Next Article in Journal
Effect of Additives on Stability of Alumina—Waste Alumina Suspension for Slip Casting: Optimization Using Box-Behnken Design
Next Article in Special Issue
Powder Pressed Cuprous Iodide (CuI) as A Hole Transporting Material for Perovskite Solar Cells
Previous Article in Journal
Thermal Characterization of Low-Dimensional Materials by Resistance Thermometers
Previous Article in Special Issue
n-i-p Nanocrystalline Hydrogenated Silicon Solar Cells with RF-Magnetron Sputtered Absorbers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Metal-Free Organic Chromophores Featuring an Ethynyl-Thienothiophene Linker with an n-Hexyl Chain for Translucent Dye-Sensitized Solar Cells

1
Department of Chemistry, Chung-Ang University, Seoul 06974, Korea
2
WestCHEM, School of Chemistry, University of Glasgow, Glasgow G12 8QQ, UK
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Materials 2019, 12(11), 1741; https://doi.org/10.3390/ma12111741
Submission received: 14 February 2019 / Revised: 8 May 2019 / Accepted: 21 May 2019 / Published: 29 May 2019
(This article belongs to the Special Issue Materials for Photovoltaic Applications)

Abstract

:
We report the simple synthesis of two organic chromophores featuring an ethynyl-thienothiophene linker with an n-hexyl chain (CSD-03 and CSD-04), their optical and electrochemical properties, and their use as photosensitizers in dye-sensitized solar cells (DSSCs). Our theoretical and experimental studies show that adding the second thienothiophene allows for narrowing the bandgap of the molecule and thus ensuring more light harvesting in the visible region. The efficiencies of both CSD-03 (5.46 ± 0.03%) and CSD-04 (5.20 ± 0.03%) are comparable to that of N719 (5.92 ± 0.01%) in translucent DSSCs fabricated with 5 μm-thick TiO2 photoanodes.

1. Introduction

The prosperity of a human society is mainly dependent on the supply of sustainable energy. Because of inevitable and imminent shortages of fossil fuels together with global warming, we should look for affordable and clean energy resources. Among renewable energy resources, solar energy is considered a promising way to meet future energy demand. Since the dye-sensitized solar cell (DSSC) was invented in 1991 [1], numerous efforts on materials and device architecture have been made to achieve high performance and long-term stability [2,3,4]. In DSSCs, the chromophore responsible for capturing solar light and producing photocurrent is a critical component to enhance their power conversion efficiency (PCE). For example, Ru-complexes have attained remarkable efficiencies over 11% due to their full absorption window from the visible to the near-infrared (NIR) region [5,6]. However, conventional Ru-complexes have significant drawbacks, such as scarcity, complicated synthesis, and toxicity [2]. Metal-free organic chromophores have been investigated as alternatives because of their many synthetic protocols, high molar extinction coefficients, and reduced toxicity issues [7,8,9]. Among these organic chromophores, a donor-π bridge-acceptor (D-π-A) architecture has been mostly adopted, and a structural modification can tune their photophysical properties to the architecture. Triarylamine-based moieties have been popularly selected as the donor unit because of their expedient synthesis and strong electron donating ability. Cyanoacrylic acid has been extensively used as the acceptor unit due to its direct synthesis from aldehydes and anchoring ability to TiO2. Methine or aromatic units have been introduced as π-spacers. However, the selection of an appropriate π-spacer is still of paramount importance in modulating the photophysical properties and enhancing the light-harvesting ability in the visible region. In particular, a planar π-spacer can increase electron delocalization and thus facilitate the electron transfer from the donor to acceptor for higher efficiency.
Thiophene and its derivatives have been widely used as π-bridges because of their low steric hindrance between adjacent molecules and a strong conjugation effect. Also, alkyne units have been used to maintain the planar structure of the D-π-A architecture while increasing the distance between D and A moieties. Recently, ethynyl-thiophene and ethynyl-thienothiophene spacers have been explored theoretically and experimentally. Wang’s group employed thienothiophene and bisthienothiophene linkers to achieve a high molar extinction coefficient of the metal-free organic sensitizers [10]. Recently, we also reported the facile synthesis of two organic dyes featuring ethynyl-thiophene from commercially available starting materials [11,12]. Zhao et al. introduced ethynyl-thiophene or 2-ethynyl-6-methylthieno[3,4-b]thiophene to zinc-porphyrin dyes to enhance the light-harvesting capability and prevent the intermolecular aggregation [13]. Fernandes et al. synthesized three push-pull organic dyes featuring 2,2’-bithiophene, ethynyl-2,2’-bithiophene or ethynyl-thieno[3,2-b]thiophene spacers [14]. In this study, we report two organic dyes, coded as CSD-03 and CSD-04 (Scheme 1), which can be synthesized in three steps from commercially available thienothiophene derivatives. The derivative has an n-hexyl side chain in order to improve the solubility of the final compounds. Importantly, the acetylene linker has been included to promote the planarity between the D and A units and thus increase electron delocalization. We also investigate the effect of the number of thienothiophene units on the optical, redox, and photovoltaic properties.

