Next Article in Journal
Characterization of Camphene- and Fenchol-Based Hydrophobic Eutectic Solvents and Their Application in Aldehyde Extraction
Previous Article in Journal
A Novel Ternary Catalyst PW4@MOF-808@SBA-15 for Deep Extraction Oxidation Desulfurization of Model Diesel
Previous Article in Special Issue
Introduction of Electron Donor Groups into the Azulene Structure: The Appearance of Intense Absorption and Emission in the Visible Region
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Impact of Different π-Bridges on the Photovoltaic Performance of A-D-D′-D-A Small Molecule-Based Donors

by
Lingjun Yang
1,†,
Yu Wu
1,2,†,
Pachaiyappan Murugan
1,†,
Peng Liu
1,
Yulong Peng
1,
Zhiyong Qiu
1,
Zaifang Li
2,*,
Changlin Yu
3 and
Shiyong Liu
1,*
1
Jiangxi Provincial Key Laboratory of Functional Molecular Materials Chemistry, Department of Chemistry and Chemical Engineering, Jiangxi University of Science and Technology, Ganzhou 341000, China
2
China-Australia Institute for Advanced Materials and Manufacturing (IAMM), Jiaxing University, Jiaxing 314001, China
3
Guangdong Provincial Key Laboratory of Advanced Green Lubricating Materials, Maoming 525000, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2024, 29(17), 4231; https://doi.org/10.3390/molecules29174231
Submission received: 14 August 2024 / Revised: 2 September 2024 / Accepted: 4 September 2024 / Published: 6 September 2024
(This article belongs to the Special Issue π-Conjugated Functional Molecules & Polymers)

Abstract

:
Three small donor molecule materials (S1, S2, S3) based on dithiophene [2,3-d:2′,3′-d′]dithiophene [1,2-b:4,5-b′]dithiophene (DTBDT) utilized in this study were synthesized using the Vilsmeier–Haack reaction, traditional Stille coupling, and Knoevenagel condensation. Then, a variety of characterization methods were applied to study the differences in optical properties and photovoltaic devices among the three. By synthesizing S2 using a thiophene π-bridge based on S1, the blue shift in ultraviolet absorption can be enhanced, the band gap and energy level can be reduced, the open circuit voltage (VOC) can be increased to 0.75 V using the S2:Y6 device, and a power conversion efficiency (PCE) of 3% can be achieved. Also, after developing the device using Y6, S3 introduced the alkyl chain of thiophene π-bridge to S2, which improved the solubility of tiny donor molecules, achieved the maximum short-circuit current (JSC = 10.59 mA/cm2), filling factor (FF = 49.72%), and PCE (4.25%). Thus, a viable option for future design and synthesis of small donor molecule materials is to incorporate thiophene π-bridges into these materials, along with alkyl chains, in order to enhance the device’s morphology and charge transfer behavior.

Graphical Abstract

1. Introduction

The power conversion efficiency (PCE) of single-junction polymer organic solar cells (PSCs) has been substantially enhanced in recent years due to the introduction of novel photovoltaic organic materials, particularly Y-series non-fullerene small molecule materials [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17]. The industrialization of mass production has brought about a new challenge: even though polymer materials have an excellent process and high efficiency when it comes to creating membranes [18,19,20,21,22], because their chemical structure is not very apparent and their repeatability is poor, these problems in the manufacturing process can be better addressed by small molecules [23,24,25,26]. There has been significant advancement in the PCE of all small molecule organic solar cells (ASM OSCs) recently [27,28,29,30,31,32,33,34], but the performance of these cells is still inferior to that of PSCs. This is primarily because the donor and acceptor materials have a short and similar acceptor–donor–acceptor (A-D-A)-type molecular structure, which results in similar physical and chemical properties and makes it difficult to control the crystallinity of these materials. Because it is difficult for the blending membrane to create a nano-scale interpenetrating network structure, both short-circuit current (JSC) and filling factor (FF) are low.
An important component of the donor (D′) in the DTBDT unit—which stands for dithiophene [2,3-d:2′,3′-d′]dithiophene [1,2-b:4,5-b′]dithiophene—is capable of donating electrons and transporting carriers with remarkable efficiency [35,36,37,38,39,40,41,42,43,44]. By incorporating the 4-(2-ethylhexyl)-4H-dithieno [3,2-b:2′,3′-d]pyrrole (DTP46) into the molecular framework, the transport efficiency of carriers can be enhanced. It is well known that adding π-bridges to materials with tiny donor molecules changes their absorbance and crystallinity, and this is because DTP46’s stiff planar-conjugated structure promotes the coplanarity of molecules [45,46,47,48,49,50,51]. In its interaction with the central conjugated unit, 3-butyl-2-thioxothiazolidin-4-one (RDN-34) can form an A-D-A type structure that, thanks to its electron absorption capacity (A), has a superior intramolecular charge transfer effect (ICT) effect; this process can also prolong electron cloud delocalization [52,53,54,55,56,57,58,59,60,61,62]. By incorporating thiophene as a π-bridge into the small donor material, the JSC and FF of organic solar cells (OSCs) can be enhanced, resulting in a larger PCE. Furthermore, via the addition of alkyl chains to the thiophene π-bridge, the π-π packing of the molecules can be enhanced, resulting in more compact layer packing and enhanced crystallization and aggregation of the material [63,64,65,66,67].
This study details the successful creation of three tiny donor molecular materials based on DTBDT, designated as S1, S2, and S3 (Scheme 1). S2 is an example of this; it expands the initial conjugated plane by adding a thiophene π-bridge to S1, which is based on S1. In addition to improving charge transport and exciton ionization, this also leads to an improvement in the morphology of the composite film, which results in an increase in JSC (8.08 mA/cm2 to 9.23 mA/cm2), FF (42.16% to 42.93%), and PCE (2.42% to 3%). A combination of an alkyl chain and the thiophene π-bridge allows for the production of S3, as S1 and S2 are poorly soluble in common organic solvents such as dichloromethane, trichloromethane, and chlorobenzene. When S3 and acceptor molecule Y6 are mixed, the film becomes free of insoluble particles. The device has several advantages, such as better carrier mobility, the ability to suppress bimolecular recombination, improved exciton dissociation and charge collection, and more balanced mobility. As a consequence of this, the S3:Y6 devices exhibit an improved open circuit voltage (VOC = 0.80 V), JSC (10.64 mA/cm2), and FF (49.72%), while simultaneously succeeding in achieving an ideal PCE (4.25%).

2. Results and Discussion

2.1. Synthesis and Characterization

The three-part synthesis method of S1, S2, and S3 is depicted in Scheme 2 and Scheme 3. Following the pre-synthesis of the DTP46 precursor (Scheme 2), the initial step was to produce an aldehyde intermediate 1 (DTP46-CHO) using the Vilsmeier–Haack reaction of DTP46. Subsequently, the DTP46-CHO was brominated with the N-bromosuccinimide (NBS) reagent, which resulted in the formation of intermediate 2 (Br-DTP46-CHO). Traditional Stille coupling is employed to synthesize aldehyde intermediates 3 and 4 with thiophene π-bridges in order to investigate the impact of thiophene π-bridges on small donor molecule materials. Derivatives 5 and 6 are produced by brominating the 3 or 4 using NBS reagent. Then, 7, 8, and 9 are formed by combining 2, 5, or 6 with BDT-2SnMe3, which is 4,8-(5-(2-(thio-2-ethylhexyl))4-chlor-othiophene-2-yl)-[1,2-b:4,5-b′]thiophene-[3,4-c]dithiobenzene, using the classic Stille coupling method followed by the target products S1S3 were synthesized by Knoevenagel condensation. S1 and S2 do not dissolve well in common solvents like dichloromethane, trichloromethane, chlorobenzene, and others because they have fewer alkyl chains. When mixed with acceptor Y6, the resulting film has small particles that cannot dissolve. S3 was synthesized by replacing the original thiophene π-bridge with thiophene with an alkyl chain based on S2. This resulted in S3 having good solubility in common solvents. The great purity of all these tiny donor molecule materials is confirmed by 1H NMR (Supplementary Information, Figures S1–S12).

