M. L. Keshtov*a,
S. A. Kuklina,
N. A. Radychevb,
I. E. Ostapova,
A. Y. Nikolaeva,
I. O. Konstantinovc,
M. M. Krayushkind,
E. N. Koukarasde,
Abhishek Sharmaf and
G. D. Sharma*f
aInstitute of Organoelement Compounds of the Russian Academy of Sciences, Vavilova St., 28, 119991 Moscow, Russian Federation. E-mail: keshtov@ineos.ac.ru
bCarl von Ossietzky University of Oldenburg, Oldenburg, 26129, Germany
cN.D. Zelinsky Institute of Organic Chemistry of the Russian Academy of Sciences, Leninsky Prospect 47, 119991 Moscow, Russian Federation
dNanotechnology and Advanced Materials Laboratory, Department of Chemical Engineering, University of Patras, Patras, 26500 GR, Greece
eMolecular Engineering Laboratory, Department of Physics, University of Patras, Patras, 26500 GR, Greece
fDepartment of Physics, The LNM Institute of Information Technology, Jamdoli, Jaipur, (Raj.) 302031, India. E-mail: gdsharma273@gmail.com
First published on 18th July 2016
Two low bandgap D-A1–D-A2 conjugated copolymers, namely denoted as P1 (non-fluorine substituted thiadiazoloquinoxaline A2) and P2 (fluorine substituted thiadiazoloquinoxaline A2) with the same D (thiophene) and A1 (benzothiadiazole) groups were synthesized in order to investigate the effect of fluorine atoms on the photovoltaic performance of polymer solar cells. The electrochemical properties demonstrate that the highest occupied molecular orbital (HOMO) energy level lowered from −5.08 eV (for P1) to −5.16 eV (for P2), whereas the lowest unoccupied molecular orbital (LUMO) energy levels remain nearly the same. These copolymers showed strong absorption in the wavelength range 300–1100 nm and have a bandgap of around 1.08 and 1.11 eV for P1 and P2, respectively. After the optimization of the weight ratio and concentration of solvent additives 1-chloronaphathalene (CN), the highest power conversion efficiencies of bulk heterojunction polymer solar cells achieved were up to 5.30% and 7.21% for P1 and P2 as donor and PC71BM as acceptor. The enhanced Voc and Jsc for the P2 based device can be mainly ascribed to the lower HOMO energy levels and higher hole mobility and better morphology of the fluorinated copolymer P2:
PC71BM blend.
To date, breakthrough in PCE rapid increase was mainly achieved via D–A approach with strict electron-donor and electron-acceptor units alternation in the polymer chain which is the most attractive and successful strategy for controlling the energy levels and optical bandgap of copolymers via intramolecular charge transfer. While D–A polymers bandgap and energy levels can be controlled by combination of different electron-donor and electron-acceptor structures, nevertheless the conventional D–A copolymers still doesn't provide sufficient opportunities for obtaining suitable energy levels and broad solar radiation absorption, since D–A possess rather narrow absorption bands with 200 nm half-maximum width. Improving PSC efficiency via common D–A copolymers structure modification of D and A units combination now seems to have reached a plateau. In order to overcome these limitations of D–A polymers we follow a straightforward extension approach to develop new regular alternating “D-A1–A2-D” polymer structure. A lot of important factors for highly efficient electron donors such as light absorption, HOMO/LUMO levels, hole mobility and D–A copolymers solubility of can carefully varied through second component introduction via a D-A1–D-A2 approach. Unlike conventional D–A polymers containing only one electron-acceptor block in repeating unit, introducing two electron-acceptor units with different absorption capacity into alternating D-A1–D-A2 copolymer structure will provide broader absorption spectrum than its components. Using additional electron-acceptor block as the second A2 unit in D–A block copolymer can broaden the absorption spectrum up to near-infrared region due to appearance of new intramolecular charge transfer ICT bands, and also to cause a synergistic effect in optical, electrical and structural properties of copolymers. Thus, by applying this approach, we can utilize advantages reported in earlier experiments and D and A units to develop their new combinations to meet conjugated polymers desired properties. One of main reasons for using multiple semiconductor components in polymers synthesis is that it can lead to absorption broadening to favor photons absorption in photovoltaic cells and molecular packing configuration by introducing third component that facilitates stacking.