2. Results and Discussion

2.1. Synthesis

The two compounds, CSD-03 and CSD-04, were directly synthesized in a three-step protocol (Scheme 1). The Vilsmeier–Hack reaction of commercially available thienothiophene derivatives yielded compounds 1 and 2 as the aldehyde. The Sonogashira coupling reactions of compound 1 and 2 with 4-ethynyl-N,N-dimethylbenzenamine provided compounds 3 and 4, respectively. Subsequently, compounds 3 and 4 were condensed with cyanoacetic acid to obtain CSD-03 and CSD-04, respectively, under Knoevenagel condensation conditions. The two final dyes were fully characterized by 1H nuclear magnetic resonance (NMR) spectroscopy, 13C NMR spectroscopy, heteronuclear single quantum coherence (HSQC) spectroscopy, Fourier-transform infrared (FTIR) spectroscopy, and high-resolution mass spectrometry (HR-MS) in Figures S1 and S2.

2.2. Characterization

The UV-visible spectra of the two compounds in dimethylformamide (DMF) and dye-grafted mesoporous TiO2 films are shown in Figure 1a,b, respectively. In Figure 1a, two broad absorption bands of each chromophore were observed between 300 nm and 600 nm. The absorption band at lower wavelengths is attributed to localized aromatic π–π* transition, while that at higher wavelengths is ascribed to the intramolecular charge transfer (ICT) from donor to acceptor. The addition of another thienothiophene unit resulted in red-shift in the maximum absorption wavelength (λmax) from 419 nm to 455 nm, presumably a result of elongating π-conjugation. The molar extinction coefficients of CSD-03 and CSD-04 in DMF were 46,700 M−1 cm−1 and 50,400 M−1 cm−1, respectively. Their coefficients are higher than those of the ruthenium-complex dyes N3 (13,900 M−1 cm−1) [15] and N719 (14,000 M−1 cm−1) [16]. The optically determined bandgaps were 2.48 eV and 2.25 eV for CSD-03 and CSD-04, respectively. The incorporation of n-hexyl to thienothiophene made the absorption spectra red-shifted and broadened when compared to dye 6 in Reference [14] because of the weaker intermolecular π–π stacking [17]. Compared with the spectra in DMF, the ICT absorption band of the two dyes on TiO2 films exhibits significant spectral broadening because of J-aggregation of chromophore molecules and the interaction of the anchoring moieties with the TiO2 surface (Figure 1b).
The electrochemical behavior of the two dyes was investigated by square wave voltammetry (SWV), as shown in Figure 2. Both dyes exhibited a single oxidation wave and a single reduction wave. The addition of the second thienothiophene unit results in more positive reduction potential, ensuring a narrow fundamental energy gap, which correlates to the narrowing of the optical bandgap. According to the energy level of Fc/Fc+ redox couple (−4.8 eV under vacuum level [18]), the estimated ionization potentials (IPs) and electron affinities (EAs) are summarized in Table 1. The IPs are slightly lower than the energy level of the I/I3 redox couple (−4.92 eV), confirming that the regeneration of the two dyes is energetically favorable. The EAs are almost 1 eV above the TiO2 conduction band (-4.26 eV), indicating efficient electron transfer from the lowest unoccupied molecular orbital (LUMO) of the organic dye to the conduction band of nanocrystalline TiO2.

2.3. Theoretical Calculations

We performed computational calculations based on density functional theory (DFT) to recognize electron distribution in the frontier molecular orbitals. Figure 3a shows the ground-state optimized geometries of the CSD-03 and CSD-04 chromophores. They are fully conjugated throughout the D, π-spacer, and A moieties. The ethynyl-thienothiophene linker allows for maintaining a coplanar molecular structure, which can facilitate the ICT and suppress rotational disorder, even if the n-hexyl chain is on the periphery of the thienothiophene. Interestingly, the ground-state structure of CSD-04 has a twist angle of 13.4° between two thienothiophene moieties, and thus CSD-03 has a more planar geometry due to small torsion angles. In Figure 3b, the electrons are distributed from the donor to the π-bridge for both chromophores at the highest occupied molecular orbital (HOMO) level, while the excited electrons are localized on the cyanoacrylic acid at the LUMO level. It is beneficial to the efficient charge transfer from the dye to TiO2 upon irradiation. The addition of another thienothiophene unit drives up both HOMO and LUMO levels, in particular, due to a stronger influence on the HOMO. It should be noted that the larger difference between I3/I potential and the HOMO level allows more efficient dye regeneration because of preventing the geminate recombination between oxidized chromophore molecules and existing electrons in the TiO2 under solar illumination.