2.2. Opto-Electronic Properties

Table 1 displays the detailed data, and Figure 1a,b displays the results of the UV-vis absorption spectra testing of three small donor molecules (S1, S2, and S3) in a CHCl3 solution and thin film. S1, S2, and S3 display a broad and extensive absorption spectrum between 450 nm and 650 nm in a CHCl3 solution. The blue shift in S2’s ultraviolet absorption curves can be explained by the fact that the electron-donor central unit BDT-2SnMe3 and the peripheral weakly electron-absorbing RDN-34 terminal group have a weaker D-A interaction, as revealed by comparing the UV absorption curves of S1 and S2. The addition of alkyl chains to the thiophene π-bridge enhances the D-A effect between groups and the degree of molecular chain order. As a result, S3 exhibits a slight redshift phenomenon and a similar absorption trend to S1 [68]. This phenomenon is further supported by the A0-0/A0-1 ratio, which measures the relative sizes of the right and left acromion, and the A0-0/A0-1 ratio of S2 is smaller compared to other small donor molecules. The addition of the thiophene π-bridge reduces the strength of the D-A effect between the groups, as indicated by the weak π-π interaction between the chains. With the introduction of thiophene π-bridge, the optical bandgaps of S1, S2, and S3 gradually narrow (1.75 to 1.69 eV) in Table 1. In solid films, S1, S2, and S3 exhibit an absorption range that is wider and wider than in solution (430–720 nm). Furthermore, the maximum absorption is red-shifted by more than 90 nm in comparison to the solution. This is typically the result of enhanced main chain intermolecular interactions of small donor molecules and stronger aggregation effects in condensed matter.
The cyclic voltammetry (CV) technique was utilized in order to determine the frontier molecular orbital (FMO) levels of S1, S2, and S3, as can be seen in Figure 1c. In Figure 1d, it is demonstrated that the addition of a thiophene π-bridge lowers the highest occupied molecular orbital (HOMO) level of small donor molecule materials (−5.14 to −5.18 eV). The addition of alkyl chains to the thiophene π-bridge further lowers the HOMO level (−5.18 to −5.20 eV). As a result of this gradual decrease in HOMO level, the VOC of S1, S2, and S3 gradually increases (0.71–0.80 V). There is not a significant difference between the HOMO energy levels of S1, S2, and S3 due to the fact that all three of these atoms share the same weak electron-absorbing end group, which is denoted by RDN-34.

2.3. Photovoltaic Performances

During the process of evaluating the photovoltaic performance, the OSCs device was built with a structure consisting of ITO/PEDOT:PSS/Donor:Y6/PDINO/Ag, as demonstrated in Figure 2a. A 1.2:1 donor-to-acceptor ratio, a 0.5% (v/v) additive of 1-chloronaphthalene (1-CN), an annealing temperature of 110 °C, and a rotation speed of 2500 rpm/40 s were the ideal preparation conditions for the optimized device. The J-V curve, which was produced after optimizing the device settings, is shown in Figure 2b and Table 2, where the data acquired are presented in detail. The values of VOC, JSC, and FF increase from 0.71 V, 8.08 mA/cm2, and 42.16% to 0.80 V, 10.64 mA/cm2, and 49.72%, respectively, as the series progresses from S1 to S3. This suggests that the performance of OSC devices can be positively impacted by reasonable molecular design, specifically the addition of an alkyl chain and the thiophene π-bridge, since the PCE of S3 (4.25%) is nearly 1.8 times higher than that of S1 (2.42%). It should be mentioned that S3 has a smaller band gap (1.69 eV) and a lower HOMO energy level (−5.20 V), so the S3:Y6 device has a higher VOC (0.80 V). On the other hand, adding an alkyl chain improves the solubility of the small donor molecule material, which in turn improves the morphology of the device film and leads to better JSC (10.64 mA/cm2) and FF (49.52%). Figure 2c shows that the three small donor molecules all have a strong light absorption from 480 nm to 830 nm in the 300–950 nm range of the external quantum efficiency (EQE) curve. This suggests that adding thiophene as a π-bridge and a sufficiently long alkyl chain to it can enhance the light absorption response intensity of the small donor molecule materials. Each and every integrated JSC that was estimated by utilizing the EQE curves is in agreement with the JSC that was measured, and this agreement is within the error range by a margin of less than five percent.
Using the J-V characteristic curve in dark conditions, the charge transfer process in OSCs devices is examined, as illustrated in Figure 3a. The device based on S3:Y6 exhibits the highest dark current density in the presence of a positive bias. Conversely, when the bias is negative, the dark current density is in the middle. The device based on S3:Y6 possesses a more asymmetrical J-V characteristic curve, which makes charge extraction and collection easier [69]. Additionally, it inhibits current leakage, resulting in a higher JSC (10.64 mA/cm2). Based on the findings presented in Figure 3b, the photogenerated current density profile and the effective voltage curve (JphVeff) were utilized in order to investigate the charge extraction and exciton dissociation capabilities of OSCs. Among them, Jph = JLJD, where JL and JD are the current densities under 100 mW cm−2 light and darkness, and Veff = V0Va, where V0 is the voltage when Jph is zero and Va is the applied voltage. The devices that were based on S3:Y6 had a greater exciton dissociation efficiency, which resulted in increased hole and electron mobility. This was in contrast to the devices that were based on S1:Y6, S2:Y6, and S3:Y6 organic semiconductor capacitors, which had dissociation efficiencies of 46.74%, 58.36%, and 62.91%, respectively.
As shown in Figure 3c,d, the charge recombination mode in OSCs was studied through the curve of VOC and JSC with light intensity. The curve fitting between VOC and light intensity was VOC∝n(KBT/q)lnPlight. When n = 1, it indicated that only bimolecular recombination existed in the device; when n = 2, it indicated that trap-assisted recombination was dominant in the device. The slope of the S3:Y6 device is 2.85 KT/q, which is much lower than that of the S2:Y6 device (4.08 KT/q), indicating that the introduction of alkyl chain can effectively inhibit the trap-assisted recombination behavior based on small donor molecule devices. Similarly, the slope of the S2:Y6 device is smaller than that of the S1:Y6 device (4.12 KT/q), indicating that the introduction of thiophene π-bridge plays a weak positive role in effectively inhibiting the trap assist behavior in the device. The curve fitting between JSC and light intensity was JSC∝Plightα. The α values of the S1:Y6, S2:Y6, and S3:Y6 OSCs devices were 0.854, 0.836, and 0.856, respectively. These α values were similar, indicating that they all had similar weak bimolecular recombination behavior.
As shown in Figure 3e,f, hole mobility (μh) and electron mobility (μe) of the device were measured by the SCLC method to study charge transport in the device, and the detailed data are shown in Table 3. The device with pure hole (ITO/PEDOT:PSS/Donor:Y6 (1.2:1)/MoO3/Ag) structure has been prepared to measure μh and pure electron (ITO/ZnO/Donor:Y6 (1.2:1)/PDINO/Ag) structure of the device measurement of μe. Based on S1:Y6, S2:Y6, and S3:Y6 OSCs devices, μh and μe are 6.19 × 10−4 cm2 V−1 s−1/5.70 × 10−4 cm2 V−1 s−1 (μh/μe = 1.09), 6.31 × 10−4 cm2 V−1 s−1/5.93 × 10−4 cm2 V−1 s−1 (μh/μe = 1.06), and 7.22 × 10−4 cm2 V−1 s−1/6.98 × 10−4 cm2 V−1 s−1 (μh/μe = 1.03), respectively. It can be seen from the data that the device based on S3:Y6 has the highest and balanced μh and μe, so, based on the S3:Y6 device, it has a higher FF.

