Bispentafluorophenyl-Containing Additive: Enhancing Efficiency and

Nov 22, 2017 - A new class of additive materials bis(pentafluorophenyl) diesters (BFEs) where the two pentafluorophenyl (C6F5) moieties are attached a...
0 downloads 7 Views 6MB Size
Research Article Cite This: ACS Appl. Mater. Interfaces 2017, 9, 43861−43870

www.acsami.org

Bispentafluorophenyl-Containing Additive: Enhancing Efficiency and Morphological Stability of Polymer Solar Cells via Hand-GrabbingLike Supramolecular Pentafluorophenyl−Fullerene Interactions Kai-En Hung,† Che-En Tsai,† Shao-Ling Chang,† Yu-Ying Lai,‡ U-Ser Jeng,§,∥ Fong-Yi Cao,† Chain-Shu Hsu,† Chun-Jen Su,§ and Yen-Ju Cheng*,† †

National Chiao Tung University, Department of Applied Chemistry, University Road, Hsinchu 30010, Taiwan Institute of Polymer Science and Engineering, National Taiwan University, No.1, Sec.4, Roosevelt Rd, Taipei 10617, Taiwan § National Synchrotron Radiation Research Center, 101 Hsin-Ann Road, Hsinchu Science Park, Hsinchu 30076, Taiwan ∥ National Tsing Hua University, Department of Chemical Engineering, 101, Section 2, Kuang-Fu Road, Hsinchu 30013, Taiwan ‡

S Supporting Information *

ABSTRACT: A new class of additive materials bis(pentafluorophenyl) diesters (BFEs) where the two pentafluorophenyl (C6F5) moieties are attached at the both ends of a linear aliphatic chain with tunable tether lengths (BF5, BF7, and BF13) were designed and synthesized. In the presence of BF7 to restrict the migration of fullerene by hand-grabbing-like supramolecular interactions induced between the C6F5 groups and the surface of fullerene, the P3HT:PC61BM:BF7 device showed stable device characteristics after thermal heating at 150 °C for 25 h. The morphologies of the active layers were systematically investigated by optical microscopy, grazing-incidence small-angle X-ray scattering (GISAXS), and atomic force microscopy. The tether length between the two C6F5 groups plays a pivotal role in controlling the intermolecular attractions. BF13 with a long and flexible tether might form a BF13−fullerene sandwich complex that fails to prevent fullerene’s movement and aggregation, while BF5 with too short tether length decreases the possibility of interactions between the C6F5 groups and the fullerenes. BF7 with the optimal tether length has the best ability to stabilize the morphology. In sharp contrast, the nonfluorinated BP7 analogue without C6F5−C60 physical interactions does not have the capability of morphological stabilization, unambiguously revealing the necessity of the C6F5 group. Most importantly, the function of BF7 can be also applied to the high-performance PffBT4BT-2OD:PC71BM system, which exhibited an original PCE of 8.80%. After thermal heating at 85 °C for 200 h, the efficiency of the PffBT4BT-2OD:PC71BM:BF7 device only decreased slightly to 7.73%, maintaining 88% of its original efficiency. To the best of our knowledge, this is the first time that the thermal-driven morphological evolution of the high-performance PffBT4BT-2OD polymer has been investigated, and its morphological stability in the inverted device can be successfully preserved by the incorporation of BF7. This research also demonstrates that BF7 is not only effective with PC61BM but also to PC71BM. KEYWORDS: additives, fullerene aggregation, morphological stability, supramolecular interactions, polymer solar cells



1).12−14 Mild thermal annealing of P3HT:PC61BM blend induces higher crystallinity of P3HT for efficient charge transport.15−18 Nevertheless, if the thermal heating is applied persistently, the spherical PC61BM is prone to diffuse out of the P3HT domain and form large PC61BM clusters.19−21 Such progressive evolution gives rise to the micrometer-sized phase segregation, dramatically reducing D−A interfacial area and device performance.22,23 Long-term sunlight irradiation inevitably accumulates heat that adversely deteriorates the optimized

INTRODUCTION

Solution-processed polymer solar cells (PSCs) with advantageous features of lightweight and flexible substrates are promising for renewable energy.1−5 Bulk heterojunction (BHJ) solar cells with maximum donor−acceptor interfacial area are the widely adopted configuration for efficient charge separation. Although extensive efforts have been made on nanostructural engineering to tailor the morphology of active layers,6−11 maintaining the optimized morphology under longterm thermal heating is rather challenging. Poly(3-hexylthiophene) (P3HT) as the electron donor and [6,6]-phenyl-C61 butyric acid methyl ester (PC61BM) as the electron acceptor are the most investigated system in BHJ solar cells (Figure © 2017 American Chemical Society

Received: September 5, 2017 Accepted: November 22, 2017 Published: November 22, 2017 43861

DOI: 10.1021/acsami.7b13426 ACS Appl. Mater. Interfaces 2017, 9, 43861−43870

Research Article

ACS Applied Materials & Interfaces

evolution, the P3HT/PC61BM-based devices incorporating PC61BPF as the additive successfully exhibited stable device characteristics against thermal treatment.59 By smartly adopting the concept of C6F5−C60 interactions, we further target development of synthetically more accessible and cost-effective materials. To this end, we have developed a new class of bis(pentafluorophenyl) diesters (BFEs) where a linear aliphatic chain is end-capped by two C6F5 groups via ester linkages. The presence of two C6F5 groups in BFEs is intended to induce hand-grabbing-like interactions to physically hold fullerenes from migration-driven aggregation via the C6F5−C60 interactions. The length of the carbon-chain tether connecting two C6F5 groups in BFEs could sterically affect the miscibility and interactions with the photoactive materials, thus determining the effectiveness of morphological stabilization. To systematically investigate the length-dependent effect, three symmetrical BFEs, bis(pentafluorophenyl)glutarate with a 5-carbon chain (denoted as BF5), bis(pentafluorophenyl)pimelate (BF7) with a 7-carbon chain, and bis(pentafluorophenyl)brassylate (BF13) with a 13-carbon chain, were designed.