2.4. Dye-Sensitized Solar Cells

Figure 4a displays J-V characteristics of translucent DSSCs (5 μm-thick TiO2) sensitized with CSD-03 and CSD-04 without any co-adsorbent. The photovoltaic parameters are summarized in Table 2. Adding the second thienothiophene unit resulted in higher Jsc (from 12.55 mA/cm2 to 14.85 mA/cm2) due to stronger ICT absorption and higher extinction coefficient in the visible region but lower Voc (from 0.627 V to 0.587 V) due to the up-shift of the HOMO energy level. The photovoltaic properties depending on dye-soaking time were also measured and summarized in Table 2. Interestingly, when compared to N719, extended dye-soaking time of CSD-03 and CSD-04 resulted in lowering DSSC performance because of the faster aggregation of organic dyes on the TiO2 surface. It should be noted that the maximum efficiencies of both CSD-03 (5.46%) and CSD-04 (5.20%) are close to that of N719 (5.92%) in the translucent DSSCs. Figure 4b exhibits incident photon-to-current efficiency (IPCE) spectra of the semitransparent DSSCs. CSD-04 provides broader IPCE spectrum than CSD-03, and thus this agrees with the enhancement of Jsc. However, because of lower Voc and fill factor (FF), CSD-04 unexpectedly exhibited lower power-conversion-efficiency (η) than CSD-03. FF is mainly affected by many factors, such as the internal resistance in the DSSC relating on the charge transportation at the electrolyte/chromophore/TiO2 interface, the charge transfer process at Pt counter electrode and diffusion in the electrolyte and the conductance of transparent electrodes [19]. Also, the dye-containing n-hexyl side chain had better DSSC performance, and IPCE than its counterpart of the nonalkyl chain (dye 6 in Reference [14]) because of its enhanced solubility for dye loading and suppressed π–π stacking. Electrochemical impedance spectroscopy (EIS) can offer a deeper understanding of interfacial charge-transfer processes [20]. Figure 4c shows the Nyquist plots of the DSSCs in the dark. The EIS parameters fitted by an equivalent circuit are summarized in Table 3. The radius of the semicircle in the middle frequency region increases in the order of CSD-03 (84.46 Ω) < CSD-04 (111.20 Ω), yielding recombination resistance (Rrec). Large Rrec helps the charge recombination at the electrolyte/chromophore/TiO2 interface to be hampered. Unfortunately, CSD-04 had higher transport resistance (Rtr), shorter electron lifetime (τe), and lower charge collection efficiency (ηcc) than CSD-03 and consequently provided the lower power-conversion-efficiency (η) of the DSSC.

3. Materials and Methods

3.1. General

1H and 13C NMR spectra were recorded on a Bruker 500 MHz instrument (Billerica, MA, USA) with tetramethylsilane (TMS) as an internal standard. MS data were obtained using either a JEOL-700 MStation (EI, Akishima, Tokyo, Japan) or a Thermo Scientific LTQ Orbitrap XL (Waltham, MA, USA) with negative ion mode nanoelectrospray ionization (nESI). Infrared (IR) spectra were recorded on a Perkin-Elmer FTIR (Waltham, MA, USA). UV-visible spectra were obtained using a JASCO UV/Vis/NIR spectrophotometer (Tokyo, Japan). Optical bandgaps (Eg) were determined using the edge of the longest wavelength absorption (λ) using Eg = 1240/λ. A three-electrode system consisting of a Pt working electrode, a Pt wire counter electrode, and an Ag wire pseudoreference electrode was used for cyclic voltammetry measurements using a CHI 440A electrochemical analyzer (Austin, TX, USA). The redox potential of the dyes was measured in DMF containing 0.1 M tetrabutylammonium hexafluorophosphate, and the potential was calibrated against the ferrocene/ferrocenium (Fc/Fc+). Before recording the electrochemical data, the solutions were carefully purged with nitrogen gas.