3. Materials and Methods

3.1. Materials

The BDT-2SnMe3 was acquired from SunaTech Inc (SooChow, China), while the non-fullerene acceptor Y6 was acquired from Shuo Lun Organic Optoelectronic Technology (Jiaxing, China). The electronic transport layer was composed of PDINO, which was acquired from Shanghai Dusen Optoelectronic Materials (Shanghai, China), and the hole transport layer was composed of PEDOT:PSS (4083) from Xi’an Baolaite Optoelectronic Technology (Xi’an, China). Unless otherwise specified, all of the other drugs and reagents that were utilized in the synthesis were acquired from Shanghai Anneji Chemical (Shanghai, China). These pharmaceuticals and reagents can be utilized without the need for additional purification. Chromatoform, methylethanol, zinc acetate dihydrate, and other solvents were purchased from Sigma-Aldrich (St. Louis, MO, USA) in order to be utilized in the process of preparing the ZnO electronic transport layer solution. After treating with calcium hydride, anhydrous chloroform and toluene can be obtained through distillation.

3.2. Characterization

The Bruker 400 and 600 spectrometers (Bruker, Karlsruhe, Germany) were utilized in order to acquire the 1H NMR spectrum in CDCl3. The cyclic voltammetry (CV) was performed with the help of a CHI661C electrochemical workstation, and the UV-vis absorption spectra were analyzed with a Shimadzu UV-2450 spectrophotometer (UV-2600, SHIMADZU, Kyoto, Japan). The CV curve calibration was carried out using an external reference, a ferrocene/ferrocenium (Fc/Fc+) redox couple, and a carbon glass electrode covered with S1, S2, or S3 films. The working electrode was placed in a 0.1 mol/L acetonitrile solution of tetrabutylammonium phosphorus hexafluoride (Bu4NPF6). The reference electrode was an Ag/Ag+ electrode, and the working electrode was a carbon glass electrode.

3.3. Synthesis of Compound 1 (DTP46-CHO)

A reaction tube with a capacity of 25 milliliters was initially filled with DMF (90.27 mg, 1.2351 mmol, 1.2 equiv.) and POCl3 (189.40 mg, 1.2351 mmol, 1.2 equiv.) in an atmosphere of argon. The reaction tube was then stirred at 0 ℃ for 4 h after dissolving DTP46 (300 mg, 1.0293 mmol) in DCE (2.64 mg, 26.6580 mmol, 25.9 equiv.) and was stirred for 10 min. Then, for the quenching step, we filled the reaction tube with 5 mL of saturated potassium acetate water, stirred it, and let it sit at room temperature for 2 h. After that, we used CH2Cl2 and saturated salt water to extract the reaction liquid. Next, we spun the organic phase that had been purified on a rotary evaporator with a 3:1 ratio of dichloromethane or petroleum ether. Thin-layer chromatography was used to purify the dark green viscous compound 1, which had a mass of 258.5 mg and a yield of 78.6%. 1H NMR (400 MHz, CDCl3): δ 9.82 (s, 1H), 7.57 (s, 1H), 7.33 (d, J = 5.4 Hz, 1H), 6.96 (d, J = 5.4 Hz, 1H), 4.03 (dd, J = 7.3, 3.5 Hz, 2H), 1.30 (ddd, J = 18.4, 8.8, 3.8 Hz, 9H), and 0.94–0.81 (m, 6H).

3.4. Synthesis of Compound 2 (Br-DTP46-CHO)

To a 7 mL anhydrous CHCl3 solution, 250.00 mg of Compound 1 and 139.28 mg of NBS, each with a molecular weight of 0.7825 mmol, were added, making up 1 equivalent. Deionized water was used to quench the reaction after the reaction tube had been wrapped in tin foil and stirred at room temperature for 5.5 h while at a dark reaction. Using CH2Cl2 as an extractant, the raw material was isolated in saturated salt water. After subjecting the organic phase to steaming on a rotary evaporation apparatus, compound 2, a solid with a dark green color and a yield of 70.5% (219.8 mg), was isolated and purified using thin-layer chromatography with an eluent ratio of 1:1 between dichloromethane and petroleum ether. 1H NMR (400 MHz, CDCl3) δ: 9.85 (s, 1H), 7.58 (s, 1H), 7.03 (s, 1H), 4.02 (dd, J = 7.4, 4.8 Hz, 2H), 1.37–1.21 (m, 9H), and 0.93–0.83 (m, 6H).

3.5. Synthesis of Compounds 3 and 4

Compound 2 (300.00 mg, 0.7530 mmol), 2-(tributyltin) thiophene (281.03 mg, 0.7530 mmol, 1 equiv.), and Pd(PPh3)4 (43.51 mg, 0.0377 mmol, 0.05 equiv.) were added to a Schlenk reaction tube that was 25 mL in volume. The reaction tube was then vacuumed alternately at a low temperature of 0 °C, and filled with argon to remove any air. Following this, 5 mL of anhydrous toluene was added to the reaction tube, and oxygen in the reaction tube was removed three times through a pumping air-filled argon cycle. Finally, the mixture was stirred at a temperature of 100 °C under an atmosphere of argon. The final step in the reaction was to extract the crude product by evaporating the solvent toluene. The third compound, a yellow-green solid, was isolated by thin-layer chromatography from 227.6 mg of a mixture of dichloromethane and petroleum ether in a ratio of 1:1. The yield was 70.5%. Chemical compound 4, which has a yellow-green solid state and a yield of 63.0% (270.4 mg), was also produced using the same procedure.
Compound 3: 1H NMR (400 MHz, CDCl3) δ: 9.84 (s, 1H), 7.56 (s, 1H), 7.28 (d, J = 0.9 Hz, 1H), 7.25 (s, 1H), 7.06–7.01 (m, 2H), 4.05 (dd, J = 7.5, 4.2 Hz, 2H), 1.34–1.27 (m, 9H), and 0.93–0.87 (m, 6H).
Compound 4: 1H NMR (400 MHz, CDCl3) δ: 9.85 (s, 1H), 7.58 (s, 1H), 7.06 (d, J = 1.2 Hz, 1H), 7.01 (s, 1H), 6.84 (s, 1H), 4.07 (dd, J = 7.4, 4.8 Hz, 2H), 1.30 (dt, J = 12.3, 6.8 Hz, 28H), and 0.93–0.86 (m, 12H).

3.6. Synthesis of Compounds 5 and 6

During the reaction, compound 3 (200.00 mg, 0.4980 mmol) and NBS (88.63 mg, 0.4980 mmol, 1 equiv.) were added to 10 mL of anhydrous CHCl3 solution. The reaction tube was then wrapped in tin foil for 4.5 h at room temperature, and the reaction was finally quenched with deionized water. An extractant called CH2Cl2 was used to extract the crude product from saturated salt water. Next, the organic phase that had been purified was heated in a rotary evaporation apparatus. Then, compound 5, a solid yellow substance with a mass of 187.8 mg and a yield of 78.5%, was isolated using thin-layer chromatography with an eluent ratio of 1:1 between dichloromethane and petroleum ether. In the same way, compound 6, a solid yellow colorant, was produced (145.0 mg, yield 63.7%).
Compound 5: 1H NMR (400 MHz, CDCl3) δ: 9.83 (s, 1H), 7.55 (s, 1H), 7.06–6.90 (m, 3H), 4.11–3.96 (m, 2H), 1.30 (dd, J = 13.9, 6.4 Hz, 9H), and 0.88 (dt, J = 6.9, 4.3 Hz, 6H).
Compound 6: 1H NMR (400 MHz, CDCl3) δ: 9.85 (s, 1H), 7.58 (s, 1H), 7.01 (s, 1H), 6.84 (s, 1H), 4.08 (td, J = 7.5, 4.1 Hz, 2H), 1.35–1.23 (m, 26H), and 0.98–0.81 (m, 14H).