Recently Wang et al. have shown the advantages of the D-A1–D-A2 strategy to improve the PCE of its D-A1 and D-A2 components.12 Janssen et al. developed isomeric random and regular alternating conjugated terpolymers containing two electron-acceptors diketopyrrolopyrrole (DPP), thienopyrrolidone (TPD) and one electron donor, to study monomer units sequential distribution effect on semiconducting properties.13 It has been shown that the optical properties and electrochemical characteristics of photovoltaic regular and random terpolymers significantly differ from each other. Randomly spaced donor and acceptor structure along the polymer chain has a significant negative impact on the photovoltaic performance resulting in PSC efficiency for random terpolymers to reach just 1.0%, while reaching more than 5.3% for regular terpolymers.
Li recently developed semi-crystalline conjugated “D-A1–D-A2” polymer structure on basis of two electron-acceptor pentacycliclactam and DPP units for use in field effect transistors and polymer solar cells.14 Polymer absorbs light over a wide range of solar spectrum from 350 to near-infrared up to 900 nm and exhibits high hole mobility of 0.81 cm2 V−1 s−1 and PSC efficiency of 4.7%.
Fluorine introduction into polymer chain is considered as one of the most effective ways for simultaneous reducing HOMO and LUMO levels, while maintaining the bandgap almost constant in comparison to their corresponding non-fluorinated analogues, which significantly increase Voc and PSC efficiency as a result, and also increase of intra-intermolecular interaction of polymers due to strong induced C–F bond that leads to high charge mobility and transport. Furthermore strong F⋯H/F⋯S interaction and well-resolved fluoropolymers fibril structure can make significant contribution to active level morphology and to suppress charge recombination. In this regard, fluoride utilization can lead to synergism of the resulting material optoelectronic characteristics.
Inspired by these results, we have prepared two “D-A1–D-A2” structure copolymers P1 and P2 fluorinated analogue by adding two fluorine atoms to thiadiazoloquinoxaline fragment (Scheme 2). In this work, two thiadiazoloquinoxaline (TDQ) and benzodithiazole (BT) electron-acceptor units were included into conjugated P1 and P2 polymers for PSC applications. TDQ is a strong electron acceptor due to its rigid planar structure and the presence of four imino groups with unique properties such as high charge carrier mobility, extending absorption up to near infrared region and ease of functionalization with various aryl and alkyl groups for improving solubility. TDQ-based conjugated polymers demonstrated conversion efficiency of more than 3.58% with photoresponse to 1200 nm.15 BT-based polymers exhibit broad absorption up to near-infrared region up to 1000 nm with efficiency of more than 5–6%.16–18 It was interesting to investigate D–A conjugated polymers containing both TDQ and BT units. As usual, two thiophene groups are introduced on the both sides of the BT group to reduce the steric hindrance between two acceptor units, which also affects the properties such as charge carrier mobility, optical absorption and energy levels of the resulting copolymers. Novel P1 and P2 polymers (Scheme 1) have narrow bandgaps. PSCs based on them have a strong tendency to form crystalline structures with photoresponse in thin films up to 1100 nm. Bulk heterojunction polymer solar cells were fabricated using these copolymers as donors along with PC71BM as acceptor to investigate the effect of substitution of fluorine atoms in TDQ acceptor units of polymer backbone. After the optimization of weight ratios and solvent concentration, P2 possesses superior photovoltaic characteristics with PCE of 7.21% compared to its non-fluorinated analog P1 with PCE of 5.30%. Our results show that conjugated polymers with electron-acceptor TDQ and BT units may demonstrate narrow bandgap and strong semi-crystallinity and therefore can potentially be used as electron donor in high performance organic electronic devices.
The polymerization was performed by a Stille coupling reaction. In a 50 mL flask, M1 (0.6738 g, 0.5 mmol) and M3 (0.3130 g, 0.5 mmol) were dissolved in 15 mL toluene, and the solution was flushed with argon for 15 min, then 27 mg Pd(PPh3)4 was added into the solution. The mixture was again flushed with argon for 20 min. The reaction mixture was heated to reflux for 48 h. The reaction mixture was cooled to room temperature and added dropwise to 400 mL methanol. The precipitate was collected and further purified by Soxhlet extraction with methanol, hexane, and chloroform in sequence. The chloroform fraction was concentrated and added drop-wise into methanol. Finally, the precipitates were collected and dried under vacuum overnight to get polymer P1 as a black solid (0.52 g, yield 71%). 1H NMR (CDCl3, 400 MHz): δ (ppm) 8.00–7.40 (br, 20H, ArH), 2.20–1.95 (br, 8H, CH2), 1.90–0.70 (br, 92H, CH3, CH2). Elem. anal. for (C96H120N6S4)n calc.: C, 77.58; H, 8.14; N, 5.65; S, 8.63. Found: C, 77.19; H, 8.18; N, 5.34; S, 8.19.