3.2. Syntheses

Synthesis of 5-bromo-6-hexylthieno[3,2-b]thiophene-2-carbaldehyde (1). 2-bromo-3-hexylthieno [3,2-b]thiophene (1.35 g, 4.44 mmol) and phosphorus oxychloride (2.04 g, 13.4 mmol) were dissolved in DMF (16 mL). The solution was stirred for 30 min at 85 °C and then cooled to room temperature, quenched with H2O (20 mL). Dichloromethane (DCM) (200 mL) and H2O (200 mL) were added, and the organic layers were separated, dried over MgSO4 and filtered. The solvent was evaporated, and the residue was purified by silica gel column chromatography (Hex/EA = 20:1) to give 1 as a dark red oil (0.87 g, 59%). 1HNMR (400 MHz, CDCl3) δ 0.88 (t, J = 6.78 Hz, 3 H) 1.23–1.45 (m, 4 H) 1.48–1.59 (m, 2 H) 1.71–1.83 (m, 2 H) 3.32 (t, J = 7.53 Hz, 2 H) 8.04 (s, 1 H) 9.56 (s, 1 H) ppm; MS m/z(EI+) calculated for C13H1581BrOS2 331.97 found 332.00.
Synthesis of 5-bromo-6,6′-dihexyl-2,2′-bithieno[3,2-b]thiophene-5′-carbaldehyde (2). 5-bromo-6,6′- dihexyl-2,2′-bithieno[3,2-b]thiophene (4.24 g, 8.07 mmol) and phosphorus oxychloride (24.7 g, 161 mmol) were dissolved in DMF (40 mL). The solution was stirred for 20 h at 95 °C and then cooled to room temperature, quenched with H2O (20 mL). Dichloromethane (MC) (500 mL) and H2O (400 mL) were added. The organic layer was separated, dried over MgSO4 and filtered. The solvent was evaporated and the residue was purified by silica gel column chromatography (Hex/EA = 10:1) to give 2 as a yellow solid (0.8 g, 18%). 1H (400 MHz, CDCl3) δ 10.05, 7.32, 7.30, 3.06 (t), 2.75–2.67 (m), 1.79 (q), 1.74–1.66 (m), 1.41–1.29 (m), 0.93–0.86 (m) ppm; MS m/z(EI+) calculated for C25H2981BrOS4 554.03 found 554.10.
Synthesis of 5-((4-(dimethylamino)phenyl)ethynyl)-6-hexylthieno[3,2-b]thiophene-2-carbaldehyde (3). A stirred mixture of 4-ethynyl-N,N-dimethylbenzenamine(0.46 g, 3.2 mmol), 1 (1.00 g, 3.02 mmol), copper(I) iodide (0.012 g, 0.060 mmol), Triphenylphosphine (0.01 6g, 0.060 mmol) and bis(triphenylphosphine)palladium(II) dichloride (0.11 g, 0.039 mmol) in trimethylamine (30 mL) was refluxed for 3 h. After cooling the solution, MC (200 mL) and H2O (200 mL) were added. The organic layer was separated, dried over MgSO4 and filtered. The solvent was evaporated and the residue was purified by silica gel column chromatography (Hex/EA = 7:1) to give 3 as an orange solid (1.1 g, 92%). 1H NMR (400 MHz, CDCl3) δ 9.93(s,1H), 7.81(s,1H),6.67(d, J = 8.9 Hz, 2H), 6.61 (d, J = 8.9 Hz, 2H), 3.01 (s, 6H), 2.91 (t, 2H), 1.79 (p, J = 7.6 Hz, 2H), 1.46 – 1.26 (m, 7H), 0.89 (t, J = 7.1 Hz, 3H) ppm; MS m/z(EI+) calculated for C23H25NOS2 395.14 found 394.70.
Synthesis of 5′-((4-(dimethylamino)phenyl)ethynyl)-6,6′-dihexyl-[2,2′-bithieno[3,2-b]thiophene]-5-carbaldehyde (4). A stirred mixture of 4-ethynyl-N,N-dimethylbenzenamine (0.154 g, 1.06 mmol), 3 (0.56 g, 1.01 mmol), copper(I) iodide (7.7 mg, 0.041 mmol), triphenylphosphine (5.3 mg, 0.020 mmol), bis(triphenylphosphine)palladium(II) dichloride (35.5 mg, 0.0506 mmol) in triethylamine (30 mL) was refluxed for 6 h. After cooling the solution, MC (200 mL) and H2O (200 mL) were added. The organic layers were separated, dried over MgSO4 and filtered. The solvent was evaporated and the residue was purified by silica gel column chromatography (Hex/EA = 7:1) to give 4 as a yellow solid (0.31 g, 50%).1H NMR (400 MHz, DMSO-d6) δ 0.83 (t, J = 7.03 Hz, 3 H) 1.23–1.34 (m, 9 H) 2.90 (t, J = 7.28 Hz, 2 H) 2.98 (s, 6 H) 6.74 (m, J = 8.53 Hz, 2 H) 7.39 (m, J = 8.53 Hz, 2 H) 8.28 (s, 1 H) 8.57 (s, 1 H) ppm; MS m/z(EI+) calculated for C35H39NOS4 617.96 found 617.50.
Synthesis of (E)-2-cyano-3-(5-((4-(dimethylamino)phenyl)ethynyl)-6-hexylthieno[3,2-b]thiophen-2-yl) acrylic acid (CSD-03). A mixture of 2 (1.1 g, 2.8 mmol) and cyanoacetic acid (0.50 g, 5.6 mmol) was dissolved in CH3CN (50 mL) containing piperidine (1.42 g, 16.7 mmol). The solution was refluxed for 3 h. After cooling the solution, it was poured into HCl (2 M, 200 mL) and MC (500 mL). The organic layer was separated, washed with water, dried over MgSO4, and filtered. Then the solvent was removed in vacuo to afford CSD-03 as a dark red solid (0.3 g, 23%). 1H NMR (500 MHz, DMSO-d6) δ 8.47 (s, 1H), 8.20 (s, 1H), 7.39–7.37 (m, 2H), 6.74–6.72 (m, 2H), 2.97 (s, 6H), 2.89 (t, J = 7.4 Hz, 2H), 1.76–1.70 (m, 2H), 1.34 – 1.22 (m, 6H), 0.83 (t, J = 7.2 Hz, 3H) ppm.; 13C(126 MHz, DMSO-d6) δ 163.8, 151.0, 145.9, 138.7, 138.0, 137,2, 132.9, 131.8, 117.6, 112.3, 107.6, 101.7, 100.0, 81.0, 40.1, 31.2, 28.6, 28.5, 28.2, 22.4, 14.4 ppm; FT-IR (ATR) υ 2920, 2852, 2181, 1672, 1603, 1568, 1533, 1470, 1418, 1368, 13.12, 1244, 1192, 1148, 1125, 945, 832, 812, 758, 745, 608, 584 cm−1; HRMS m/z (FAB+) calculated for C26H26N2O2S2 [M]+ 462.1436, found 462.1439.
Synthesis of (E)-2-cyano-3-(5-((4-(dimethylamino)phenyl)ethynyl)-6,6′-dihexyl-2,2′-bithieno[3,2-b] thiophen-5′-yl)acrylic acid (CSD-04). A mixture of 4 and cyanoacetic acid (0.0173 g, 0.204 mmol) was dissolved in CH3CN (40 mL) containing piperidine (0.052 g, 0.061 mmol). The solution was refluxed for 5 h. After cooling the solution, it was poured into HCl (2 M, 200 mL) and MC (500 mL). The organic layer was separated and washed with water, dried over MgSO4, and filtered. Then, the solvent was evaporated in vacuo to afford CSD-04 as a red solid (0.056 g, 80%). 1H NMR (500 MHz, DMSO-d6) δ 8.22 (s, 1H), 7.79 (s, 1H), 7.76 (s, 1H), 7.45 (d, J = 9.0 Hz, 2H), 7.01 (s, 1H), 6.72 (d, J = 9.0 Hz, 2H), 2.96 (s, 6H), 2.89–2.86 (m, 4H), 1.69–1.63 (m, 12H), 0.86–0.83 (m, 6H) ppm.; 13C(151 MHz, DMSO-d6) δ 168.6, 150.4, 145.5, 141.8, 141.7, 140.6,139.5, 138.6, 138.3, 137.9, 136.9, 134.7, 133.0, 132.2, 125.6, 121.3, 119.2, 118.1, 117.0, 112.0, 104.6, 40.1, 31.0, 30.9, 29.1, 28.8, 28.4, 28.4, 27.9, 27.5, 22.0, 21.6, 14.0, 13.9, 13.9 ppm; FT-IR (ATR) υ 2911, 2854, 2202, 1763, 1603, 1521, 1458, 1427, 1363, 1312, 1283, 1204, 1182, 1105, 1057, 945, 810 cm−1; HRMS m/z (FAB+) calculated for C38H41N2O2S4 [M+H]+ 685.2051, found 685.0078.