3.7. Synthesis of Compounds 7, 8, and 9

The following ingredients were added to a 25 mL Schlenk reaction tube: BDT-2SnMe3 (100.00 mg, 0.0964 mmol), compound 2 (84.47 mg, 0.2120 mmol, 2.2 equiv.), and Pd(PPh3)4 (5.57 mg, 0.0048 mmol, 0.05 equiv.). The tube was then vacuumed alternately at low temperature and filled with argon to remove air and 5 mL of anhydrous toluene was added. Oxygen was removed from the reaction tube by pumping the air-filled argon cycle three times. The mixture was then stirred at 100 °C under an argon atmosphere for 48 h. When the reaction was complete, the solvent toluene was evaporated in order to yield a crude product, which has a ratio of dichloromethane to petroleum ether that is equal to two to one. Thin-layer chromatography was utilized in order to purify the red solid chemical 7 (79.9 mg, yield 61.7%). The identical procedure was used to produce red solid compounds 8 and 9, with respective yields of 52.4% and 48.8%. The dosages of these compounds were 76.2 mg and 86.8 mg.
Compound 7: 1H NMR (400 MHz, CDCl3) δ: 9.69 (s, 2H), 7.45 (s, 2H), 7.34 (d, J = 14.9 Hz, 4H), 6.96 (s, 2H), 4.04 (dd, J = 16.4, 5.4 Hz, 4H), 3.04 (d, J = 6.1 Hz, 4H), 1.55 (dt, J = 19.8, 6.4 Hz, 10H), 1.32 (ddd, J = 22.9, 11.9, 7.9 Hz, 30H), and 1.00–0.90 (m, 20H).
Compound 8: 1H NMR (400 MHz, CDCl3) δ: 9.65 (s, 2H), 7.38–7.27 (m, 6H), 6.98 (t, J = 15.6 Hz, 4H), 6.78 (d, J = 25.2 Hz, 2H), 3.96 (s, 4H), 3.06 (dd, J = 15.6, 5.8 Hz, 4H), 1.62–1.51 (m, 10H), 1.34–1.24 (m, 21H), and 0.92 (ddd, J = 15.7, 10.0, 5.6 Hz, 20H).
Compound 9: 1H NMR (400 MHz, CDCl3) δ: 9.65 (s, 2H), 7.34 (s, 4H), 7.00 (d, J = 7.9 Hz, 4H), 6.81 (s, 2H), 3.96 (s, 4H), 3.04 (d, J = 6.1 Hz, 4H), 1.38–1.25 (m, 34H), and 1.00–0.89 (m, 26H).

3.8. Synthesis of S1, S2, and S3

In a round-bottom flask, compound 7 (50.00 mg, 0.0372 mmol), RDN-34 (42.25 mg, 0.2232 mmol, 6 equiv.), 2.5 mL of anhydrous toluene, 50 μL of acetic anhydride, and 50 μL of BF3-OEt2 were added. To determine whether or not the reaction of the raw materials is adequate, a capillary tube is used to perform the sampling point plate of the reaction liquid at any point in time throughout the reaction process. Following the evaporation of the solvent toluene at the conclusion of the reaction, which results in the formation of a crude product, dichloromethane/petroleum ether = 2:1 is used as the eluent, and the blue-violet solid substance S1 (40.1 mg, yield 63.8%) is purified using thin-layer chromatography. The same procedure was used to manufacture the blue-violet solid substance S2 (34.8 mg, yield 56.6%) and the blue-violet solid compound S3 (29.3 mg, yield 63.8%) with the same results.
S1: 1H NMR (400 MHz, CDCl3) δ: 7.44 (s, 2H), 7.29 (s, 2H), 7.10 (s, 2H), 6.77 (d, J = 39.1 Hz, 4H), 3.93 (s, 8H), 3.09 (d, J = 6.0 Hz, 4H), 1.65–1.51 (m, 14H), 1.47–1.25 (m, 32H), and 0.97 (tdd, J = 22.3, 14.8, 7.4 Hz, 32H).
S2: 1H NMR (400 MHz, CDCl3) δ: 7.78–7.28 (m, 4H), 7.14–6.44 (m, 10H), 3.80 (d, J = 35.4 Hz, 8H), and 1.66–0.87 (m, 78H).
S3: 1H NMR (400 MHz, CDCl3) δ: 7.84 (s, 2H), 7.55 (s, 2H), 7.35 (s, 2H), 7.16 (s, 2H), 7.00 (d, J = 21.2 Hz, 4H), 4.14–4.00 (m, 8H), 3.01 (d, J = 6.1 Hz, 4H), 1.57–1.49 (m, 8H), 1.32–1.22 (m, 40H), and 0.87 (ddd, J = 20.1, 10.0, 4.5 Hz, 30H).

3.9. Device Fabrication

The indium tin oxide (ITO)-coated glass was pre-cleaned in an ultrasonic bath with deionized water, acetone, and ethanol multiple times after the BHJ OPV devices based on the ITO/PEDOT:PSS/active layer/PDINO/Ag structure were prepared. The next step was a five-minute plasma treatment of the glass. On the ITO substrates, the PEDOT:PSS layer was spin-coated at a speed of 2500 revolutions per minute for forty seconds. After that, the layer was annealed at a temperature of 110 degrees Celsius for ten minutes. By dissolving Donor:Y6 (1.2:1, weight%) in CF at a concentration of 17.6 mg/mL, as well as adding 0.5% 1-CN and agitating the mixture for more than 30 min, an active layer solution was created. After the spin-coated active layer underwent thermal annealing in a nitrogen glove box, a 0.043 cm2 area was vacuum-deposited with 10 nm PDINO and 100 nm Ag. Under a monochromatic-lit 150 W xenon lamp with a Cornerstone 74004 lens, we measured the external quantum efficiency (EQE), and we analyzed the J-V curves in a glovebox with an AM 1.5 G solar spectrum (100 mW/cm2).

4. Conclusions

In this study, three different small donor molecule materials were synthesized using the Vilsmeier–Haack reaction, classical Stille coupling, and Knoevenagel condensation. These approaches were utilized to construct the final target small donor molecule materials (S1, S2, and S3), which were then analyzed using UV-vis, CV, J-V, and SCLC. The incorporation of a thiophene π-bridge into S2 from S1 enhances the blue shift of UV absorption, narrows the band gap, and lowers the HOMO energy level. As a consequence, the VOC of S2:Y6 devices increases to 0.75 V, and a 3% PCE is achieved. S3, which was created by adding an alkyl chain to a thiophene π-bridge, outperforms S2 in terms of solubility and π-π stacking, leading to enhanced exciton dissociation and charge collection capabilities following device preparation. Additionally, after blending with Y6, the highest JSC (10.59 mA/cm2), FF (49.72%), and PCE (4.25%) were achieved. According to recent studies, when designing molecules for molecular simulation, it is important to first check if the length of the alkyl chain can guarantee enough solubility. If it cannot, then the molecular accumulation during device preparation will be too strong, which will impact charge transport and make it impossible to harvest more effective current and voltage, ultimately hindering the device’s efficiency. The incorporation of an alkyl chain into the thiophene π-bridge and a thiophene π-bridge can enhance the device’s morphology and charge transfer behavior. This could be a promising strategy for the future design and synthesis of materials including tiny donor molecules.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29174231/s1. Figures S1–S12: 1H NMR spectra of all compounds.

Author Contributions

L.Y. conducted the main experimental part and wrote the draft of the manuscript. Y.W. tested BHJ devices; P.M. contributed to conceptualization and proofreading; P.L., Y.P. and Z.Q. carried out part of the experiments. S.L., Z.L. and C.Y. supervised the project. S.L. conceptualized the project, undertook funding acquisition, provided guidance during all stages, and contributed to the writing and editing of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The National Natural Science Foundation of China (No. 22169009), Jiangxi Provincial Natural Science Foundation (No. 20212ACB204007), Jiangxi Provincial Key Laboratory of Functional Molecular Materials Chemistry (20212BCD42018), and the Fundamental Research Funds for the Jiaxing University (Grant No. CD70519048) are appreciated for their financial support.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Material, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare that they have no conflicts of interest.

