM7:Y6 and ZR1:Y6 devices are shown with the linear extrapolation of the high temperature (300–260 K) experimental data to 0 K (dashed lines) yielding 1.08 V for ZR1:Y6 and 1.07 V for PM7:Y6. |
To assign the low μ of the low disorder ASM system, we turn to the definition of charge transport. μ is widely calculated from the Miller–Abrahams hopping rate of charges from an initial state i with the energy Ei to the next state j with energy Ej (eqn (4)) in the established transport models for organic semiconductors.
(4) |
In pursuit of elucidating the underlying factors contributing to the lower mobility, we have conducted morphology analysis using grazing incident wide angle X-ray scattering (GIWAXS). The GIWAXS data in Fig. 3a show that the Y6 neat film has face-on π–π stacking preference, as predicted.25 On the other hand, the ZR1 neat film has significantly higher order and crystallinity compared to Y6 but has an edge-on π–π stacking preference, consistent with literature reports.45–47 Blending ZR1 with Y6 extensively suppresses the crystallinity of both ZR1 and Y6 (Fig. 3c and d). Fig. S8 (ESI†) shows that PM6 neat film, like Y6 film, has face-on π–π stacking order. The OoP GIWAXS profile of PM6:Y6 (shown in Fig. S8, ESI†), unlike in ZR1:Y6, shows a prominent π–π stacking peak. This PM6:Y6 blend π–π stacking peak is shifted between the original π–π stacking peaks of PM6 and Y6 indicating that both PM6 and Y6 π–π stacking are preserved in the blend and contribute to the blend π–π peak. The main difference here is that the face-on π–π stacking peak remains prominent in the PM6:Y6 blend, but almost disappears in ZR1:Y6 blend (Fig. 3d). For both PM7:Y6 and PM6:Y6 blends our data and literature reports indicate dominant face-on orientation of the π–π stacking with respect of the substrate.48–50
Fig. 3 GIWAXS 2D scans for (a) a neat film of Y6, (b) neat film of ZR1 and (c) ZR1:Y6 blend. (d) 1D GIWAXS profiles extracted from a, b and c scans: in the in-plane direction and in the out-of-plane (OoP) direction. (e) A TEM scan of ZR1:Y6 film (f) a TEM scan of PM6:Y6 film. The scale bars in TEM sans are 100 nm (see Fig. S9, ESI† for zoom series). (g) An AFM scan of ZR1:Y6 film showing an RMS roughness of 1.2 nm. (h) An AFM scan of a PM6:Y6 film showing an RMS roughness of 1 nm (h) the scale bars in the AFM sans are 500 nm. |
We further investigate the active layer mesoscale morphology here since it is known to significantly influence charge carrier mobility and the charge generation in organic blends.51 The transmission electron microscopy (TEM) and atomic force microscopy (AFM) scans in Fig. 3(e)–(h) show that both PM6:Y6 and ZR1:Y6 films have relatively smooth surfaces and have features on the order of 10s nm. That is consistent with previous reports.25,45 The ZR1:Y6 film seems slightly rougher than PM6:Y6 with secondary large features on the order of 100s nm, which we interpret as height fluctuations based on the textural resemblance between its AFM and TEM scans (shown in (Fig. 3e and g). Additionally, the surface texture of the ZR1:Y6 film (Fig. 3g) shows no signs of the platelet crystallites that were observed in the neat ZR1 film (Fig. S10, ESI†), in agreement with the GIWAXS results of suppressed crystallinity in ZR1:Y6. We next turn to resonant soft X-ray scattering (RSoXS) to get more insights into domain purity and size of PM6:Y6 and ZR1:Y6 blends. From the RSoXS profiles a scattering feature arises in both blends at the resonant X-ray energy (285.3 eV), Fig. S11 (ESI†), which do not appear at the non-resonant energy (270 eV). This indicates the presence of molecular domains that are uncorrelated with roughness. From these features, characteristic lengths (Lc) which are related to domain size show that ZR1:Y6 has a longer Lc ∼ 60 nm than PM6:Y6 (Lc ∼ 45 nm). Additionally, the total scattering intensity, which is proportional to domain purity, is lower for the ZR1:Y6 film than the PM6:Y6 film, which indicates increased mixing of the ZR1:Y6 domains compared with PM6:Y6. By combining the GIWAXS, RSoXS and microscopic results, we conclude that the ZR1:Y6 has mixed and larger domains than PM6:Y6. This is potentially due to an increased thermodynamic miscibility between the two small molecules and a reason for the lower crystallinity. This resulting morphology is the likely origin of ZR1:Y6 having higher recombination coefficient k2 compared to the polymer:Y6 systems.
The apparent lower charge carrier mobilities of ZR1:Y6 compared to polymer:Y6 blends can be explained with our morphological data showing the lack of parallel stacking of ZR1 and Y6 molecules.52 In this case, smaller disorder values do not help to facilitate more efficient charge transport due to the unfavorable localization length. In organic semiconductor blends, crystallinity and structural order affect charge generation and recombination. Non-geminate recombination (NGR) coefficients of the blends are measured using PIA and bias assisted charge extraction (BACE) methods. When NGR has a second order dependence on n, recombination rate is given as R = k2n2, where k2 is the bimolecular rate coefficient. Fairly second order kinetics are observed for all systems. Data for recombination rate versus carrier density is shown in Fig. S12 (ESI†). Fig. 4a depicts k2 as a function of carrier density for all the three blends. The k2 of PM6:Y6 is considerably smaller than the other two systems, in accordance with the recent literature reports on relationship between ΔES1−CT and the k2.21,30 The k2 of ZR1:Y6 is slightly higher than PM7:Y6 (Table 4). As a common reference point for the charge recombination in organic semiconductors, Langevin theory outlines the case where whenever two opposite charges meet, they recombine. Therefore, the Langevin rate constant is proportional to charge carrier mobility (kL = q(μe + μh)/(ε0εr)). A discrepancy between kL and k2 of organic solar cells is commonly observed and the difference between these coefficients is characterized as the Langevin reduction factor (γ). Origin of γ comes from the competition between the CT state decay rate constant (kf) and charge dissociation rate constant (kd). If kd ≫ kf the overall recombination of the free charges is limited by the resplitting of the CT state (γ = kf/kd + kf). In the resplitting limited regime, k2 is independent of μ.53
k2 [cm3 s−1] | γ | |
---|---|---|
ZR1:Y6 | 4.3 × 10−11 | 0.260 |
PM7:Y6 | 1.3 × 10−11 | 0.015 |
PM6:Y6 | 8 × 10−12 | 0.014 |
We obtain kd as a function of temperature from temperature-dependent k2 combined with the temperature-dependent μ values, assuming same kf of 109 s−1 for all blends, which is a reasonable value for the organic blends.54,55 Temperature dependency of kd are observed to be the same, albeit with a notably lower value for ZR1:Y6 (Fig. 4b). Using the Arrhenius relation, activation energies for charge dissociation were estimated to be similar; 31, 35, and 41 meV for PM6:Y6, PM7:Y6, and ZR1:Y6, respectively (Fig. 4c). This indicates that the lower kd does not originate from the energy barrier between CT and CS for the ZR1:Y6 blend. In order to find the possible causes of the lower kd in ZR1:Y6 we turn to the Onsager–Braun theory, which explains charge dissociation in organic blends in terms of local mobility and Eb of the CT state.
(5) |
As an outlook for future studies, it is worthwhile to consider the role of delocalization in charge generation. Our study has demonstrated that charge separation is more efficient with polymer donors compared to small molecule donors, likely due to greater hole delocalization in polymers. The significance of hole delocalization for charge separation is well-documented in the literature.57 For example, Nenashev et al. reported that delocalization of holes along polymer chains increases the probability of exciton dissociation using a semi-analytic model for polymer blends.58 Similarly, Y6, a small molecule, also shows strong polaron delocalization, which greatly enhances charge generation.59 Crystallinity and packing are crucial factors that promote higher charge delocalization in both polymers and small molecules. Therefore, future studies should investigate the effects of delocalization on polymers versus small molecules to better understand why polymer donors consistently outperform small molecule donors. A comprehensive study involving a broader selection of small molecule and polymer donors would be necessary to draw general conclusions and would be of great interest to the community.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4mh00747f |
This journal is © The Royal Society of Chemistry 2024 |