3.3. Computational Details

Theoretical calculations were performed with the Gaussian 09 [21]. The ground-state geometries were fully optimized using the DFT method with Becke’s three-parameter hybrid exchange functionals and the Lee-Yang-Parr correlation functional (B3LYP) and 6-311G (d, p) basis set for all atoms without additional diffused function. Vibrational frequencies were also calculated to confirm that the optimized geometry would correspond to the lowest point of the respective potential energy surface.

3.4. DSSC Fabrication and Photovoltaic Measurements

The acetylene-black TiO2 paste [22] was screen-printed onto fluorine-doped SnO2 (FTO) conducting glass (TEC 8, Pilkington, 2.2-mm-thick, sheet resistance = 8 Ω/sq). The printed films were placed in a muffle furnace and slowly heated to 300 °C over a 30 min period, heated at 300 °C for 1 h, heated to 575 °C for 1 h, and then cooled to room temperature after 3 h. The final thickness of the sintered films was about 5 μm, and their active area was 0.2025 cm2 (i.e., 0.45 cm×0.45 cm). The prepared TiO2 photoanodes were dipped in a 0.04 M TiCl4 solution at 75 °C for 30 min, rinsed several times with deionized water and ethanol, and then annealed at 500 °C for 30 min on a hot plate. The photoanodes were treated with the O2 plasma for 10 min, and then immersed in 0.1 M HNO3 solution for 30 min to facilitate dye adsorption. The final ones were submerged in a 0.5 mM dye-containing ethanol solution (N719, CSD-03, and CSD-04) for different soaking time. After two holes were drilled in the FTO-coated glass, chloroplatinic acid (H2PtCl6) solution in 2-propanol was spin-coated and annealed at 525 ºC for 1 h in the muffle furnace to prepare the Pt counter electrode. Both the dye-grafted photoanode and the counter electrode were assembled with 25-μm-thick Surlyn (Solaronix, Aubonne, Switzerland). A liquid electrolyte (Iodolyte AN-50, Solaronix) was injected through the holes in the backside of the counter electrode.
The photovoltaic performance of the masked devices was investigated by using a solar cell I-V measurement system (K3000 LAB, McScience, Suwon, Korea) under the standard AM 1.5G spectrum (100 mW cm−2). Short-circuit photocurrent density (Jsc), open-circuit voltage (Voc), fill factor (FF), and power conversion efficiency (η) were obtained simultaneously. The rectangular mask size (e.g., 0.6 cm × 0.6 cm) was somewhat larger than the active area of the photoanodes because this is the preferred configuration to capture both direct and diffuse lights and to minimize internal reflection [23]. The incident photon-to-current conversion efficiency (IPCE) was obtained using a spectra IPCE measurement system (K3100, McScience, Korea). EIS experiments were performed with cells biased to Voc in the dark using a frequency response analyzer (Solartron 1260, AMETEK, Leicester, UK). A sinusoidal potential modulation of 10 mV was applied over a frequency range from 0.1 Hz to 100 kHz. The obtained EIS spectra were fitted using the equivalent circuit built in the ZView software (Version 3.5f, Scribner Associates Inc., Southern Pines, NC, USA).