References

  1. Yuan, J.; Zhang, Y.; Zhou, L.; Zhang, G.; Yip, H.-L.; Lau, T.-K.; Lu, X.; Zhu, C.; Peng, H.; Johnson, P.A.; et al. Single-junction organic solar cell with over 15% efficiency using fused-ring acceptor with electron-deficient core. Joule 2019, 3, 1140–1151. [Google Scholar] [CrossRef]
  2. Cui, Y.; Yao, H.; Zhang, J.; Xian, K.; Zhang, T.; Hong, L.; Wang, Y.; Xu, Y.; Ma, K.; An, C.; et al. Single-junction organic photovoltaic cells with approaching 18% efficiency. Adv. Mater. 2020, 32, 1908205. [Google Scholar] [CrossRef]
  3. Zhu, C.; Yuan, J.; Cai, F.; Meng, L.; Zhang, H.; Chen, H.; Li, J.; Qiu, B.; Peng, H.; Chen, S.; et al. Tuning the electron-deficient core of a non-fullerene acceptor to achieve over 17% efficiency in a single-junction organic solar cell. Energy Environ. Sci. 2020, 13, 2459–2466. [Google Scholar] [CrossRef]
  4. Liu, H.; Li, Y.; Xu, S.; Zhou, Y.; Li, Z. Emerging chemistry in enhancing the chemical and photochemical stabilities of fused-ring electron acceptors in organic solar cells. Adv. Funct. Mater. 2021, 31, 2106735. [Google Scholar] [CrossRef]
  5. Chen, N.; Yang, L.-J.; Chen, Y.; Wu, Y.; Huang, X.-M.; Liu, H.; Xie, H.-Y.; Hu, L.; Li, Z.; Liu, S.-Y. PBDB-T accessed via direct C−H arylation polymerization for organic photovoltaic application. ACS Appl. Polym. Mater. 2022, 4, 7282–7289. [Google Scholar] [CrossRef]
  6. Murugan, P.; Hu, T.; Hu, X.; Chen, Y. Fused ring A–DA′D–A (Y-series) non-fullerene acceptors: Recent developments and design strategies for organic photovoltaics. J. Mater. Chem. A 2022, 10, 17968–17987. [Google Scholar] [CrossRef]
  7. Yang, L.-J.; Chen, N.; Huang, X.-M.; Wu, Y.; Liu, H.; Liu, P.; Hu, L.; Li, Z.-F.; Liu, S.-Y. Direct C−H arylation-derived donor polymers afford PCEs over 10% for organic solar cells. ACS Appl. Polym. Mater. 2023, 5, 7340–7349. [Google Scholar] [CrossRef]
  8. Fu, J.; Fong, P.W.K.; Liu, H.; Huang, C.-S.; Lu, X.; Lu, S.; Abdelsamie, M.; Kodalle, T.; Sutter-Fella, C.M.; Yang, Y.; et al. 19.31% binary organic solar cell and low non-radiative recombination enabled by non-monotonic intermediate state transition. Nat. Commun. 2023, 14, 1760. [Google Scholar] [CrossRef]
  9. Chen, T.; Li, S.; Li, Y.; Chen, Z.; Wu, H.; Lin, Y.; Gao, Y.; Wang, M.; Ding, G.; Min, J.; et al. Compromising charge generation and recombination of organic photovoltaics with mixed diluent strategy for certified 19.4% efficiency. Adv. Mater. 2023, 35, 2300400. [Google Scholar] [CrossRef]
  10. Wang, J.; Wang, Y.; Bi, P.; Chen, Z.; Qiao, J.; Li, J.; Wang, W.; Zheng, Z.; Zhang, S.; Hao, X.; et al. Binary organic solar cells with 19.2% efficiency enabled by solid additive. Adv. Mater. 2023, 35, 2301583. [Google Scholar] [CrossRef]
  11. Wu, Y.; He, X.-Y.; Huang, X.-M.; Yang, L.-J.; Liu, P.; Chen, N.; Li, C.-Z.; Liu, S.-Y. Synthesis of long-chain oligomeric donor and acceptors via direct arylation for organic solar cells. Chin. J. Chem. 2024, 42, 523–532. [Google Scholar] [CrossRef]
  12. Yang, L.-J.; Wu, Y.; Murugan, P.; Liu, P.; Qiu, Z.-Y.; Peng, Y.-L.; Li, Z.-F.; Liu, S.-Y. Advancing Integration of direct C−H arylation-derived star-shaped oligomers as second acceptors for ternary organic solar cells. ACS Appl. Mater. Interfaces 2024, 16, 26348–26359. [Google Scholar] [CrossRef] [PubMed]
  13. Hu, H.; Liu, S.; Xu, J.; Ma, R.; Peng, Z.; Peña, T.A.D.; Cui, Y.; Liang, W.; Zhou, X.; Luo, S.; et al. Over 19 % efficiency organic solar cells enabled by manipulating the intermolecular interactions through side chain fluorine functionalization. Angew. Chem. Int. Ed. 2024, 63, e202400086. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, J.; Zhu, J.; Li, C.; Lin, Y.; Li, X.; Zhang, H.; Ma, Z.; Lu, Y. Effect of nitro-substituted ending group on the photovoltaic and photocatalytic performance of non-fullerene acceptors. Chem. Eng. J. 2024, 490, 151467. [Google Scholar] [CrossRef]
  15. Yang, L.-J.; Chen, N.; Murugan, P.; Wu, Y.; Liu, P.; Huang, X.-M.; Li, Z.-F.; Liu, S.-Y. Direct C−H arylation vs Stille polymeri zation: Rational design, synthesis, and systematic examinations of π-conjugated polymers for organic solar cells. Mater. Today Chem. 2024, 37, 101987. [Google Scholar] [CrossRef]
  16. Hu, D.; Tang, H.; Chen, C.; Huang, P.; Shen, Z.; Li, H.; Liu, H.; Petoukhoff, C.E.; Jurado, J.P.; Luo, Y.; et al. Insights into preaggregation control of Y-series nonfullerene acceptors in liquid state for highly efficient binary organic solar cells. Adv. Mater. 2024, 36, 2402833. [Google Scholar] [CrossRef]
  17. Xu, W.; Du, L.; Du, Z.; He, W.; Li, H.; Li, G.; Yang, C.; Cheng, P.; Cao, Z.; Yu, D. Enhanced photovoltaic performance of asym metrical benzo dithiophene homopolymer donor materials in nonfullerene acceptor-based organic photovoltaics. Molecules 2024, 29, 1332. [Google Scholar] [CrossRef]
  18. Qian, D.; Ye, L.; Zhang, M.; Liang, Y.; Li, L.; Huang, Y.; Guo, X.; Zhang, S.; Tan, Z.; Hou, J.A.; et al. Design, application, and morphology study of a new photovoltaic polymer with strong aggregation in solution state. Macromolecules 2012, 45, 9611–9617. [Google Scholar] [CrossRef]
  19. Zhang, M.; Guo, X.; Ma, W.; Ade, H.; Hou, J. A large-bandgap conjugated polymer for versatile photovoltaic applications with high performance. Adv. Mater. 2015, 27, 4655–4660. [Google Scholar] [CrossRef]
  20. Li, Z.; Chueh, C.-C.; Jen, A.K.-Y. Recent advances in molecular design of functional conjugated polymers for high-performance polymer solar cells. Prog. Polym. Sci. 2019, 99, 101175. [Google Scholar] [CrossRef]
  21. Tan, Z.-R.; Xing, Y.-Q.; Cheng, J.-Z.; Zhang, G.; Shen, Z.-Q.; Zhang, Y.-J.; Liao, G.; Chen, L.; Liu, S.-Y. EDOT-based conjugated polymers accessed via C−H direct arylation for efficient photocatalytic hydrogen production. Chem. Sci. 2022, 13, 1725–1733. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, X.; Feng, L.; Zhang, K.; Liu, S.-Y. Carbazole and diketopyrrolopyrrole-based D-A π-conjugated oligomers accessed via direct C–H arylation for opto-electronic property and performance study. Molecules 2022, 27, 9031. [Google Scholar] [CrossRef]
  23. Tang, H.; Xu, T.; Yan, C.; Gao, J.; Yin, H.; Lv, J.; Singh, R.; Kumar, M.; Duan, T.; Kan, Z.; et al. Donor derivative incorpo ration: An effective strategy toward high performance all-small-molecule ternary organic solar cells. Adv. Sci. 2019, 6, 1901613. [Google Scholar] [CrossRef] [PubMed]
  24. Hu, D.; Yang, Q.; Chen, H.; Wobben, F.; Le Corre, V.M.; Singh, R.; Liu, T.; Ma, R.; Tang, H.; Koster, L.J.A.; et al. 15.34% efficiency all-small-molecule organic solar cells with an improved fill factor enabled by a fullerene additive. Energy Environ. Sci. 2020, 13, 2134–2141. [Google Scholar] [CrossRef]
  25. Han, D.; Wen, S.; Bi, F.; Shang, C.; Ding, X.; Saparbaev, A.; Zakhidov, E.; Kuvondikov, V.; Yang, C.; Sun, M. Small molecular donor materials based on electron withdrawing benzobisthiazole core unit enable an efficiency of 11.8% for organic solar cells. Chem. Eng. J. 2023, 463, 142400. [Google Scholar] [CrossRef]
  26. Yang, X.; Gao, Y.; Xu, L.-Y.; Wu, X.; Chen, X.; Shao, Y.; Xiao, B.; Liu, S.; Xia, J.; Sun, R.; et al. Efficient and stable all-small-molecule solar cells enabled by incorporating a designed giant molecule acceptor. Energy Environ. Sci. 2024, 17, 5962–5971. [Google Scholar] [CrossRef]
  27. Duan, T.; Gao, J.; Babics, M.; Kan, Z.; Zhong, C.; Singh, R.; Yu, D.; Lee, J.; Xiao, Z.; Lu, S. Difluorinated oligothiophenes for high-efficiency all-small-molecule organic solar cells: Positional isomeric effect of fluorine substitution on performance variations. Sol. RRL 2020, 4, 1900472. [Google Scholar] [CrossRef]
  28. Hu, D.; Yang, Q.; Zheng, Y.; Tang, H.; Chung, S.; Singh, R.; Lv, J.; Fu, J.; Kan, Z.; Qin, B.; et al. 15.3% efficiency all-small-molecule organic solar cells achieved by a locally asymmetric F, Cl disubstitution strategy. Adv. Sci. 2021, 8, 2004262. [Google Scholar] [CrossRef]
  29. Sun, Y.; Nian, L.; Kan, Y.; Ren, Y.; Chen, Z.; Zhu, L.; Zhang, M.; Yin, H.; Xu, H.; Li, J. Rational control of sequential morphology evolution and vertical distribution toward 17.18% efficiency all-small-molecule organic solar cells. Joule 2022, 6, 2835–2848. [Google Scholar] [CrossRef]
  30. Wu, S.; Feng, W.; Meng, L.; Zhang, Z.; Si, X.; Chen, Y.; Wan, X.; Li, C.; Yao, Z.; Chen, Y. 15.51% efficiency all-small-molecule organic solar cells achieved by symmetric thiazolyl substitution. Nano Energy 2022, 103, 107801. [Google Scholar] [CrossRef]
  31. Cai, S.; Huang, P.; Cai, G.; Lu, X.; Hu, D.; Hu, C.; Lu, S. Symmetrically fluorinated benzo[1,2-b:4,5-b′]dithiophene-cored donor for high-performance all-small-molecule organic solar cells with improved active layer morphology and crystallinity. ACS Appl. Mater. Interfaces 2022, 14, 14532–14540. [Google Scholar] [CrossRef]
  32. Ma, X.; Wang, C.; Deng, D.; Zhang, H.; Zhang, L.; Zhang, J.; Yang, Y.; Wei, Z. Small molecule donors design rules for non-halogen solvent fabricated organic solar cells. Small 2024, 20, 2309042. [Google Scholar] [CrossRef] [PubMed]
  33. Alam, S.; Lee, J. Progress and future potential of all-small-molecule organic solar cells based on the benzodithiophene donor material. Molecules 2023, 28, 3171. [Google Scholar] [CrossRef]
  34. Murugan, P.; Ravindran, E.; Sangeetha, V.; Liu, S.-Y.; Jung, J.W. Perylene-diimide for organic solar cells: Current scenario and prospects in molecular geometric, functionalization, and optoelectronic properties. J. Mater. Chem. A 2023, 11, 26393–26425. [Google Scholar] [CrossRef]
  35. Qiu, B.; Chen, Z.; Qin, S.; Yao, J.; Huang, W.; Meng, L.; Zhu, H.; Yang, Y.; Zhang, Z.-G.; Li, Y. Highly efficient all-small-molecule organic solar cells with appropriate active layer morphology by side chain engineering of donor molecules and thermal annealing. Adv. Mater. 2020, 32, 1908373. [Google Scholar] [CrossRef] [PubMed]
  36. Ge, J.; Xie, L.; Peng, R.; Fanady, B.; Huang, J.; Song, W.; Yan, T.; Zhang, W.; Ge, Z. 13.34% efficiency non-fullerene all-small-molecule organic solar cells enabled by modulating the crystallinity of donors via a fluorination strategy. Angew. Chem. Int. Ed. 2020, 59, 2808–2815. [Google Scholar] [CrossRef]
  37. Bin, H.; Angunawela, I.; Qiu, B.; Colberts, F.J.M.; Li, M.; Dyson, M.J.; Wienk, M.M.; Ade, H.; Li, Y.; Janssen, R.A.J. Precise control of phase separation enables 12% efficiency in all small molecule solar cells. Adv. Energy Mater. 2020, 10, 2001589. [Google Scholar] [CrossRef]
  38. Liu, C.; Qiu, N.; Sun, Y.; Ke, X.; Zhang, H.; Li, C.; Wan, X.; Chen, Y. All-small-molecule organic solar cells based on a fluorinated small molecule donor with high open-circuit voltage of 1.07 V. Front. Chem. 2020, 8, 329. [Google Scholar] [CrossRef] [PubMed]
  39. Guo, J.; Balakirev, D.O.; Gu, C.; Peregudova, S.M.; Ponomarenko, S.A.; Liu, Z.; Luponosov, Y.N.; Min, J.; Lei, A. End group tuning in small molecule donors for non-fullerene organic solar cells. Dyes Pigments 2020, 175, 108078. [Google Scholar] [CrossRef]
  40. Meng, W.; Lv, J.; Duan, T.; Kan, Z.; Lu, S.; Dai, X.; Li, Z. Small molecule donor based on alkoxylated benzothiadiazole unit: Synthesis and photovoltaics properties. Mater. Chem. Phys. 2020, 247, 122874. [Google Scholar] [CrossRef]
  41. Wu, Q.; Deng, D.; Zhou, R.; Zhang, J.; Zou, W.; Liu, L.; Wu, S.; Lu, K.; Wei, Z. Modulation of donor alkyl terminal chains with the shifting branching point leads to the optimized morphology and efficient all-small-molecule organic solar cells. ACS Appl. Mater. Interfaces 2020, 12, 25100–25107. [Google Scholar] [CrossRef]
  42. Dong, X.; Yang, K.; Tang, H.; Hu, D.; Chen, S.; Zhang, J.; Kan, Z.; Duan, T.; Hu, C.; Dai, X.; et al. Improving molecular planarity by changing alky chain position enables 12.3% efficiency all-small-molecule organic solar cells with enhanced carrier lifetime and reduced recombination. Sol. RRL 2020, 4, 1900326. [Google Scholar] [CrossRef]
  43. Xu, T.; Lv, J.; Yang, K.; He, Y.; Yang, Q.; Chen, H.; Chen, Q.; Liao, Z.; Kan, Z.; Duan, T.; et al. 15.8% efficiency binary all-small-molecule organic solar cells enabled by a selenophene substituted sematic liquid crystalline donor. Energy Environ. Sci. 2021, 14, 5366–5376. [Google Scholar] [CrossRef]
  44. Wang, Q.; Zhang, X.; Miao, Y.; Jiang, X.; Wang, X.; Zhang, Z.; Lv, Z.; Liu, T.; Zou, B.; Xu, H.; et al. Synergy effect of symmetry-breaking and end-group engineering enables 16.06% efficiency for all-small-molecule organic solar cells. ACS Mater. Lett. 2024, 6, 713–719. [Google Scholar] [CrossRef]
  45. Rasmussen, S.C.; Evenson, S.J.; McCausland, C.B. Fluorescent thiophene-based materials and their outlook for emissive appli cations. Chem. Commun. 2015, 51, 4528–4543. [Google Scholar] [CrossRef] [PubMed]
  46. Wolfe, R.M.; Culver, E.W.; Rasmussen, S.C. Synthesis and characterization of bis[1]benzothieno[3,2-b:2′,3′-d]pyrroles: Quantitative effects of benzannulation on dithieno[3,2-b:2′,3′-d]pyrroles. Molecules 2018, 23, 2279. [Google Scholar] [CrossRef]
  47. Rybakiewicz, R.; Skórka, Ł.; Gańczarczyk, R. Dithienopyrrole-based organic electroactive materials and their photovoltaic as pects. Curr. Org. Chem. 2020, 24, 2695–2736. [Google Scholar] [CrossRef]
  48. Bulumulla, C.; Gunawardhana, R.; Gamage, P.L.; Miller, J.T.; Kularatne, R.N.; Biewer, M.C.; Stefan, M.C. Pyrrole-containing semiconducting materials: Synthesis and applications in organic photovoltaics and organic field-effect transistors. ACS Appl. Mater. Interfaces 2020, 12, 32209–32232. [Google Scholar] [CrossRef]
  49. Vogt, A.; Schwer, F.; Förtsch, S.; Lorenz, C.; Mena-Osteritz, E.; Aubele, A.; Kraus, T.; Bäuerle, P. Broadly applicable synthesis of arylated dithieno[3,2-b:2′,3′-d]pyrroles as building blocks for organic electronic materials. Chem. Eur. J. 2021, 27, 12362–12370. [Google Scholar] [CrossRef]
  50. Mei, S.; Shao, W.; Huang, S.; Kong, X.; Hu, Z.; Yang, M.; Wu, W.; Tan, H. Novel D-A-π-A organic dyes with phenoxazine as a donor unit for dye-sensitized solar cells: The effect of an ethynyl group on performance. Energy Fuels 2021, 35, 19748–19755. [Google Scholar] [CrossRef]
  51. Zheng, T.; Huang, S.; Liu, Y.; Li, Z.; Kong, X.; Qin, N.; Tan, H. Molecular engineering strategies of spectral matching and structure optimization for efficient metal-free organic dyes in dye-sensitized solar cells: A theoretical study. J. Phys. Chem. A 2024, 128, 5861–5872. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, T.; Han, L.; Wei, H.; Zhu, D.; Bao, X.; Qiao, S.; Sun, W.; Chen, W.; Yang, R. Influence of a π-bridge dependent molecular configuration on the optical and electrical characteristics of organic solar cells. J. Mater. Chem. A 2016, 4, 8784–8792. [Google Scholar] [CrossRef]