The hole-only and electron-only devices with ITO/PEODT:
PSS/copolymer
:
PC71BM/Au and Al/copolymer
:
PC71BM/Al architectures were also fabricated in an analogous way, in order to measure the hole and electron mobility, respectively. The impedance data were measured with an impedance analyser (frequency 1 mHz and 1000 kHz), under illumination.
The composition and structure of intermediate products 2–9 and target compound M2 were confirmed by elemental analysis, 1H, 13C and 19F NMR spectroscopy. In particular, there are three multiplets in the proton spectrum of M2 at δ 7.02–7.10 (m, 4H), 7.70–7.62 (m, 4H), 7.80–7.90 (m, 4H), related to the 12 aromatic protons, triplet at δ 1.86 (t, J = 8.19 Hz, 8H), related to the methylene groups directly adjacent to the fluorene moiety, whereas in the interval of δ 1.40–0.60 ppm there are signals related to the other 92 protons of the alkyl substituents. Although the 1H NMR spectrum is rather complicated, the ratio of the intensity values of the aromatic part of the spectrum to aliphatic corresponds well to the proposed structure (Fig. S13 and ESI†). The carbon spectrum of M2 contains 10 signals in the range of 113–165 ppm related to 10 different aromatic quaternary carbons. Signals of the atoms C3, C1 and C6 are doublets with JC–F = 247, 7.6 and 1.6 Hz, respectively, due to the spin–spin interaction with F atom. In the region of 108–132 ppm there are 6 signals related to six tertiary carbon atoms, wherein the signals of atoms C5, C4 and C2 are doublets with JC–F = 9.0, 23.0, 23.0 Hz, respectively, due to the spin–spin interaction with a fluorine atom. In the region of 55.46 and 14.11 ppm there are characteristic signals related to the carbon atom of cyclopentadiene moiety of fluorine and terminal CH3 groups of the alkyl substituents, respectively. In the region of 41–22 ppm there are signals related to the rest of the aliphatic carbon atoms, which once again confirms the proposed structure (Fig. S13b, ESI†). The 19F NMR spectra contain a singlet signal at −112.94 ppm relating to fluorine atom of M2 (Fig. S13c, ESI†).
The synthetic routes for copolymers are shown in Scheme 2 and the detailed synthetic process and characterization are described in the ESI.† Copolymer P1 and P2 was synthesized in good yields of 71–76% via Stille cross-coupling reaction of corresponding dibromo monomers (M1 and M2) with bisstanyl monomers M3 using toluene as the solvent and Pd(PPh3)4 catalyst at 110 °C. The copolymers were purified by Soxhlet extraction with methanol, hexane and chloroform to remove small molecular and residual catalyst. Chloroform fraction was precipitated in methanol again and dark purple polymer was obtained after dried in a vacuum oven. Copolymeric were characterized by 1H NMR spectroscopy (Fig. S14a and b, ESI†) and elemental analysis. Both the copolymers are readily soluble in common organic solvents such as chloroform, chlorobenzene and o-dichlorobenzene. The number-average molecular weights (Mn), weight-average molecular weight (Mw) and polydispersity indices (PDIs) of copolymers were determined by gel permeation chromatography (GPC) analysis with a polystyrene standard calibration and chloroform as the eluent. P1 and P2 display high Mn 14.7 and 18.6 kDa, and PDI 1.93 and. 1.81, relatively (Table 1). The molecular weight of conjugated copolymers plays a key role in determining their optical, electrical properties and charge carrier mobilities. A high molecular weight is required for the polymer to ensure the long conjugation length and promote the intermolecular charge transport and consequently enhance the device efficiency.
Copolymer | Yield (%) | Mn (kDa) | Mw (kDa) | PDI Mw/Mn | Td (°C) |
---|---|---|---|---|---|
P1 | 71 | 14.7 | 28.3 | 1.93 | 336 |
P2 | 75 | 18.6 | 33.7 | 1.81 | 347 |
Copolymer | λsolvmax (nm) | λfilmmax (nm) | Eopg (eV) | Eoxonset (V) | Eredonset (V) | HOMO (eV) | LUMO (eV) | Eecg (eV) |
---|---|---|---|---|---|---|---|---|
P1 | 313, 400, 904 | 313, 400, 952 | 1.08 | 0.60 | −0.67 | −5.08 | −3.81 | 1.27 |
P2 | 313, 400, 908 | 313, 398, 946 | 1.11 | 0.68 | −0.68 | −5.16 | −3.80 | 1.36 |
EHOMO = −q (Eoxonset + 4.48) eV |
ELUMO = −q (Eredonset + 4.48) eV |
Eelecg = EHOMO − ELUMO |
The calculated EHOMO and ELUMO energy levels from the onset oxidation and reduction potential of P1 are −5.08 and −3.81 eV, respectively. After adding two fluorine atoms, P2 shows a much deeper HOMO level of −5.16 eV, whereas the LUMO level (−3.80 eV) is almost unchanged. Evidently, the electron-withdrawing nature of the fluorine atom lowers the HOMO energy level of the fluorinated polymer compared with that of the non-fluorinated analog. As expected, this indicated that the introduction of two fluorine atoms onto the thiadiazoloqunoxaline unit can effectively lower the HOMO energy level of the copolymer by ∼0.08 eV, which leads to an increase in Voc of BHJ devices fabricated by using these fluorinated materials and to ensure better environmental stability of the material. On the other hand, LUMO energy levels of P1 and P2 were calculated to be −3.81 eV and −3.80 eV, respectively. Thus, one can find that the LUMO energy levels were almost not affected by the introduction of fluorine atoms. Both LUMO energy levels are 0.3 eV higher than that of the acceptor PC71BM (−4.2 eV), ensuring energetically favorable electron transfer from the polymer donor to the PC71BM acceptor in PSCs. Electrochemical band gaps are then calculated to be 1.27 and 1.36 eV for P1 and P2 respectively. On account of the larger influence on the HOMO energy level in comparison with the LUMO, the electrochemical bandgap (Eelecg) slightly increases upon fluorination. The difference between the optical and electrochemical band gap could be explained by the exciton binding energy of the copolymers and/or the interfacial barriers for charge injection.21,22
TD-DFT excited state calculations were performed to calculate the optical gaps of the P1 and P2 using the same functionals and basis set on the corresponding ground state structures. The UV/vis spectra were calculated using the B3LYP and M06 functionals. The first rounds of geometry optimization were performed using the Turbomole package.29 All of the follow up calculations were performed using the Gaussian package.28
The first rounds of calculations were the geometry optimizations of the P1 and P2 structures. To increase the computational efficiency the alkyl groups were truncated to ethyl groups. Vibrational analysis on all of the optimized structures did not reveal any vibrational modes with imaginary Eigen frequencies, i.e. the final optimized structures are true local (if not global) minima. The main part of the structures that consists of the benzothiadiazole (BT) and thiadiazoloquinoxaline (TDAQ) moieties, as well as the linking thiophenes, are planar. We have calculated the HOMO and LUMO energy levels and the optical gaps, defined here as the energetically lowest allowed vertical electronic excitation, employing the PBE, M06, and B3LYP functionals. In Table 3, in addition to the frontier orbitals' energy levels, we also provide the optical gap, the main contributions to the first excitation, as well as the wavelength of the first excitation and of the excitations with the largest oscillator strengths.
HOMO (eV) | LUMO (eV) | HL (eV) | OG (eV) | λ1st/max (nm) | f | Main contributions | μ (D) | |
---|---|---|---|---|---|---|---|---|
a Values when solvent effects are taken into account for chloroform. | ||||||||
P1 | ||||||||
PBE | −4.61 | −3.67 | 0.93 | 1.33 | 931 | 0.45 | H → L (89%), H → L+1 (6%) | 4.76 |
−4.74a | −3.79a | 0.95a | 1.27a | 975a | 0.62a | H → L (94%)a | 6.24a | |
B3LYP | −5.16 | −3.26 | 1.89 | 1.68 | 736/581/530/401/349 | 0.62 | H → L (99%) | 4.53 |
−5.29a | −3.37a | 1.91a | 1.63a | 762/580/550/390/351a | 0.75a | H → L (99%)a | 5.69a | |
M06 | −5.44 | −3.19 | 2.25 | 1.77 | 700/546/496/336/324 | 0.63 | H → L (98%) | 4.67 |
−5.59a | −3.31a | 2.28a | 1.72a | 721/546/511/378/337a | 0.76a | H → L (98%)a | 5.88a | |
![]() |
||||||||
P2 | ||||||||
PBE | −4.67 | −3.74 | 0.92 | 1.32 | 936 | 0.44 | H → L (89%), H → L+1 (6%), H−1 → L (5%) | 3.48 |
−4.76a | −3.82a | 0.94a | 1.27a | 978a | 0.61a | H → L (94%)a | 4.79a | |
B3LYP | −5.22 | −3.34 | 1.88 | 1.67 | 741/579/517/411/352 | 0.61 | H → L (99%) | 2.97 |
−5.31a | −3.41a | 1.91a | 1.62a | 764/578/533/351/341a | 0.75a | H → L (99%)a | 3.72a | |
M06 | −5.50 | −3.27 | 2.24 | 1.76 | 705/544/487/394/337 | 0.62 | H → L (98%) | 3.09 |
−5.63a | −3.34a | 2.29a | 1.73a | 717/541/494/367/337a | 0.78a | H → L (98%)a | 3.23a |
In addition to the B3LYP functional we have also performed our calculations employing the M06 functional. The M06 meta-hybrid functional was chosen since it provides leveled performance over transition types.30,31 We provide results using all three functionals, which can additionally be used for comparison with the literature. The HOMO–LUMO (HL) gap of each structure calculated using the hybrid B3LYP functional is notably smaller, by ∼0.35 eV, than that using the meta-hybrid M06 functional however the calculated optical gaps are only marginally smaller, with a difference ∼0.1 eV. In Table 3 we also provide the character of the first allowed excitations only for contributions larger than 4%. The first excitation, as calculated by each of the functional for all three structures, clearly exhibits a single-configuration character.
In Fig. 4, we have plotted the isosurfaces (isovalue = 0.02) of the HOMO and LUMO for all three structures. For both of the structures the HOMO extends over the main body and the LUMO is delocalized mainly over the TDAQ moieties. To quantify the contributions of the moieties to the frontier orbitals we have calculated the total and partial density of states (PDOS). The PDOSs for P1 and P2 are shown in Fig. S15.† We partition all of the structures into the linking thiophenes, the TDAQ and BTDA moieties and aliphatic groups for both structures, and additionally to fluorene (FL) for P1 and di-fluoro fluorene (FFL) for P2. Both structures have high contributions to the HOMO from the TDAQ, and BTDA, specifically at 27.9% and 24.6% for P1, respectively, and at 26.4% and 25.5% for P2, respectively. The linking thiophenes also have high contributions to the HOMO, at 46.5% for P1 and at 45.5% for P2, which results in an extensive delocalization along the main chain of the structure. The situation is different for the LUMO of the structures, which in both cases extends almost exclusively over the TDAQ moiety, at 72.9% for P1, and at 74.2% for P2. The remaining minor contributions are from the BTDA and linking thiophene moieties at 11.2% and 10.6% for P1, and 11.4% and 10.3% for P2. These are in agreement with our earlier observations on the orbital delocalization.
In Fig. 5, we show the UV/Visual absorption spectra of the P1 and P2 structures calculated at the TD-DFT/M06 level of theory, both accounting for solvent effects for CF and in gas phase. The spectra have been produced by convoluting Gaussian functions with HWHM = 0.22 eV centered at the excitation wavenumbers. In Fig. S16 (see ESI†) we also provide the corresponding spectra calculated using the B3LYP functional which is in good agreement with the spectra using the M06 functional and only slightly overestimate in the long wavelength region by ∼30 nm and in the short wavelength region ∼10 nm.
![]() | ||
Fig. 5 Theoretical UV/vis absorption spectrum of (a) P1, and (b) P2 (calculated using the M06 functional). |
The calculated absorption spectra of P1 and P2 have significant similarities; they exhibit three main bands of high absorbance; one centered at large wavelengths ∼720 nm, as well as two centered at smaller wavelengths at 340 nm and ∼520 nm. In the ESI† we provide all of the high intensity peaks with the corresponding oscillator strengths (Table S1 and S2†).
![]() | ||
Fig. 6 (a) Current density–voltage characteristics under illumination of AM1.5G, 100 mW cm2 and (b) IPCE spectra of the PSCs based on copolymer![]() ![]() ![]() ![]() |
X-ray diffraction (XRD) measurements were carried out for two copolymer films in order to characterize their structure, such as crystallite size, crystallite orientation and intermolecular distance. The XRD profiles of P1 and P2 films processed with CF solvent were shown in Fig. 7a and shows the semi-crystalline nature of the copolymers. Both the polymers show pronounced (100) reflection peaks, indicating the ordered lamellar stacking of the copolymer chains in solid films. The (100) reflection peaks at 2θ = 4.78° and 4.96° for P1 and P2, respectively, corresponds to lamellar spacing of 19.38 Å and 18.46 Å, respectively. The lamellar spacing of P1 is slightly larger than P2 attributed to the fact that F atom in the fluorene units extends the length of side chains in the P2 film so as to stretch the lamellar distance of polymer chains in P2 film. Both the (010) reflection peaks of P1 and P2 films were located at 2θ = 23.38° which corresponds to the π–π stacking distance of 4.46 Å in both P1 and P2 films. The π–π stacking distance of two copolymers are almost the same, suggesting that the introduction of F atom in fluorene unit on copolymer side chains has little influence on π–π stacking in copolymer film. However, the intensity of the both (010) and (100) reflection peaks of P2 film is stronger than that of P1 film, indicating a more ordered π–π stacking and crystallinity of P2 as compared to P1. This indicates that the introduction of fluorine atoms into copolymer backbones has a significant effect on the crystalline property of copolymers as well as the resulting device performance.
![]() | ||
Fig. 7 X-ray diffraction pattern of (a) pristine P1 and P2 and (b) P2![]() ![]() |
The charge transport property in the PSCs is generally one of the important factors for determining the Jsc and FF. The hole mobility was measured in the hole only devices using space charge limited current (SCLC) method (Fig. 8). The hole mobility of P1 and P2 in the BHJ active layers are 2.78 × 10−5 cm2 V−1 s−1 and 5.76 × 10−5 cm2 V−1 s−1, respectively. However, the electron mobility in the BHJ active layer based on P1 and P2 are almost same (2.46 × 10−4 cm2 V−1 s−1). The higher PCE for PSC based on P2 than P1, may be attributed to the larger hole mobility in the P2:
PC71BM as compared to that for P1
:
PC71BM, leading the better balanced charge transport. The higher value of hole mobility is consistent with the more crystalline nature of P2 as inferred from the XRD data.
The devices based on these two devices exhibited low PCE values, which are due to the low Jsc and FF values. This might be attributed to poor nanophase morphology and poor exciton dissociation and charge transport. We have used the solvent additive method to improve the PCE of the PSCs. Therefore, PSCs based on P1 and P2 were fabricated with the same blend ratio and different amounts of CN additive were tried and we found that a 3% by volume gave the best photovoltaic response. The J–V characteristics of the devices under illumination are shown in Fig. 6a. The PCE of the devices significantly improved with the introduction of the solvent additive, leading to enhancement of the FF and Jsc at a slight loss in the Voc (for P1 Jsc = 10.86 mA cm−2, Voc = 0.74 V, FF = 0.66, PCE = 5.30%) and (for P2 Jsc 12.34 mA cm−2, Voc = 0.80 V, FF = 0.68, PCE = 7.21%). These values are higher than that reported in literature using similar copolymers.15–17 The increase in Jsc and FF is associated with loss in Voc, mainly because CN possesses a high boiling point and is able to solvate the PC71BM leading to ample impact on Jsc by providing a more optimal morphology for facilitating charge transportation.32,33 However the loss of Voc can be attributed to the lowering of charge-separated and charge-transfer-state energies upon additive addition.34 Simultaneously, we speculate that the addition of CN decreases the size of fullerene domains and facilitates the formation of a bicontinuous inter-penetrating donor–acceptor network (as discussed in the later part of the discussion).
In order to get the accuracy of the photovoltaic performance, incident photon to current conversion efficiency (IPCE) spectra of the PCEs with P1 and P2 were measured and shown in Fig. 6b. The IPCE spectra cover a broad band spectral response range from 300 nm to 300 nm in agreement with the corresponding absorption spectra of active layer (Fig. S16, ESI only for P2:
PC71BM†). The PSC based on P2 showed higher values of IPCEs than the P1, which is consistent with the higher value of Jsc for P2 based device. The Jsc values estimated from the integration of the IPCE spectra (as shown in Table 4) agree well with those obtained from J–V characteristics.
To get information about the effect of the CN additive on the hole and electron transport properties of the active layers, we also investigated the hole and electron mobilities of P1:
PC71BM and P2
:
PC71BM blend films by SCLC method (Fig. 8 for hole only device). After the CN additive, both P1
:
PC71BM and P2
:
PC71BM blended films exhibited an increased hole mobility of 8.83 × 10−5 cm2 V−1 s−1 and 1.45 × 10−4 cm2 V−1 s−1, respectively. However, the electron mobility of the active layers does change much after using the CN additive. The increase in hole mobility and similar value of electron mobility leads to a more balanced charge transport within the active layer and is one of the main reasons for enhanced Jsc, FF and PCE of the corresponding PSCs.
We have used impedance spectroscopy (IS) to investigate the electrical characteristics of donor/acceptor interfaces, active layer/electrode interfaces and series resistance using an equivalent circuit models in PSCs.35,36 Fig. 9a shows the Nyquist plots of IS for the device for P2:
PC71BM active layer processed with and without CN additives and fitted with the equivalent circuit model (Fig. 9b). The corresponding device parameters are derived by fitting the experimental Nyquist plots with equivalent circuit. The shunt pair R1 and C1 corresponds to the D/A interfaces within the BHJ active layer film whereas, the other shunt pair of R2 and C2 corresponds the interfaces of BHJ active layer with anode (PEDOT
:
PSS) and cathode (Al). R3 corresponds to the resistance of electrodes and wires. The PSC based on the approach without CN additives exhibits higher R1 (1038 Ω) and R2 (206 Ω) values and lower C1 (6.81 × 10−8 F) and C2 (3.46 × 10−7 F) values than that when using CN additives (R1 = 834 Ω, R2 = 158 Ω, C1 = 6.81 × 10−8 F and C2 = 6.38 × 10−8 F). These differences can be attributed to the variation of interface area between the P2 and PC71BM in the active layer processed with and without CN. The better nanoscale morphology of the active layer processed with CN additive corresponds to the smaller R1 and higher C1 compared to that for without CN additive counterpart. Moreover, the value of R3 is lower for the active layer processed with CN as compared to the approach without additive. These findings from the IS measurements provide strong evidence that favorable nanoscale morphology in the active layer processed with CN additive give rise to the observed improvement in the Jsc and FF in PSCs. The BHJ active layer can be considered as a shunt pair of a resistance and capacitance in equivalent circuit, the average carrier transition time at shunt conditions corresponds to the carrier lifetime (τ in the active layer). The longer τ means a lower recombination rate and a better chance that carriers could reach the electrodes. The τ values are 2.8 × 10−4 s and 7.06 × 10−5 s (τ = R1C1) for the devices processed with and without CN additives, respectively. The longer value of τ for the device processed with CN additive than without additive suggests that the charge carriers in the former blend have a greater opportunity to reach the respective electrodes resulting in higher Jsc and FF.
The nanoscale morphology of the active layer plays an important role in the photovoltaic performance of the device. Proper morphology is not only beneficial for exciton dissociation into free charge carriers but also necessary for charge transport to respective electrodes for efficient collection. The atomic force microscopy (AFM) images in tapping mode for the blended films cast from with and without additive are shown in Fig. 10. It can be seen from these images that the films cast from CF only showed inferior nanoscale morphology of high surface roughness (rms values are 4.02 nm and 3.74 nm for P1 and P2, respectively), which may destroy the continuous percolative pathways for hole and electron transport to the corresponding electrodes, consequently increasing the chance of recombination of charge carriers and reducing the Jsc accounting for the low PCE. Moreover, the high surface roughness may also reduce the D/A interfacial area in the active blend layer and hamper the exciton dissociation. However, the surface roughness of the blended films was significantly reduced (2.12 nm and 1.54 nm for P1 and P2, respectively), when processed with solvent additive. The reduction in the surface roughness might lead to an increase the exciton dissociation and charge transport that resulted in the improvement of both Jsc and FF.
![]() | ||
Fig. 10 AFM images (3 × 3 μm) without (top) and with CN additives (bottom) (a and c) for P1![]() ![]() ![]() ![]() |
The morphology of the active layer was further investigated using transmission electron microscopy (TEM) image from which are shown in Fig. 11. Copolymer:
PC71BM films without CN additive have coarse dispersive domains that can be a factor in hampering exciton diffusion and charge transport. However, the morphology of the film after adding CN was significantly changed. The large crystallites and phase separation domains disappeared and a fine nanoscale morphology within the exciton diffusion length was observed, resulting in a favorable exciton dissociation and charge transport. The film P2
:
PC71BM showed a much finer and denser texture and nanoscale separation between copolymer and PC71BM, which could result in a high PCE with larger Jsc value.37
![]() | ||
Fig. 11 TEM images (3 × 3 μm) without (top) and with CN additives (bottom) (a and c) for P1![]() ![]() ![]() ![]() |
The crystallinity of the active layer is another factor that affects the charge transport and collection, thereby disturbing the overall performance of the BHJ organic solar cell. To obtain information about the change in crystallinity of the copolymer:
PC71BM, we recorded the X-ray diffraction patterns of the blend films processed with CF and CN/CF solvents (Fig. 7b, only shown for P2
:
PC71BM). The film cast from CF showed a broad diffraction peak at 2θ = 4.96°, corresponding to the crystalline domain of P2. For the film cast from CN/CF this diffraction peak becomes narrower, that is related to the increase in the crystalline nature of P2 in the active layer that facilitated the hole transport in the active layer more efficiently.
In order to assess the effect of the solvent additive on the exciton generation, dissociation and charge transport properties, we have measured the photocurrent density (Jph) versus the effective voltage (Veff) of the devices and the results are shown in Fig. 12. Jph is defined as Jph = JL − JD, where JD and JL are the current density in dark and under illumination, respectively. Veff is defined and Vo − Vappl, where Vo is the voltage at which the photocurrent is zero and Vappl is the external applied voltage.38 The Veff determines the electric field in the bulk region and thereby affects the carrier transport and photocurrent extraction. At low Veff, the Jph varies linearly with Veff and then tends to saturate as the Veff further increases. At high Veff values mobile charge carriers quickly move toward the corresponding electrode with minimum recombination. At full saturation, we assume that all the excitons generated after the absorption of light by the active layer are dissociated into free charge carriers and consequently collected by the respective electrodes. The Jph saturate at lower Veff for the device processed with solvent additive as compared to the device processed without solvent additive, indicating that the low Veff is needed for the exciton dissociation for former device. Therefore at high Veff, the saturation photocurrent density (Jphsat) is limited by the total number of absorbed photons and independent of bias voltage. We have estimated the maximum generation rate of free charge carriers (Gmax) according to Jphsat = qGmaxL, where q is the electronic charge and L is the thickness of the active layer.39 The values of Gmax for the devices based on P1:
PC71BM (as cast), P1
:
PC71BM (SA) and P2
:
PC71BM (as cast) and P2
:
PC71BM (SA) are 8.4 × 1021 cm−3, 9.11 × 1021 cm−3, 9.68 × 1021 cm−3 and 10.25 × 1021 cm−3, respectively. Gmax is related to the number of photon absorbed by the active layer and such changes in the Gmax are consistent with the absorption spectra of corresponding active layers (see the Fig. S16†). The charge collection probability Pc has been estimated according to Pc = Jsc/Jphsat (ref. 40) and are 0.67, 0.78, 0.73 and 0.84 for P1
:
PC71BM (as cast), P1
:
PC71BM (SA), P2
:
PC71BM (as cast) and P2
:
PC71BM (SA), respectively. The higher value of Pc for the devices with solvent additives than the corresponding devices without solvent additives is attributed to the better phase separation, increased hole mobility and improved balance in charge transport, and enhancement in light harvesting efficiency.
![]() | ||
Fig. 12 Variation of photocurrent density (Jph) with effective voltage (Veff) for the PSCs based in different active layers. |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra14537j |
This journal is © The Royal Society of Chemistry 2016 |