4. Conclusions

In this article, we report the facile synthesis of the two organic dyes featuring an ethynyl–thienothiophene linker with an n-hexyl chain, which will add impetus to develop lower cost photosensitizers for the DSSCs. UV-vis spectroscopy and square wave voltammetry have shown that adding the second thienothiophene allowed for narrowing the bandgap of the molecule and ensuring better light harvesting in the DSSCs. Unexpectedly, the extended π-conjugation negatively affects the structural planarity and HOMO level of CSD-04. Therefore, the photovoltaic performance of CSD-03 (5.45 ± 0.03%) is better than that of CSD-04 (5.07 ± 0.18%).

Supplementary Materials

The following are available online at https://www.mdpi.com/1996-1944/12/11/1741/s1, Figure S1: (a) FTIR, (b) 1H NMR, (c) 13C NMR, and (d) HSQC spectra of CSD-03, Figure S2: (a) FTIR, (b) 1H NMR, (c) 13C NMR, and (d) HSQC spectra of CSD-04.

Author Contributions

J.H. conceived and designed the experiments; D.-S.L. synthesized the molecules; K.-W.P. conducted DFT/TDDFT calculations; A.A.W. performed electrochemical experiments and analyzed the data; D.-S.L. and K.-W.P. performed photovoltaic experiments. D.-S.L. and K.-W.P. prepared the manuscript; A.A.W. and J.H. revised the manuscript.