  53. Piradi, V.; Zhang, G.; Li, T.; Zhang, M.; Peng, Q.; Zhan, X.; Zhu, X. Side-chain engineering of benzodithiophene-bridged dimeric porphyrin donors for all-small-molecule organic solar cells. ACS Appl. Mater. Interfaces 2020, 12, 41506–41514. [Google Scholar] [CrossRef]
  54. Wu, L.N.; Li, M.Y.; Sui, M.Y.; Huang, J.C.; Sun, G.Y.; Cheng, L. Achieve panchromatic absorption for all-small-molecule organic solar cells based on mono-porphyrin molecules by π-bridge modification. Mater. Today Energy 2021, 20, 100658. [Google Scholar] [CrossRef]
  55. Shen, H.; Ren, Y.; Li, J.; Xu, Y.; Han, C.; Zou, W.; Xu, H.; Sun, Y.; Kan, Y.; Gao, K. Enhanced performance via π-bridge alteration of porphyrin-based donors for all-small-molecule organic solar Cells. Chin. J. Chem. 2023, 41, 644–650. [Google Scholar] [CrossRef]
  56. Wang, P.; Li, Y.; Han, C.; Wang, J.; Bi, F.; Zheng, N.; Yang, J.; Wang, J.; Bao, X. Rationally regulating the π-bridge of small molecule acceptors for efficient organic solar cells. J. Mater. Chem. A 2022, 10, 17808–17816. [Google Scholar] [CrossRef]
  57. Wang, P.; Bi, F.; Li, Y.; Han, C.; Zheng, N.; Zhang, S.; Wang, J.; Wu, Y.; Bao, X. Manipulating the intermolecular interactions through side chain engineering and unilateral π-bridge strategy for efficient small molecular photovoltaic acceptor. Adv. Funct. Mater. 2022, 32, 2200166. [Google Scholar] [CrossRef]
  58. Wang, X.; Li, Y.; Li, J.; Zhang, Y.; Shao, J.; Li, Y. Direct arylation synthesis of small molecular acceptors for organic solar cells. Molecules 2023, 28, 3515. [Google Scholar] [CrossRef]
  59. Kang, X.; Ding, X.; Du, Z.; Bi, F.; Sun, M.; Bao, X. Understanding the impact of thiazole unit sequence in π-bridge on the perfor mance of small molecule donor materials. Synthetic Met. 2024, 307, 117657. [Google Scholar] [CrossRef]
  60. Chen, B.-W.; Cao, K.; Wang, X.; Chen, Z.-C.; Jeong, S.Y.; Qiu, Z.-L.; Dai, L.-S.; Li, Y.-F.; Yang, K.-Y.; Yun, D.-Q.; et al. Design and performance of small-molecule donors with donor–π-acceptor architecture toward vacuum-deposited organic photovoltaics having heretofore highest short-circuit current density. Small 2024, 2403486. [Google Scholar] [CrossRef]
  61. Oh, S.; Ryu, D.H.; Jin, S.-M.; Kim, Y.; Kim, J.Y.; Jahankhan, M.; Lee, S.; Song, C.E.; Lee, H.K.; Shin, W.S.; et al. Effect of π-bridge ring fusion on the wide-bandgap donor and its application for high-efficiency all small-molecule tandem solar cells. ACS Appl. Energy Mater. 2024, 7, 5005–5017. [Google Scholar] [CrossRef]
  62. Crociani, L. The double-cross of benzotriazole-based polymers as donors and acceptors in non-fullerene organic solar cells. Molecules 2024, 29, 3625. [Google Scholar] [CrossRef]
  63. Ou, Z.; Qin, J.; Jin, K.; Zhang, J.; Zhang, L.; Yi, C.; Jin, Z.; Song, Q.; Sun, K.; Yang, J.; et al. Engineering of the alkyl chain branching point on a lactone polymer donor yields 17.81% efficiency. J. Mater. Chem. A 2022, 10, 3314–3320. [Google Scholar] [CrossRef]
  64. Shen, S.; Yang, L.; Mi, Y.; Zhou, Y.; Li, M.; Zhang, J.; Ye, L.; Song, J. Alkyl branching sites on π-spacers for dipyran-based high-efficiency organic solar cells. ACS Appl. Energy Mater. 2023, 6, 1066–1075. [Google Scholar] [CrossRef]
  65. Yang, D.; Li, S.; Guo, J.; Zhuo, H.; Zhou, L.; Lai, W.; Zhang, Z.; Li, X.; Meng, L.; Li, Y. Enhanced photovoltaic performance of 9,10-difluorodithieno[3,2-a:2′,3′-c]phenazine-based polymer donors by the synergistic effect of alkyl-thiophene π-bridges and halogen atom modification. ACS Appl. Polym. Mater. 2023, 5, 10315–10323. [Google Scholar] [CrossRef]
  66. Li, J.; Li, K.; Mei, Z.; Chen, Y.; Xie, Q.; Yu, C.; Liu, X.; Wang, Y.; Wu, Y.; Liao, Q.; et al. Designing simple non-fused terthiophene-based electron acceptors for efficient organic solar cells. J. Energy Chem. 2024, 96, 501–508. [Google Scholar] [CrossRef]
  67. Cao, M.; Zhang, X.; Tang, X.; Meng, S.; Cao, J.; Zhang, Z.-G.; Weng, C.; Shen, P. Low-cost and completely non-fused small-molecule acceptors based on tetrathiophene featuring 3,5-dialkylthiophene side chain enable efficient organic solar cells. Chem. Eng. J. 2024, 484, 149584. [Google Scholar] [CrossRef]
  68. Ren, J.; Gao, S. Effect of an external electric field on the ordered structures of blended donor polymers in solar cells. J. Phys. Chem. C 2022, 126, 11318–11329. [Google Scholar] [CrossRef]
  69. Rodríguez-Martínez, X.; Hartnagel, P.; Riera-Galindo, S.; Beket, G.; Österberg, T.; Gao, F.; Kirchartz, T.; Inganäs, O. Air processing of thick and semitransparent laminated polymer: Non-fullerene acceptor blends introduce asymmetric current-voltage characteristics. Adv. Funct. Mater. 2023, 33, 2301192. [Google Scholar] [CrossRef]
Scheme 1. Structure of S1, S2, and S3.
Scheme 1. Structure of S1, S2, and S3.
Molecules 29 04231 sch001
Scheme 2. Synthesis of DTP46 precursor. (i) 1.2 eq DMF, 1.2 eq POCl3, 25.9 eq anhydrous DCE, 0 °C, 4 h; (ii) 1 eq NBS, anhydrous CHCl3, r.t.; (iii) 5% Pd(PPh3)4, anhydrous toluene, 110 °C, 48 h.
Scheme 2. Synthesis of DTP46 precursor. (i) 1.2 eq DMF, 1.2 eq POCl3, 25.9 eq anhydrous DCE, 0 °C, 4 h; (ii) 1 eq NBS, anhydrous CHCl3, r.t.; (iii) 5% Pd(PPh3)4, anhydrous toluene, 110 °C, 48 h.
Molecules 29 04231 sch002
Scheme 3. Synthesis of S1, S2, and S3. (iv) 5% Pd(PPh3)4, anhydrous toluene, 100 °C, 48 h; (v) anhydrous toluene, acetic anhydride, BF3-OEt2, r.t.
Scheme 3. Synthesis of S1, S2, and S3. (iv) 5% Pd(PPh3)4, anhydrous toluene, 100 °C, 48 h; (v) anhydrous toluene, acetic anhydride, BF3-OEt2, r.t.
Molecules 29 04231 sch003
Figure 1. UV-vis absorption spectra of S1, S2, and S3 (a) in CHCl3 solution and (b) in thin films; (c) CV curves; (d) FMO alignments.
Figure 1. UV-vis absorption spectra of S1, S2, and S3 (a) in CHCl3 solution and (b) in thin films; (c) CV curves; (d) FMO alignments.
Molecules 29 04231 g001
Figure 2. (a) Schematic device structure and chemical structures of S1, S2, S3, and Y6; (b) J-V curves; (c) corresponding EQE spectra of the OPVs.
Figure 2. (a) Schematic device structure and chemical structures of S1, S2, S3, and Y6; (b) J-V curves; (c) corresponding EQE spectra of the OPVs.
Molecules 29 04231 g002
Figure 3. (a) Semilogarithmic plot of J-V characteristics of devices in the dark; (b) Jph-Veff characteristics; (c) VOC and (d) JSC versus light intensity of the OSCs; (e) J-V curves of the hole-only device; (f) J-V curves of the electron-only device.
Figure 3. (a) Semilogarithmic plot of J-V characteristics of devices in the dark; (b) Jph-Veff characteristics; (c) VOC and (d) JSC versus light intensity of the OSCs; (e) J-V curves of the hole-only device; (f) J-V curves of the electron-only device.
Molecules 29 04231 g003
Table 1. Optical and electrochemical data of S1, S2, and S3.
Table 1. Optical and electrochemical data of S1, S2, and S3.
λmax s
(nm) a
λmax f
(nm) b
λonset
(nm) b
Eg
(eV) c
EHOMO
(eV) d
ELUMO
(eV) d
S15915807081.75−5.14−3.61
S25805857241.71−5.18−3.68
S35955867321.69−5.20−3.72
a Measured in chloroform solutions. b Cast from chloroform solutions. c Optical band gaps estimated from the onset wavelength (λonset): Eg = 1240/λedge. d Measured by CV. s In CHCl3 solution. f In thin films.
Table 2. Photovoltaic performance of the device by ITO/PEDOT:PSS/Donor:Y6/PDINO/Ag.
Table 2. Photovoltaic performance of the device by ITO/PEDOT:PSS/Donor:Y6/PDINO/Ag.
VOC
(V)
JSC
(mA/cm2)
JSC a
(mA/cm2)
FF
(%)
PCE b
(%)