Figure 1. Polymer structures used in this research.

morphology.24,25 Several strategies such as reducing the crystallinity of polymers26,27 or fullerenes,28−30 incorporating a compatibilizer,31 and cross-linking between components in active layers were described to stabilize the morphology of active layers.32−35 Restricting PC61BM from severe diffusion is the most effective approach to fix the optimal morphology. The in situ cross-linking reaction between fullerene-based materials has effectively suppressed thermal-driven phase separation.35 The optimized morphology of the active layer can be thus preserved through covalent fixation. However, precisely controlling the temperature to trigger the occurrence of chemical cross-linking is difficult. In comparison with the covalent bonding strategy to fix the morphology, utilization of noncovalent interactions emerges as a simpler and smarter alternative. Introducing additives to achieve the optimal morphology of active layers has been widely implemented.36−47 47 Despite additive strategies gaining substantial success in improving efficiency, the capability of preserving the optimized morphology against thermal heating is imperative but still lacking. Benzene and hexafluorobenzene spontaneously assemble into an alternating face-to-face stacking in order to form complementary quadrupole−quadrupole attractions.48−57 Similar favorable interactions have been demonstrated between a pentafluorophenyl (C6F5) motif and the surface of a fullerene.58 Recently, we developed a PC61BM-based material, PC61BPF, attaching a pentafluorophenyl moiety.59 Through the intermolecular attractions between the fullerene materials and the C6F5 moiety of PC61BPF to suppress the morphological



RESULTS AND DISCUSSION Synthesis of BFEs. The synthesis of BFEs is simple and straightforward. As depicted in Scheme 1, condensation of glutaric acid, pimelic acid, and brassylic acid with 2 equiv pentafluorophenol in the presence of dicyclohexylcarbodiimide (DCC) and 4-dimethylaminopyridine (DMAP) led to the formation of BF5, BF7, and BF13, respectively, with moderate yields. In a similar manner, bis(phenyl)pimelate (BP7) with two nonfluorinated phenyl groups was also synthesized as a reference compound for comparison. Thermal Properties. Differential scanning calorimetry (DSC) measurements showed that the melting point of BFEs decreases as the tether length increases (63, 53, and 40 °C for BF5, BF7, and BF13, respectively, Figure 2a). BF13 has the lowest melting point due to its highest structural flexibility.

Scheme 1. Synthesis of BF5, BF7, and BF13 and BP7

43862

DOI: 10.1021/acsami.7b13426 ACS Appl. Mater. Interfaces 2017, 9, 43861−43870

Research Article

ACS Applied Materials & Interfaces

Figure 2. Differential scanning calorimetry (a) heating curves of BF5, BF7, and BF13 and BP7 at a ramping rate of 10 °C/min and (b) cooling curves of neat PC61BM and PC61BM with 5 wt % BF5, BF7, and BF13 with a cooling rate of 10 °C/min.

Table 1. SCLC Hole and Electron Mobility of P3HT:PC61BM Blends without or with 7 wt % BF5, BF7, and BF13 no additive (cm2 s−1 V−1) hole mobility, μh electron mobility, μe

−5

5.96 × 10 1.08 × 10−4

with BF5 (cm2 s−1 V−1) −5

3.93 × 10 7.88 × 10−5

with BF7 (cm2 s−1 V−1) −5

5.01 × 10 5.09 × 10−5

with BF13 (cm2 s−1 V−1) 4.34 × 10−5 3.08 × 10−5

Figure 3. J−V curves of the ITO/PEDOT:PSS/P3HT:PC61BM/Ca/Al devices with thermal heating at 150 °C for different times (a) without additive, (b) with 7 wt % BF5, (c) with 7 wt % BF7, (d) with 7 wt % BF13, or (e) with 7 wt % BP7.

crystallization point, Tc, at 242 °C with the heat of crystallization (ΔHc) of 12.29 W/g during the cooling of the

Meanwhile, the melting point of the nonfluorinated BP7 is ca. 47 °C. On the other hand, the neat PC61BM showed a clear 43863

DOI: 10.1021/acsami.7b13426 ACS Appl. Mater. Interfaces 2017, 9, 43861−43870

Research Article

ACS Applied Materials & Interfaces DSC measurement. To investigate the influence of the additive on the thermal transition of PC61BM, 5 wt % BFEs were added into PC61BM for the DSC measurements (Figure 2b). After the incorporation of BF5, BF7, and BF13, the Tc of PC61BM gradually decreased to 222, 218, and 203 °C, while the heat of crystallization (ΔHc) also decreased to 8.37, 8.23, and 6.28 W/ g, respectively. The reduced crystallization ability of PC61BM is associated with the presence of C6F5−C60 interactions. Moreover, the longer tether chain of the BFEs results in lower Tc, broader crystallization peak, and smaller ΔHc. Charge Mobility Measurements. To investigate the influence of BFEs on the charge transporting characteristics of the P3HT:PC61BM blend, the space-charge-limited-current (SCLC) measurements using ITO/PEDOT:PSS/active layer/ Au and ITO/Al/active layer/Al device structures were employed to evaluate the hole mobility (μh) and electron mobility (μe), respectively. The results are summarized in Table 1. After introduction of 7 wt % BF5, BF7, and BF13, the μh changed only marginally, while the μe decreased slightly with increasing tether length of the BFEs from 1.08 × 10−4 to 7.88 × 10−5, 5.09 × 10−5, and 3.08 × 10−5 for P3HT:PC61BM, P3HT:PC61BM:BF5, P3HT:PC61BM:BF7, and P3HT:PC61BM:BF13 blends, respectively. These results indicate that the incorporation of the C6F5-containing BFEs only slightly affects the electron mobility of the P3HT:PC61BM system (Table 1). Thermal Stability of the P3HT:PC61BM-Based System. Traditional P3HT:PC61BM-based devices with ITO/PEDOT:PSS/P3HT:PC61BM/Ca/Al with configuration were used to evaluate the ability of the C6F5-containing BFEs for morphological stabilization against thermal treatment. BF5, BF7, and BF13 (7 wt %) additives were doped into the P3HT:PC61BM (1:1 in wt) blend to form ternary-blend active l a y e r s ( P 3 H T : P C 6 1 B M : B F 5 , P3 H T : P C 6 1 B M : B F7 , P3HT:PC61BM:BF13, respectively) which were further heated at 150 °C for 0, 5, 10, 15, 20, and 25 h before depositing top electrodes to complete the devices. Under identical device conditions, reference P3HT:PC61BM devices without introducing any additive were also fabricated for comparison. The J−V curves of all the devices and their corresponding photovoltaic parameters (PCE, Voc, Jsc, and FF) as a function of heating time (5, 10, 15, 20, and 25 h) at 150 °C are shown in Figure 3a−e and Tables S1−S5 (see the Supporting Information, SI). Table 2 only summarizes the device data before and after heating at 150 °C for 25 h. Before the thermal treatment, the devices using P3HT:PC61BM, P3HT:PC61BM:BF5, P3HT:PC61BM:BF7, and P3HT:PC61BM:BF13 blends showed similar efficiencies of 3.81%, 3.73%, 3.70%, and 3.53%, respectively, indicating that the incorporation of 7 wt % C6F5-containing additives does not influence the device performance (Table 2). The efficiency of the P3HT:PC61BM device degraded gradually from 3.81% to 0.57% after heating at 150 °C for 25 h (Figure 3a), which represents an 85% loss in efficiency. The dramatically reduced Jsc but unchanged Voc confirm that the performance deterioration is mainly attributed to the thermal-driven phase separation and PC 61 BM aggregation. However, in the presence of 7 wt % BF5, the P3HT:PC61BM:BF5 device showed a 35% efficiency loss from 3.73% to 2.43% after heating at 150 °C for 25 h (Figure 3b), indicating that BF5 additive already exerts a pronounced effect on alleviating the morphological degradation. Most importantly, the efficiency of the P3HT:PC61BM:BF7 device only decays slightly (8% loss) from 3.70% to 3.43% after 25 h