Funding

This work was supported by the National Research Foundation of Korea (NRF) Grant funded by the Ministry of Science and ICT for First-Mover Program for Accelerating Disruptive Technology Development (NRF-2018M3C1B9088457) and the Chung-Ang University Research Grants in 2015.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. O’Regan, B.; Grätzel, M. A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature 1991, 353, 737–740. [Google Scholar] [CrossRef]
  2. Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Petterson, H. Dye-sensitized solar cells. Chem. Rev. 2010, 110, 6595–6663. [Google Scholar] [CrossRef]
  3. Fakharuddin, A.; Brown, T.M.; Fabregat-Santiago, F.; Bisquert, J. A perspective on the production of dye-sensitized solar modules. Energ. Environ. Sci. 2014, 7, 3952–3981. [Google Scholar] [CrossRef]
  4. Freitag, M.; Teuscher, J.; Saygil, Y.; Zhang, X.; Giordano, F.; Liska, P.; Hua, J.; Zakeeruddin, S.M.; Moser, J.-E.; Grätzel, M.; et al. Dye-sensitized solar cells for efficient power generation under ambient lightning. Nat. Photonics 2017, 11, 372–378. [Google Scholar] [CrossRef]
  5. Nazeeruddin, M.K.; Angelis, F.D.; Fantacci, S.; Selloni, A.; Viscardi, G.; Liska, P.; Ito, S.; Takeru, B.; Grätzel, M. Combined experimental and DFT-TDDFT computational study of photoelectrochemical cell ruthenium sensitizers. J. Am. Chem. Soc. 2005, 127, 16835–16847. [Google Scholar] [CrossRef]
  6. Robertson, N. Optimizing dyes for dye-sensitized solar cells. Angew. Chem. Int. Ed. 2006, 45, 2338–2345. [Google Scholar] [CrossRef]
  7. Kanaparthi, R.K.; Kandhadi, J.; Giribabu, L. Metal-free organic dyes for dye-sensitized solar cells: Recent advances. Tetrahedron 2012, 68, 8383–8393. [Google Scholar] [CrossRef]
  8. Liang, M.; Chen, J. Arylamine organic dyes for dye-sensitized solar cells. Chem. Soc. Rev. 2013, 42, 3453–3488. [Google Scholar] [CrossRef] [PubMed]
  9. Mahmood, A. Triphenylamine based dyes for dye sensitized solar cells: A review. Sol. Energy 2016, 123, 127–144. [Google Scholar] [CrossRef]
  10. Zhang, G.; Bai, Y.; Shi, D.; Wegner, S.; Zakeeruddin, S.M.; Grätzel, M.; Wang, P. Employ a bisthienothiophene linker to construct an organic chromophore for efficient and stable dye-sensitized solar cells. Energy Environ. Sci. 2009, 2, 92–95. [Google Scholar] [CrossRef]
  11. Al-Eid, M.; Lim, S.H.; Park, K.-W.; Fitzpatrick, B.; Han, C.-H.; Kwak, K.; Hong, J.; Cooke, G. Facile synthesis of metal-free organic dyes featuring a thienylethynyl spacer for dye sensitized solar cells. Dyes Pigm. 2014, 104, 197–203. [Google Scholar] [CrossRef] [Green Version]
  12. Park, K.-W.; Ahn, S.; Baek, M.H.; Lim, D.-S.; Wiles, A.A.; Kim, M.G.; Hong, J. Coplanar D-π-A organic sensitizers featuring a thienylethynyl spacer for efficient dye-sensitized solar cells. Mater. Express 2017, 7, 43–50. [Google Scholar] [CrossRef]
  13. Zhao, H.; Long, J.; Zhao, B.; Tan, S. 2-ethynyl-6-methylthieno[3,2-b]thiophene as an efficient π spacer for porphyrin-based dyes. Dyes Pigm. 2015, 122, 168–176. [Google Scholar] [CrossRef]
  14. Fernandes, S.S.M.; Mesquita, I.; Andrade, L.; Mendes, A.; Justino, L.L.G.; Burrows, H.D.; Raposo, M.M.M. Synthesis and characterization of push-pull bithiophene and thieno[3,2-b]thiophene derivatives bearing an ethyne linker as sensitizers for dye-sensitized solar cells. Org. Electron. 2017, 49, 194–205. [Google Scholar] [CrossRef]
  15. Horiuchi, T.; Miura, H.; Uchida, S. Highly-efficient metal-free organic dyes for dye-sensitized solar cells. Chem. Commun. 2003, 9, 3036–3037. [Google Scholar] [CrossRef]
  16. Wang, P.; Klein, C.; Humphry-Baker, R.; Zakeeruddin, S.M.; Grätzel, M. A high molar extinction coefficient sensitizer for stable dye-sensitized solar cells. J. Am. Chem. Soc. 2005, 127, 808–809. [Google Scholar] [CrossRef]
  17. Won, Y.S.; Yang, Y.S.; Kim, J.H.; Ryu, J.-H.; Kim, K.K.; Park, S.S. Organic photosensitizers based on terthiophene with alkyl chain and double acceptors for application in dye-sensitized solar cells. Energy Fuels 2010, 24, 3676–3681. [Google Scholar] [CrossRef]
  18. Pommerehne, J.; Vestweber, H.; Guss, W.; Mahrt, R.F.; Bässler, H.; Porsch, M.; Daub, J. Efficient two layer leds on a polymer blend basis. Adv. Mater. 1995, 7, 551–554. [Google Scholar] [CrossRef]
  19. Han, L.; Koide, N.; Chiba, Y.; Islam, A.; Komiya, R.; Fuke, N.; Fukui, A.; Yamanaka, R. Improvement of efficiency of dye-sensitized solar cells by reduction of internal resistance. Appl. Phys. Lett. 2005, 86, 213501. [Google Scholar] [CrossRef]