S1:Y6
0.71
(0.70 ± 0.005)
8.08
(7.46 ± 0.48)
7.9342.16
(41.25 ± 0.89)
2.42
(2.27 ± 0.14)

S2:Y6
0.75
(0.75 ± 0.004)
9.23
(8.91 ± 0.33)
9.2042.93
(42.59 ± 0.46)
3.00
(2.85 ± 0.11)

S3:Y6
0.80
(0.79 ± 0.002)
10.64
(10.17 ± 0.51)
10.5949.72
(49.62 ± 0.39)
4.25
(4.21 ± 0.07)
The values in brackets are the average values and standard deviations. a Integral JSC from EQE curves. b Average PCEs based on 20 independent devices.
Table 3. Based on S1:Y6, S2:Y6, and S3:Y6 OSCs devices for electron mobility and hole migration.
Table 3. Based on S1:Y6, S2:Y6, and S3:Y6 OSCs devices for electron mobility and hole migration.
μh (cm2 V−1 s−1)μe (cm2 V−1 s−1)μh/μe
S1:Y66.19 × 10−45.70 × 10−41.09
S2:Y66.31 × 10−45.93 × 10−41.06
S3:Y67.22 × 10−46.98 × 10−41.03
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yang, L.; Wu, Y.; Murugan, P.; Liu, P.; Peng, Y.; Qiu, Z.; Li, Z.; Yu, C.; Liu, S. Impact of Different π-Bridges on the Photovoltaic Performance of A-D-D′-D-A Small Molecule-Based Donors. Molecules 2024, 29, 4231. https://doi.org/10.3390/molecules29174231

AMA Style

Yang L, Wu Y, Murugan P, Liu P, Peng Y, Qiu Z, Li Z, Yu C, Liu S. Impact of Different π-Bridges on the Photovoltaic Performance of A-D-D′-D-A Small Molecule-Based Donors. Molecules. 2024; 29(17):4231. https://doi.org/10.3390/molecules29174231

Chicago/Turabian Style

Yang, Lingjun, Yu Wu, Pachaiyappan Murugan, Peng Liu, Yulong Peng, Zhiyong Qiu, Zaifang Li, Changlin Yu, and Shiyong Liu. 2024. "Impact of Different π-Bridges on the Photovoltaic Performance of A-D-D′-D-A Small Molecule-Based Donors" Molecules 29, no. 17: 4231. https://doi.org/10.3390/molecules29174231

Article Metrics

Back to TopTop