Table 2. Characteristics of the ITO/PEDOT:PSS/ P3HT:PC61BM with 7 wt % or without BF5, BF7, BF13, and BP7 Additive/Ca/Al Devices Where the Active Layers Were Isothermally Heated at 150 °C for 25 h additive none BF5 BF7 BF13 BP7 a

heating time (h)

Voc (V)

Jsc (mA/cm2)

FF (%)

PCEmax (%)

0 25 0 25 0 25 0 25 0 25

0.60 0.62 0.60 0.60 0.60 0.62 0.60 0.60 0.60 0.62

8.90 2.05 8.71 6.17 8.65 8.16 8.86 1.18 8.45 3.93

71.3 44.6 71.4 63.6 71.2 67.7 66.4 44.1 68.6 57.5

3.81 0.57 3.73 2.43 3.70 3.43 3.53 0.31 3.48 1.40

PCEavga (%) 3.67 0.50 3.56 2.31 3.66 3.40 3.18 0.25 3.08 1.18

± ± ± ± ± ± ± ± ± ±

0.13 0.05 0.16 0.10 0.02 0.01 0.29 0.03 0.35 0.20

Average PCE obtained from 10 devices.

heating (Figure 3c), which suggests that BF7 with longer tether length has better capability of stabilizing the morphology. Interestingly, when the tether length increases to 13 carbons in BF13, the morphological stabilization is totally lost (Figure 3d). An extremely low efficiency of 0.31% was obtained for the P3HT:PC61BM:BF13 device after 25 h heating. These results manifest that the tether length associated with the structural flexibility is an important structural factor for morphological stabilization. The BF13 with the longest aliphatic length might be too flexible to restrict the migration of fullerenes. In contrast, the 5-carbon length in BF5 might be too short for the two C6F5 moieties to easily interact with the fullerenes. The 7-carbon tether in BF7 turns out to have the optimal distance and suitable flexibility to efficiently exert the C6F5−C60 attractions in the bulk. Considering that BF7 is the most effective additive material, we intentionally synthesized a nonfluorinated analogue of BF7, bisphenyl pimelate (BP7), where two nonfluorinated phenyl groups are connected by a 7-carbon tether. In sharp contrast, the device using 7 wt % BP7 showed much lower efficiency of 1.4% after 25 h heating (Figure 3e). These experimental comparisons verify that the existence of pentafluorophenyl groups in BF7 for inducing the C6F5−C60 attractions is imperative to fix the morphology. The PCE decay of P3HT:PC61BM devices without and with 7 wt % BF5, BF7, BF13, and BP7 as a function of heating time at 150 °C is shown in Figure 4. Optical Microscopy. Optical microscopy (OM) was utilized to investigate the morphological change of the P3HT:PC61BM, P3HT:PC61BM:BF5, P3HT:PC61BM:BF7, P3HT:PC61BM:BP7, and P3HT:PC61BM:BF13 thin films, which were prepared identically to those for the device fabrication. The OM images before and after the heating at 150 °C for 25 h are shown in Figure 5. Similar to the prior reports,19 the homogeneity of the P3HT:PC61BM blend was altered subsequent to the heating and numerous long string-like PC61BM aggregates were observed (Figure 5a). When 7 wt % BF5 was introduced, the PC 61 BM aggregates in the P3HT:PC61BM:BF5 film were greatly reduced after heating (Figure 5b), revealing that BF5 has certain capability of preventing PC61BM aggregation. Moreover, the PC61BM aggregates in the P3HT:PC61BM:BF7 blend can be completely suppressed after thermal treatment (Figure 5c), indicating that BF7 stabilizes the morphology most effectively. When BP7, a nonfluorinated version of BF7, was employed, the thermal43864

DOI: 10.1021/acsami.7b13426 ACS Appl. Mater. Interfaces 2017, 9, 43861−43870

Research Article

ACS Applied Materials & Interfaces

than traditional DFT methods.60 Pristine C60 was employed as a simplified model for the fullerene acceptors so as to reduce the computational cost. The optimized geometries of stretching BF5, BF7, and BF13 are depicted in Figure 6a, of which all CH2 moieties adopt staggered conformation. The distances between the two oxygens of O−C6F5 are 7.22, 9.78, and 17.44 Å for BF5, BF7, and BF13, respectively. It was found that the molecular length of BF13 enables its two pentafluorophenyl groups to encapsulate a C60 to form a host−guest sandwich complex. Given that the calculated diameter of C60 is ca. 7 Å and the π stacking distance is within the range of 3−4 Å, the length of a tether must be at least 13 Å long (7 + 3 + 3 Å) to form such a host−guest sandwich complex. Among the three BFEs, only BF13 meets this structural condition. This possibility is further supported by a local minimum of the potential energy surface for the C60−BF13 sandwich complex whose optimized geometry is illustrated in Figure 6b. One C6F5 group is at a distance of 3.50 Å from C60 and the other one is 3.52 Å. The stabilization energy for the formation of the C60− BF13 sandwich complex is estimated by its association energy to be −30.35 kcal mol−1, which is much larger than the previously calculated intermolecular strength (−10.58 kcal mol−1) between C60 and methyl pentafluorobenzonate (Figure 6c) under the same level of theory,59 suggesting that the formation of the C60−BF13 sandwich complex is much more thermodynamically favorable than that of the nonsandwich complex. The OM images in Figure 5d clearly reveal that the P3HT:PC61BM blends with 7 wt % of BF13 subsequent to thermal treatment exhibited large-area spherulite-like crystallization, while others showed scattered the axialite-like crystallization. The existence of the C60−BF13 sandwich complex might be responsible for this disparity. In the sandwich configuration, the two C6F5 groups of BF13 interact with the same fullerene, thus losing the ability to hinder fullerene aggregation. Based on the OM images, it facilitated the crystallization of PC61BM instead, resulting in inadequate morphology for PSC applications. It should be also noted that the electron mobility of P3HT:PC61BM:BF13 drops almost by an order of magnitude compared to the P3HT:PC61BM blend, presumably due to the formation of PC61BM:BF13 sandwich complex having an especially large disruption. Morphological Stabilization of the PffBT4BT2OD:PC71BM-Based System. To date, research on morphological stabilization in the literature has only focused on the traditional P3HT:PC61BM system. Given that a variety of highperformance p-type polymers have been developed for BHJ solar cells, further efforts should be directed toward maintaining their optimized morphology and performance against heating. A donor−acceptor copolymer PffBT4BT-2OD with strong intermolecular aggregation recently emerged as a high profile ptype material (Figure 1).61−64 Inverted devices (ITO/ZnO/ PffBT4BT-2OD:PC71BM/MoO3/Ag) using PffBT4BT-2OD as the donor and PC71BM as the acceptor have successfully accomplished high efficiencies.61−64 To further evaluate the morphological stability of the PffBT4BT-2OD:PC71BM blend, we applied isothermal annealing at 85 °C for 0, 50, 100, and 200 h to the PffBT4BT-2OD:PC71BM blends and monitored the change of efficiency. Unlike the heating conditions for the P3HT:PC61BM system, decreasing the heating temperature to 85 °C but prolonging the heating time to 200 h were adopted in order to avoid degradation at interfaces of the inverted devices during thermal treatment.24 However, the conditions are already harsh enough to mimic the real circumstance of

Figure 4. PCE decay of P3HT:PC61BM devices without and with 7 wt % BF5, BF7, BF13, and BP7 as a function of heating time at 150 °C.

Figure 5. OM images of P3HT:PC61BM films before and after 150 °C heating for 25 h (a) without additive or (b) with 7 wt % BF5, (c) with 7 wt % BF7, (d) with 7 wt % BF13, and (e) with 7 wt % BP7.

driven aggregation of PC61BM was observed again in the P3HT:PC61BM:BP7 OM image after heating, demonstrating that the C6F5 groups unambiguously play a key role in slowing down the morphological evolution (Figure 5e). In contrast, the P3HT:PC61BM:BF13 blend exhibited severe aggregation after heating, confirming that too long tether length fails to suppress the morphological evolution of the P3HT:PC61BM (Figure 5d). The OM images are fairly consistent with the device-performance results. The OM images of P3HT:PC61BM films with different heating times are shown in Figure S1. Computation of the Optimized Geometries of BFEs and Interaction with C60. To understand the tether length effect of BFEs and elucidate why BF13 does not show the ability of morphology stabilization, density-functional-theory (DFT) calculations were performed. Geometry optimizations were all carried out with the Gaussian 09 package at the wB97XD/6-311G(d,p) level of theory. The minimal nature of stationary points was confirmed by frequency analysis. The utilization of the wB97XD functional lies in the fact that it includes empirical correction for dispersion energy, which could describe the weak intermolecular interactions more accurately 43865

DOI: 10.1021/acsami.7b13426 ACS Appl. Mater. Interfaces 2017, 9, 43861−43870

Research Article

ACS Applied Materials & Interfaces

Figure 6. (a) Optimized geometries of stretching BF5, BF7, and BF13, (b) C60−BF13 sandwich complex, and (c) C60 and pentafluorophenyl complex.59

sunlight irradiation. The J−V characteristics and curves of the PffBT4BT-2OD:PC71BM films with different heating times are shown in Table 3 and Figure 7. The device initially exhibited a

polymer and fullerene, leading to the optimized morphology and performance. More importantly, the degradation rate of the device efficiency as a function of heating time is greatly decelerated in the presence of BF7. After 200 h heating (over 8 days), the device still maintained a high efficiency of 7.73%, preserving 88% of its original efficiency, demonstrating that BF7 remains effective in improving the morphological stability of the PffBT4BT-2OD:PC71BM system (Figure 7b). It should be noted that the efficiencies of the PffBT4BT2OD:PC71BM:BF7 device are almost unchanged in the range of 100−200 h heating. Figure 8 shows the PCE decay of the devices as a function of heating time. These results suggest that if the heating is applied persistently over 200 h, the device characteristics of PffBT4BT-2OD:PC71BM:BF7 are prone to be stable, whereas the PffBT4BT-2OD:PC71BM device without BF7 tends to decay continuously. Small-Angle X-ray Scattering Measurements. Grazingincidence small-angle X-ray scattering (GISAXS) was used to monitor the thermal-induced aggregation of the fullerenes in the PffBT4BT-2OD:PC71BM thin films with or without BF7. The as-cast PffBT4BT-2OD:PC71BM thin-film exhibited prominent GISAXS intensity, which should be contributed largely by PC71BM aggregates as elaborately demonstrated previously.65−69 The radius of gyration (Rg) of the PC71BM clusters can be roughly estimated by using the Guinier approximation: I(q) = I(0) exp(−Rg2q2/3) where I(q) is the in-plane scattering intensity extracted from the corresponding 2D GISAXS pattern measured, I(0) is the zero-angle scattering intensity, and q is defined as 4πλ−1 sin θ in terms of the scattering angle 2θ and the wavelength λ. The GISAXS profiles, I(qy) were extracted at the specular beam position within the region qz = 0.03 Å−1 (as marked in Figure S3). After thermal annealing for the PffBT4BT-2OD:PC71BM blends at 85 °C for 50, 100, and 150 h, the low-q GISAXS intensity steadily increased and the Rg of the PC71BM clusters estimated in the q range of 0.01−0.04 Å−1 increased from 2.1 nm (pristine) to 6.6 nm (50 h), 7.7 nm (100 h), and 8.5 nm (150 h), indicating that

Table 3. Characteristics of the ITO/ZnO/PffBT4BT2OD:PC71BM (1:1.2 in wt %) without or with 10 wt % BF7/ MoO3/Ag Devices Where the Active Layers Were Isothermally Heated at 85 °C for Various Times additive none

BF7

a

heating time (h)

Voc (V)

Jsc (mA/cm2)

FF (%)

PCEmax (%)

0 50 100 200 0 50 100 200

0.74 0.72 0.64 0.58 0.76 0.76 0.74 0.76

16.42 13.27 13.86 13.43 17.11 16.58 15.80 15.85

70.9 65.1 58.0 51.9 67.7 66.8 66.7 65.9

8.61 6.21 5.14 4.04 8.80 8.42 7.80 7.73

PCEavga (%) 8.19 5.89 4.79 3.76 8.52 8.22 7.54 7.19

± ± ± ± ± ± ± ±

0.35 0.21 0.21 0.22 0.26 0.14 0.20 0.33

Average PCE obtained from 10 devices.

PCE of 8.61% before thermal treatment. The device efficiencies degraded gradually with increasing heating time. After 200 h heating, the PCE decreased dramatically to 4.04%, which is a 53% reduction of its pristine value (Figure 7a). Similar to the P3HT:PC61BM system, we have demonstrated that the highly crystalline PffBT4BT-2OD:PC71BM blend does not have good morphological stability against thermal heating. To test the effectiveness of our C6F5-containing additive in improving morphological stability of the PffBT4BT-2OD:PC71BM system, 10 wt % BF7 was doped to form ternary PffBT4BT2OD:PC71BM (1:1 wt %):BF7 blends, which were isothermally heated at 85 °C for 0, 50, 100, and 200 h (Table 3). In the presence of BF7, the pristine device without thermal heating achieved a PCE of 8.80%, which is even higher than the device without adding BF7 (8.61%). Therefore, BF7 can simply serve as an additive to modulate the interactions between the 43866

DOI: 10.1021/acsami.7b13426 ACS Appl. Mater. Interfaces 2017, 9, 43861−43870

Research Article

ACS Applied Materials & Interfaces

Figure 7. J−V curves of the ITO/ZnO/PffBT4BT-2OD:PC71BM/MoO3/Ag devices thermally annealed at 85 °C for 50, 100, and 200 h (a) without additive and (b) with 10 wt % BF7.

indicates a structural enhancement in an even larger length scale, presumably from agglomeration of PC71BM aggregates.68 Nevertheless, when 10 wt % BF7 was introduced into the PffBT4BT-2OD:PC71BM blend, the intensity growth of the corresponding GISAXS profiles can be suppressed effectively in similar thermal annealing. Correspondingly, the PC71BM aggregation size similarly extracted in the q range of 0.01− 0.04 Å−1 remains stable in the range of 3.3 to 3.6 nm (Figure 9b, Table S6) without much growth. Furthermore, relatively moderate intensity increase in the very low-q regime near 0.003 Å reveals less confined agglomeration of PCBM aggregates. These results provide direct evidence that BF7 is very efficient in suppressing thermal-induced PC71BM aggregation in the crystalline PffBT4BT-2OD polymer during thermal treatment. Better suppression of large-scale agglomeration from PCBM aggregates, however, would require longer BFE derivatives. (Rg of PC71BM cluster aggregates in PffBT4BT-2OD with different heating times are shown in Figure S2, Figure S3, and Table S6 Supporting Information). Atomic Force Microscopy. Atomic force microscopy (AFM) was used to observe the evolution of the surface morphology of PffBT4BT-2OD:PC71BM films before and after heating. After 200 h heating at 85 °C, the Rq value of PffBT4BT-2OD:PC71BM thin film decreases from 4.19 nm

Figure 8. PCE decay of PffBT4BT-2OD:PC71BM and PffBT4BT2OD:PC71BM:10 wt % BF7 devices as a function of heating time at 85 °C.

the PC61BM is gradually aggregated into larger domains upon thermal treatment (Figure 9a, Table S6). The significantly increased intensity in the very low q region near 0.003 Å−1

Figure 9. GISAXS profiles of the as-cast and annealed PffBT4BT-2OD/PC71BM films at 85 °C for 0, 50, 100, and 150 h (a) without BF7 or (b) with 10 wt % BF7. The arrows in panel a mark the drastic intensity changes in the low- and high-q regimes owing to growth of individual PC71BM aggregates and their agglomeration in different length scales upon annealing. 43867

DOI: 10.1021/acsami.7b13426 ACS Appl. Mater. Interfaces 2017, 9, 43861−43870

Research Article

ACS Applied Materials & Interfaces

Figure 10. AFM images of PffBT4BT-2OD:PC71BM films: (a) height and (b) phase images before heating and (c) height and (d) phase images after 200 h heating at 85 °C. AFM images of PffBT4BT-2OD:PC71BM:BF7 films: (e) height and (f) phase images before heating and (g) height and (h) phase images after 200 h heating at 85 °C.

heating at 85 °C for 200 h, the efficiency only decreased slightly to 7.73%, maintaining 88% of its original efficiency. To the best of our knowledge, this is the first time that the thermal-driven morphological evolution of a high-performance PffBT4BT2OD polymer other than P3HT has been investigated, and its morphological stability in the inverted device can be successfully preserved by the introduction of BF7. We demonstrate that the function of BF7 is widely applicable to both n-type PC61BM and PC71BM with highly crystalline ptype polymers to achieve highly efficient and stable PSCs.

(pristine) to 3.65 nm (annealed for 200 h). However, the surface morphology of the PffBT4BT-2OD:PC71BM:BF7 thin film does not show notable change before and after 200 h heating at 85 °C; the Rq value of PffBT4BT-2OD:PC71BM:BF7 thin film decreases from 3.82 nm (pristine with BF7) to 3.55 nm (annealed for 200 h). It again suggests that the morphology is stabilized by BF7 (Figure 10).



CONCLUSIONS We have developed a new class of synthetically easy and costeffective materials, bis(pentafluorophenyl) diesters (BF5, BF7, and BF13), where the two pentafluorophenyl moieties are attached at both ends of a linear aliphatic chain with tunable tether lengths (5-carbon for BF5, 7-carbon for BF7, and 13carbon for BF13). These materials function as additives to be incorporated into the traditional P3HT:PC61BM blend. Through the hand-grabbing-like C6F5−C60 attractions to restrict the migration of PC61BM, BF7 can effectively suppress fullerene aggregation to slow down morphological evolution. The P3HT:PC61BM:BF7 device showed very stable device characteristics after thermal heating at 150 °C for 25 h. The tether length plays a key role in controlling the intermolecular interactions. In sharp contrast, the device incorporating BF13 did not possess any morphological stability upon thermal heating due to the fact that BF13 with long and flexible tether might encapsulate a PCBM to form a BF13−fullerene complex that fails to prevent fullerene’s movement and aggregation. On the other hand, the tether length of BF5 is relatively short, which might decrease the possibility of interactions between the C6F5 groups and the fullerenes. Therefore, the efficiency of the P3HT:PC61BM:BF5 device decreased by 35% after thermal heating at 150 °C for 25 h. BF7 with the optimal tether length has the best ability to stabilize the morphology. In sharp contrast, the nonfluorinated BP7 analogue without C6F5−C60 physical interactions does not have the ability to stabilize the morphology, unambiguously revealing the importance of the perfluorobenzyl group. Most importantly, BF7 is also effective in the high-performance PffBT4BT-2OD:PC71BM system, which exhibited an original PCE of 8.80%. After thermal



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.7b13426. Device characteristics, OM images, X-ray scattering measurements, synthesis of BFEs, computational data, and NMR spectra (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Yu-Ying Lai: 0000-0002-1921-6923 U-Ser Jeng: 0000-0003-0780-4557 Yen-Ju Cheng: 0000-0002-2247-5061 Author Contributions

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We thank the Ministry of Science and Technology and the Ministry of Education, Taiwan for financial support. We thank the National Center of High-Performance Computing (NCHC) in Taiwan for computer time and facilities. 43868

DOI: 10.1021/acsami.7b13426 ACS Appl. Mater. Interfaces 2017, 9, 43861−43870

Research Article

ACS Applied Materials & Interfaces



(19) Klimov, E.; Li, W.; Yang, X.; Hoffmann, G. G.; Loos, J. Scanning Near-Field and Confocal Raman Microscopic Investigation of P3HT− PCBM Systems for Solar Cell Applications. Macromolecules 2006, 39, 4493−4496. (20) Zhong, H.; Yang, X.; deWith, B.; Loos, J. Quantitative Insight into Morphology Evolution of Thin PPV/PCBM Composite Films upon Thermal Treatment. Macromolecules 2006, 39, 218−223. (21) Müller, C.; Ferenczi, T. A. M.; Campoy-Quiles, M.; Frost, J. M.; Bradley, D. D. C.; Smith, P.; Stingelin-Stutzmann, N.; Nelson, J. Binary Organic Photovoltaic Blends: A Simple Rationale for Optimum Compositions. Adv. Mater. 2008, 20, 3510−3515. (22) Yang, X.; van Duren, J. K. J.; Janssen, R. A. J.; Michels, M. A. J.; Loos, J. Morphology and Thermal Stability of the Active Layer in Poly(p-phenylenevinylene)/Methanofullerene Plastic Photovoltaic Devices. Macromolecules 2004, 37, 2151−2158. (23) Swinnen, A.; Haeldermans, I.; van de Ven, M.; D’Haen, J.; Vanhoyland, G.; Aresu, S.; D’Olieslaeger, M.; Manca, J. Tuning the Dimensions of C60-Based Needlelike Crystals in Blended Thin Films. Adv. Funct. Mater. 2006, 16, 760−765. (24) Jørgensen, M.; Norrman, K.; Krebs, F. C. Stability/Degradation of Polymer Solar Cells. Sol. Energy Mater. Sol. Cells 2008, 92, 686−714. (25) Bartelt, J. A.; Beiley, Z. M.; Hoke, E. T.; Mateker, W. R.; Douglas, J. D.; Collins, B. A.; Tumbleston, J. R.; Graham, K. R.; Amassian, A.; Ade, H.; Fréchet, J. M. J.; Toney, M. F.; McGehee, M. D. The Importance of Fullerene Percolation in the Mixed Regions of Polymer−Fullerene Bulk Heterojunction Solar Cells. Adv. Energy Mater. 2013, 3, 364−374. (26) Sivula, K.; Luscombe, C. K.; Thompson, B. C.; Fréchet, J. M. J. Enhancing the Thermal Stability of Polythiophene:Fullerene Solar Cells by Decreasing Effective Polymer Regioregularity. J. Am. Chem. Soc. 2006, 128, 13988−13989. (27) Bertho, S.; Campo, B.; Piersimoni, F.; Spoltore, D.; D’Haen, J.; Lutsen, L.; Maes, W.; Vanderzande, D.; Manca, J. Improved Thermal Stability of Bulk Heterojunctions Based on Side-Chain Functionalized Poly(3-alkylthiophene) Copolymers and PCBM. Sol. Energy Mater. Sol. Cells 2013, 110, 69−76. (28) Zhang, Y.; Yip, H.-L.; Acton, O.; Hau, S. K.; Huang, F.; Jen, A. K. Y. A Simple and Effective Way of Achieving Highly Efficient and Thermally Stable Bulk-Heterojunction Polymer Solar Cells Using Amorphous Fullerene Derivatives as Electron Acceptor. Chem. Mater. 2009, 21, 2598−2600. (29) Cheng, Y.-J.; Liao, M.-H.; Chang, C.-Y.; Kao, W.-S.; Wu, C.-E.; Hsu, C.-S. Di(4-methylphenyl)methano-C60 Bis-Adduct for Efficient and Stable Organic Photovoltaics with Enhanced Open-Circuit Voltage. Chem. Mater. 2011, 23, 4056−4062. (30) Li, C.-Z.; Chien, S.-C.; Yip, H.-L.; Chueh, C.-C.; Chen, F.-C.; Matsuo, Y.; Nakamura, E.; Jen, A. K. Y. Facile Synthesis of a 56πElectron 1,2-Dihydromethano-[60]PCBM and its Application for Thermally Stable Polymer Solar Cells. Chem. Commun. 2011, 47, 10082−10084. (31) Sivula, K.; Ball, Z. T.; Watanabe, N.; Fréchet, J. M. J. Amphiphilic Diblock Copolymer Compatibilizers and Their Effect on the Morphology and Performance of Polythiophene:Fullerene Solar Cells. Adv. Mater. 2006, 18, 206−210. (32) Kim, B. J.; Miyamoto, Y.; Ma, B.; Fréchet, J. M. J. Photocrosslinkable Polythiophenes for Efficient, Thermally Stable, Organic Photovoltaics. Adv. Funct. Mater. 2009, 19, 2273−2281. (33) Gholamkhass, B.; Holdcroft, S. Toward Stabilization of Domains in Polymer Bulk Heterojunction Films. Chem. Mater. 2010, 22, 5371−5376. (34) Cheng, Y.-J.; Hsieh, C.-H.; Li, P.-J.; Hsu, C.-S. Morphological Stabilization by In Situ Polymerization of Fullerene Derivatives Leading to Efficient, Thermally Stable Organic Photovoltaics. Adv. Funct. Mater. 2011, 21, 1723−1732. (35) Kim, H. J.; Han, A. R.; Cho, C.-H.; Kang, H.; Cho, H.-H.; Lee, M. Y.; Fréchet, J. M. J.; Oh, J. H.; Kim, B. J. Solvent-Resistant Organic Transistors and Thermally Stable Organic Photovoltaics Based on Cross-linkable Conjugated Polymers. Chem. Mater. 2012, 24, 215− 221.

REFERENCES

(1) Cheng, Y.-J.; Yang, S.-H.; Hsu, C.-S. Synthesis of Conjugated Polymers for Organic Solar Cell Applications. Chem. Rev. 2009, 109, 5868−5923. (2) Brabec, C. J.; Sariciftci, N. S.; Hummelen, J. C. Plastic Solar Cells. Adv. Funct. Mater. 2001, 11, 15−26. (3) Dennler, G.; Scharber, M. C.; Brabec, C. J. Polymer-Fullerene Bulk-Heterojunction Solar Cells. Adv. Mater. 2009, 21, 1323−1338. (4) Thompson, B. C.; Fréchet, J. M. J. Polymer-Fullerene Composite Solar Cells. Angew. Chem., Int. Ed. 2008, 47, 58−77. (5) Yu, G.; Gao, J.; Hummelen, J. C.; Wudl, F.; Heeger, A. J. Polymer Photovoltaic Cells: Enhanced Efficiencies via a Network of Internal Donor-Acceptor Heterojunctions. Science 1995, 270, 1789−1791. (6) Ma, W.; Yang, C.; Gong, X.; Lee, K.; Heeger, A. J. Thermally Stable, Efficient Polymer Solar Cells with NanoscaleControl of the Interpenetrating Network Morphology. Adv. Funct. Mater. 2005, 15, 1617−1621. (7) Kozub, D. R.; Vakhshouri, K.; Orme, L. M.; Wang, C.; Hexemer, A.; Gomez, E. D. Polymer Crystallization of Partially Miscible Polythiophene/Fullerene Mixtures Controls Morphology. Macromolecules 2011, 44, 5722−5726. (8) Campoy-Quiles, M.; Ferenczi, T.; Agostinelli, T.; Etchegoin, P. G.; Kim, Y.; Anthopoulos, T. D.; Stavrinou, P. N.; Bradley, D. D. C.; Nelson, J. Morphology Evolution via Self-Organization and Lateral and Vertical Diffusion in Polymer:Fullerene Solar Cell Blends. Nat. Mater. 2008, 7, 158−164. (9) Kang, H.; Lee, W.; Oh, J.; Kim, T.; Lee, C.; Kim, B. J. From Fullerene-Polymer to All-Polymer Solar Cells: The Importance of Molecular Packing, Orientation, and Morphology Control. Acc. Chem. Res. 2016, 49, 2424−2434. (10) Lee, C.; Kang, H.; Lee, W.; Kim, T.; Kim, K.-H.; Woo, H. Y.; Wang, C.; Kim, B. J. High-Performance All-Polymer Solar Cells via Side-Chain Engineering of the Polymer Acceptor: the Importance of the Polymer Packing Structure and the Nanoscale Blend Morphology. Adv. Mater. 2015, 27, 2466−2471. (11) Cai, W.; Liu, P.; Jin, Y.; Xue, Q.; Liu, F.; Russell, T. P.; Huang, F.; Yip, J.-L.; Cao, Y. Morphology Evolution in High-Performance Polymer Solar Cells Processed from Nonhalogenated Solvent. Adv. Sci. 2015, 2, 1500095−1500101. (12) Kim, Y.; Cook, S.; Tuladhar, S. M.; Choulis, S. A.; Nelson, J.; Durrant, J. R.; Bradley, D. D. C.; Giles, M.; McCulloch, I.; Ha, C.-S.; Ree, M. A Strong Regioregularity Effect in Self-Organizing Conjugated Polymer Films and High-Efficiency Polythiophene:Fullerene Solar Cells. Nat. Mater. 2006, 5, 197−203. (13) Burkhard, G. F.; Hoke, E. T.; Scully, S. R.; McGehee, M. D. Incomplete Exciton Harvesting from Fullerenes in Bulk Heterojunction Solar Cells. Nano Lett. 2009, 9, 4037−4041. (14) Guo, S.; Brandt, C.; Andreev, T.; Metwalli, E.; Wang, W.; Perlich, J.; Müller-Buschbaum, P. First Step into Space: Performance and Morphological Evolution of P3HT: PCBM Bulk Heterojunction Solar Cells under AM0 Illumination. ACS Appl. Mater. Interfaces 2014, 6, 17902−17910. (15) Nguyen, L. H.; Hoppe, H.; Erb, T.; Günes, S.; Gobsch, G.; Sariciftci, N. S. Effects of Annealing on the Nanomorphology and Performance of Poly(alkylthiophene):Fullerene Bulk-Heterojunction Solar Cells. Adv. Funct. Mater. 2007, 17, 1071−1078. (16) Li, G.; Yao, Y.; Yang, H.; Shrotriya, V.; Yang, G.; Yang, Y. Solvent Annealing” Effect in Polymer Solar Cells Basedon Poly(3hexylthiophene) and Methanofullerenes. Adv. Funct. Mater. 2007, 17, 1636−1644. (17) Treat, N. D.; Brady, M. A.; Smith, G.; Toney, M. F.; Kramer, E. J.; Hawker, C. J.; Chabinyc, M. L. Interdiffusion of PCBM and P3HT Reveals Miscibility in a Photovoltaically Active Blend. Adv. Energy Mater. 2011, 1, 82−89. (18) Kim, H. J.; Kim, J.-H.; Ryu, J.-H.; Kim, Y.; Kang, H.; Lee, W. B.; Kim, T. − S.; Kim, B. J. Architectural Engineering of Rod−Coil Compatibilizers for Producing Mechanically and Thermally Stable Polymer Solar Cells. ACS Nano 2014, 8, 10461−10470. 43869

DOI: 10.1021/acsami.7b13426 ACS Appl. Mater. Interfaces 2017, 9, 43861−43870

Research Article

ACS Applied Materials & Interfaces (36) Graham, K. R.; Wieruszewski, P. M.; Stalder, R. M.; Hartel, J.; Mei, J.; So, F.; Reynolds, J. R. Improved Performance of Molecular Bulk-Heterojunction Photovoltaic Cells through Predictable Selection of Solvent Additives. Adv. Funct. Mater. 2012, 22, 4801−4813. (37) Lee, J. K.; Ma, W. L.; Brabec, C. J.; Yuen, J.; Moon, J. S.; Kim, J. Y.; Lee, K.; Bazan, G. C.; Heeger, A. J. Processing Additives for Improved Efficiency from Bulk Heterojunction Solar Cells. J. Am. Chem. Soc. 2008, 130, 3619−3623. (38) Lou, S. J.; Szarko, J. M.; Xu, T.; Yu, L.; Marks, T. J.; Chen, L. X. Effects of Additives on the Morphology of Solution Phase Aggregates Formed by Active Layer Components of High-Efficiency Organic Solar Cells. J. Am. Chem. Soc. 2011, 133, 20661−20663. (39) Hoven, C. V.; Dang, X.-D.; Coffin, R. C.; Peet, J.; Nguyen, T.Q.; Bazan, G. C. Improved Performance of Polymer Bulk Heterojunction Solar Cells Through the Reduction of Phase Separation via Solvent Additives. Adv. Mater. 2010, 22, E63−E66. (40) Rogers, J. T.; Schmidt, K.; Toney, M. F.; Bazan, G. C.; Kramer, E. J. Time-Resolved Structural Evolution of Additive-Processed Bulk Heterojunction Solar Cells. J. Am. Chem. Soc. 2012, 134, 2884−2887. (41) Yiu, A. T.; Beaujuge, P. M.; Lee, O. P.; Woo, C. H.; Toney, M. F.; Fréchet, J. M. J. Side-Chain Tunability of Furan-Containing LowBand-Gap Polymers Provides Control of Structural Order in Efficient Solar Cells. J. Am. Chem. Soc. 2012, 134, 2180−2185. (42) Xin, H.; Guo, X.; Ren, G.; Watson, M. D.; Jenekhe, S. A. Efficient Phthalimide Copolymer-Based Bulk Heterojunction Solar Cells: How the Processing Additive Influences Nanoscale Morphology and Photovoltaic Properties. Adv. Energy Mater. 2012, 2, 575−582. (43) Chang, C.-Y.; Cheng, Y.-J.; Hung, S.-H.; Wu, J.-S.; Kao, W.-S.; Lee, C.-H.; Hsu, C.-S. Combination of Molecular, Morphological, and Interfacial Engineering to Achieve Highly Efficient and Stable Plastic Solar Cells. Adv. Mater. 2012, 24, 549−553. (44) Lobez, J. M.; Andrew, T. L.; Bulovic, V.; Swager, T. M. Improving the Performance of P3HT−Fullerene Solar Cells with SideChain-Functionalized Poly(thiophene) Additives: A New Paradigm for Polymer Design. ACS Nano 2012, 6, 3044−3056. (45) Fan, B.; Sun, C.; Jiang, X.-F.; Zhang, G.; Chen, Z.; Ying, L.; Huang, F.; Cao, Y. Improved Morphology and Efficiency of Polymer Solar Cells by Processing Donor−Acceptor Copolymer Additives. Adv. Funct. Mater. 2016, 26, 6479−6488. (46) Güldal, N. S.; Berlinghof, M.; Kassar, T.; Du, X.; Jiao, X.; Meyer, M.; Ameri, T.; Osvet, A.; Li, N.; Destri, G. L.; Fink, R. H.; Ade, H.; Unruh, T.; Brabec, C. J. Controlling Additive Behavior to Reveal an Alternative Morphology Formation Mechanism in Polymer: Fullerene Bulk-Heterojunctions. J. Mater. Chem. A 2016, 4, 16136−16147. (47) Fan, B.; Ying, L.; Huang, F.; Cao, Y. Processing a Pyridyl-Based Polymeric Additive for Improved Photovoltaic Performance of a WideBandgap π-Conjugated Polymer. J. Mater. Chem. C 2016, 4, 8052− 8060. (48) Patrick, C. R.; Prosser, G. S. A Molecular Complex of Benzene and Hexafluorobenzene. Nature 1960, 187, 1021−1021. (49) Zhou, X.-H.; Luo, J.; Huang, S.; Kim, T.-D.; Shi, Z.; Cheng, Y.J.; Jang, S.-H.; Knorr, D. B.; Overney, R. M.; Jen, A. K. Y. Supramolecular Self-Assembled Dendritic Nonlinear Optical Chromophores: Fine-Tuning of Arene−Perfluoroarene Interactions for Ultralarge Electro-Optic Activity and Enhanced Thermal Stability. Adv. Mater. 2009, 21, 1976−1981. (50) Coates, G. W.; Dunn, A. R.; Henling, L. M.; Ziller, J. W.; Lobkovsky, E. B.; Grubbs, R. H. Phenyl−Perfluorophenyl Stacking Interactions: Topochemical [2 + 2] Photodimerization and Photopolymerization of Olefinic Compounds. J. Am. Chem. Soc. 1998, 120, 3641−3649. (51) Gray, T.; Kim, T.-D.; Knorr, D. B., Jr.; Luo, J.; Jen, A. K. Y.; Overney, R. M. Mesoscale Dynamics and Cooperativity of Networking Dendronized Nonlinear Optical Molecular Glasses. Nano Lett. 2008, 8, 754−759. (52) Meyer, E. A.; Castellano, R. K.; Diederich, F. Interactions with Aromatic Rings in Chemical and Biological Recognition. Angew. Chem., Int. Ed. 2003, 42, 1210−1250.

(53) Renak, M. L.; Bartholomew, G. P.; Wang, S.; Ricatto, P. J.; Lachicotte, R. J.; Bazan, G. C. Fluorinated Distyrylbenzene Chromophores: Effect of Fluorine Regiochemistry on Molecular Properties and Solid-State Organization. J. Am. Chem. Soc. 1999, 121, 7787−7799. (54) Salonen, L. M.; Ellermann, M.; Diederich, F. Aromatic Rings in Chemical and Biological Recognition: Energetics and Structures. Angew. Chem., Int. Ed. 2011, 50, 4808−4842. (55) Kilbinger, A. F. M.; Grubbs, R. H. Arene−Perfluoroarene Interactions as Physical Cross-Links for Hydrogel Formation. Angew. Chem., Int. Ed. 2002, 41, 1563−1566. (56) Guo, S.; Ning, J.; Körstgens, V.; Yao, Y.; Herzig, E. M.; Roth, S. V.; Müller-Buschbaum, P. The Effect of Fluorination in Manipulating the Nanomorphology in PTB7:PC71BM Bulk Heterojunction. Systems. Adv. Energy Mater. 2015, 5, 1401315. (57) Kim, T.; Choi, J.; Kim, H. J.; Lee, W.; Kim, B. J. Comparative Study of Thermal Stability, Morphology, and Performance of AllPolymer, Fullerene−Polymer, and Ternary Blend Solar Cells Based on the Same Polymer Donor. Macromolecules 2017, 50, 6861−6871. (58) Li, C.-Z.; Matsuo, Y.; Niinomi, T.; Sato, Y.; Nakamura, E. Faceto-Face C6F5−[60]Fullerene Interaction for Ordering Fullerene Molecules and Application to Thin-Film Organic Photovoltaics. Chem. Commun. 2010, 46, 8582−8584. (59) Liao, M.-H.; Tsai, C.-E.; Lai, Y.-Y.; Cao, F.-Y.; Wu, J.-S.; Wang, C.-L.; Hsu, C.-S.; Liau, I.; Cheng, Y.-J. Morphological Stabilization by Supramolecular Perfluorophenyl-C60 Interactions Leading to Efficient and Thermally Stable Organic Photovoltaics. Adv. Funct. Mater. 2014, 24, 1418−1429. (60) Chai, J.-D.; Head-Gordon, M. Long-Range Corrected Hybrid Density Functionals with Damped Atom-Atom Dispersion Corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615−6620. (61) Liu, Y.; Zhao, J.; Li, Z.; Mu, C.; Ma, W.; Hu, H.; Jiang, K.; Lin, H.; Ade, H.; Yan, H. Aggregation and Morphology Control Enables Multiple Cases of High-Efficiency Polymer Solar Cells. Nat. Commun. 2014, 5, 5293−5300. (62) Zhao, J.; Li, Y.; Yang, G.; Jiang, K.; Lin, H.; Ade, H.; Ma, W.; Yan, H. Efficient Organic Solar Cells Processed from Hydrocarbon Solvents. Nat. Energy 2016, 1, 15027−15033. (63) Ma, W.; Yang, G.; Jiang, K.; Carpenter, J. H.; Wu, Y.; Meng, X.; McAfee, T.; Zhao, J.; Zhu, C.; Wang, C.; Ade, H.; Yan, H. Organic Solar Cells: Influence of Processing Parameters and Molecular Weight on the Morphology and Properties of High-Performance PffBT4T2OD: PC71BM Organic Solar Cells. Adv. Energy Mater. 2015, 5, 1501400. (64) Zhao, F.; Li, Y.; Wang, Z.; Yang, Y.; Wang, Z.; He, G.; Zhang, J.; Jiang, L.; Wang, T.; Wei, Z.; Ma, W.; Li, B.; Xia, A.; Li, Y.; Wang, C. Combining Energy Transfer and Optimized Morphology for Highly Efficient Ternary Polymer Solar Cells. Adv. Energy Mater. 2017, 7, 1602552−1602560. (65) Jørgensen, M.; Norrman, K.; Krebs, F. C. Stability/Degradation of Polymer Solar Cells. Sol. Energy Mater. Sol. Cells 2008, 92, 686−714. (66) Chirvase, D.; Parisi, J.; Hummelen, J. C.; Dyakonov, V. Influence of Nanomorphology on the Photovoltaic Action of Polymer−Fullerene Composites. Nanotechnology 2004, 15, 1317− 1323. (67) Liu, C.-M.; Su, Y.-W.; Jiang, J.-M.; Chen, H.-C.; Lin, S.-W.; Su, C.-J.; Jeng, U.; Wei, K.-H. Complementary Solvent Additives Tune the Orientation of Polymer Lamellae, Reduce the Sizes of Aggregated Fullerene Domains, and Enhance the Performance of Bulk Heterojunction Solar Cells. J. Mater. Chem. A 2014, 2, 20760−20769. (68) Jhuo, H.-J.; Liao, S.-H.; Li, Y.-L.; Yeh, P.-N.; Chen, S.-A.; Wu, W.-R.; Su, C.-J.; Lee, J.-J.; Yamada, N. L.; Jeng, U. The Novel Additive 1-Naphthalenethiol Opens a New Processing Route to EfficiencyEnhanced Polymer Solar Cells. Adv. Funct. Mater. 2016, 26, 3094− 3104. (69) Chiu, M.-Y.; Jeng, U.; Su, C. H.; Liang, K. S.; Wei, K.-H. Simultaneous Use of Small- and Wide-Angle X-ray Techniques to Analyze Nanometerscale Phase Separation in Polymer Heterojunction Solar Cells. Adv. Adv. Mater. 2008, 20, 2573−2578. 43870

DOI: 10.1021/acsami.7b13426 ACS Appl. Mater. Interfaces 2017, 9, 43861−43870