  20. Hoshikawa, T.; Yamada, M.; Kikuchi, R.; Eguchi, K. Impedance analysis of internal resistance affecting the photoelectrochemical performance of dye-sensitized solar cells. J. Electrochem. Soc. 2015, 152, E68–E73. [Google Scholar] [CrossRef]
  21. Gaussian, R.A.; Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; et al. Gaussian 09, Revision A.02; Gaussian Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  22. Cho, T.-Y.; Han, C.-W.; Jun, Y.; Yoon, S.-G. Formation of artificial pores in nano-TiO2 photo-electrode films using acetylene-black for high-efficiency dye-sensitized solar cells. Sci. Rep. 2013, 3, 1496. [Google Scholar] [CrossRef]
  23. Ito, S.; Nazeeruddin, M.K.; Liska, P.; Comte, P.; Charvet, R.; Péchy, P.; Jirousek, M.; Kay, A.; Zakeeruddin, S.M.; Grätzel, M. Photovoltaic characterization of dye-sensitized solar cells: Effect of device masking on conversion efficiency. Prog. Photovolt: Res. Appl. 2006, 14, 589–601. [Google Scholar] [CrossRef]
Scheme 1. The synthetic route towards dyes CSD-03 and CSD-04.
Scheme 1. The synthetic route towards dyes CSD-03 and CSD-04.
Materials 12 01741 sch001
Figure 1. UV-visible absorption spectra of the two dyes (a) in DMF solution (1 × 10−5 mol L−1) and (b) adsorbed onto transparent TiO2 film.
Figure 1. UV-visible absorption spectra of the two dyes (a) in DMF solution (1 × 10−5 mol L−1) and (b) adsorbed onto transparent TiO2 film.
Materials 12 01741 g001
Figure 2. Square wave voltammetry of (a) CSD-03 and (b) CSD-04.
Figure 2. Square wave voltammetry of (a) CSD-03 and (b) CSD-04.
Materials 12 01741 g002
Figure 3. (a) Optimized geometries and (b) calculated electron distribution and energy levels of CSD-03 and CSD-04.
Figure 3. (a) Optimized geometries and (b) calculated electron distribution and energy levels of CSD-03 and CSD-04.
Materials 12 01741 g003
Figure 4. (a) Current density-voltage (J–V) characteristics, (b) incident photon-to-current efficiency (IPCE) spectra, and (c) Nyquist plots in the dark of the masked dye-sensitized solar cells (DSSCs) fabricated with CSD-03 and CSD-04.
Figure 4. (a) Current density-voltage (J–V) characteristics, (b) incident photon-to-current efficiency (IPCE) spectra, and (c) Nyquist plots in the dark of the masked dye-sensitized solar cells (DSSCs) fabricated with CSD-03 and CSD-04.
Materials 12 01741 g004
Table 1. Summary of optical and electrochemical properties of the as-synthesized dyes.
Table 1. Summary of optical and electrochemical properties of the as-synthesized dyes.
Dyesλmax/λonset
(nm)
Eg (optical)
(eV)
Eox
(V)
Ered
(V)
IP
(eV)
EA
(eV)
Eg (e-chem)
(eV)
CSD-03419/4992.480.52−1.35−5.32−3.451.87
CSD-04455/5502.250.53−1.32−5.34−3.481.85
Table 2. Masked DSSC (5 μm-thick TiO2) performance parameters of CSD-03, CSD-04, and N719.
Table 2. Masked DSSC (5 μm-thick TiO2) performance parameters of CSD-03, CSD-04, and N719.
Dyes (Soaking Time)Voc (V)Jsc (mA/cm2)FFη (%)
CSD-03 (3 h)0.627 ± 0.01312.55 ± 0.1869.4 ± 1.15.46 ± 0.03
CSD-03 (6 h)0.619 ± 0.02212.64 ± 0.6869.1 ± 0.25.41 ± 0.08
CSD-03 (12 h)0.6111 ± 0.00411.15 ± 0.5168.7 ± 0.34.69 ± 0.27
CSD-04 (3 h)0.587 ± 0.00414.85 ± 0.2958.2 ± 1.45.07 ± 0.18
CSD-04 (6 h)0.576 ± 0.00913.86 ± 0.3565.2 ± 1.05.20 ± 0.03
CSD-04 (12 h)0.587 ± 0.01612.974 ± 0.4264.1 ± 0.94.88 ± 0.01
N719 (3 h)0.608 ± 0.00314.74 ± 0.4762.1 ± 0.15.57 ± 0.21
N719 (24 h)0.671 ± 0.00114.95 ± 0.3959.1 ± 0.65.92 ± 0.01
Table 3. Electrochemical impedance spectroscopy (EIS) fitting parameters of the DSSCs (5 μm-thick TiO2).
Table 3. Electrochemical impedance spectroscopy (EIS) fitting parameters of the DSSCs (5 μm-thick TiO2).
DyesRtr (Ω)Rrec (Ω)Cμ (mF)τe (ms)ηcc (%)
CSD-03 (3 h)5.6884.460.086.3493.70
CSD-04 (3 h)11.99111.200.055.6590.27

Share and Cite

MDPI and ACS Style

Lim, D.-S.; Park, K.-W.; Wiles, A.A.; Hong, J. Metal-Free Organic Chromophores Featuring an Ethynyl-Thienothiophene Linker with an n-Hexyl Chain for Translucent Dye-Sensitized Solar Cells. Materials 2019, 12, 1741. https://doi.org/10.3390/ma12111741

AMA Style

Lim D-S, Park K-W, Wiles AA, Hong J. Metal-Free Organic Chromophores Featuring an Ethynyl-Thienothiophene Linker with an n-Hexyl Chain for Translucent Dye-Sensitized Solar Cells. Materials. 2019; 12(11):1741. https://doi.org/10.3390/ma12111741

Chicago/Turabian Style

Lim, Dong-Suk, Kwang-Won Park, Alan A. Wiles, and Jongin Hong. 2019. "Metal-Free Organic Chromophores Featuring an Ethynyl-Thienothiophene Linker with an n-Hexyl Chain for Translucent Dye-Sensitized Solar Cells" Materials 12, no. 11: 1741. https://doi.org/10.3390/ma12111741

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop