Next Article in Journal
Energy-Trapping Characteristics of Lateral Field Excited GdCOB Crystal Bulk Acoustic Wave Devices Based on Stepped Electrodes
Next Article in Special Issue
Compressive Behavior of Some Balls Manufactured by 3D Printing from Ceramic–Polymer Composite Materials
Previous Article in Journal
High Performance Rotating Triboelectric Nanogenerator with Coaxial Rolling Charge Pump Strategy
Previous Article in Special Issue
Electric Field Effects on Curved Graphene Quantum Dots
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Insights into Improving the Photovoltaic Performance of Dye-Sensitized Solar Cells by Removing Platinum from the Counter Electrode Using a Graphene-MoS2 Composite or Hybrid

1
Department of Organic Colourants, Institute for Colour Science and Technology, Tehran P.O. Box 16765-654, Iran
2
Department of Nanomaterials and Nanocoatings, Institute for Colour Science and Technology, Tehran P.O. Box 16765-654, Iran
3
Department of Physics, Shahrood University of Technology, Shahrood P.O. Box 36155-316, Iran
4
Faculty of Mechanical Engineering and Design, Kaunas University of Technology, Studentu Street 56, LT 51373 Kaunas, Lithuania
5
Department of Chemistry, Faculty of Engineering and Technology, SRM Institute of Science and Technology, Kattankulathur 603203, Tamil Nadu, India
*
Author to whom correspondence should be addressed.
Micromachines 2023, 14(12), 2161; https://doi.org/10.3390/mi14122161
Submission received: 10 October 2023 / Revised: 7 November 2023 / Accepted: 24 November 2023 / Published: 26 November 2023

Abstract

:
Photovoltaic systems, such as dye-sensitized solar cells (DSSCs), are one of the useful tools for generating renewable and green energy. To develop this technology, obstacles such as cost and the use of expensive compounds must be overcome. Here, we employed a new MoS2/graphene hybrid or composite instead of platinum in the DSSCs. Furthermore, the correctness of the preparation of the MoS2/graphene hybrid or composite was evaluated by field emission scanning electron microscope (FESEM), and the results showed that the desired compound was synthesized correctly. Inexpensive organic dyes were used to prepare the DSSCs, and their chemical structure was investigated by density functional theory (DFT) and cyclic voltammetry (CV). Finally, the DSSCs were fabricated using MoS2/graphene composite or hybrid, and to compare the results, the DSSCs were also prepared using platinum. Under the same conditions, the DSSCs with MoS2/graphene composite illustrated better efficiency than MoS2/graphene hybrid or/and graphene.

1. Introduction

Population growth leads to an increase in energy demand in the areas of household consumption, industry, and technology. Therefore, it is necessary to develop new sources of energy [1,2]. Solar radiation is one of the energy sources that has the mentioned characteristics. Many studies have been conducted to harness this energy and effectively convert it into other energies, such as electrical energy [3,4]. One of the emerging energy generation technologies that uses sunlight is dye-sensitized solar cells (DSSCs) [5,6]. DSSCs are a very promising alternative to classical inorganic p–n junction solar cells as they combine molecular systems and nanoparticles to create a device that mimics photosynthesis, with the objective of turning sunlight into a renewable, reliable, and low-cost source of energy closer to existence [7]. The components of a DSSC are: photoanode, photosensitizers (dyes), electrolyte, and counter electrode. Two important challenges to the widespread development of this technology are the low efficiency and the high price of some of its components [8].
An important component that affects the cost of this technology is the counter electrode. Platinum is used to prepare the counter electrode, which is not only expensive but also an end-to-end resource [9]. Baskaran et al. considered the possibility of replacing platinum on the counter electrode with Cu2ZnSnS4/MoS4. The solar cell constructed with platinum had an efficiency of about 4.4%, while the efficiency with CZMO8 was 4.07%. The proximity of the results indicates the confirmation of the replacement of platinum with the prepared nanocomposite [10]. Sarkar et al. investigated the performance of DSSCs based on a new nanohybrid of CoNi2S4 graphene oxide as a counter electrode. The results showed that the DSSCs fabricated using RGO and CoNi2S4 have an efficiency of 3.44% and 5.78%, respectively. However, the use of the composite increases the efficiency up to 9.22%, which is in good agreement with the efficiency of platinum (9.51%) [11]. Research shows that different derivatives of graphene oxide can be a suitable option to replace platinum in the counter electrode.
In this paper, our aim was to investigate the replacement of platinum counter electrodes with new graphene-based compounds. For this purpose, a composite and a hybrid of graphene oxide (GO), MoS2, MoS2/GO nanohybrid, and MoS2/GO nano-composite were fabricated and used in a DSSC. To investigate the effect of the prepared compounds, solar cells containing both a Pt-free and a Pt counter electrode were fabricated. Organic dyes (Figure 1) and N719 were used as photosensitizers in the prepared cells. The process of synthesis of the organic dyes used as photosensitizers in this study has already been reported in ref. [12].

2. Experimental

2.1. Materials

The J-V curves were studied using the Bunko-Keiki CEP-2000 system and flower-like MoS2 was synthesized through the hydrothermal method following our previous work [13]. However, in the hybrid particles, the GO sheets were placed in a suspension of ammonium heptamolybdate and thiourea before the hydrothermal process. In this case, the MoS2 sheets formed on the graphene sheets.

2.2. X-ray Diffraction

The phase series was confirmed by X-ray diffraction (XRD) and a Philips XRD diffractometer using C u k α radiation at 40 KV, 30 mA, a step size of 0.05° (2ϴ), and a scan rate of 1°/min. X’Pert software was used for qualitative analysis and the report of the diffraction peak width (rad, β) at full width half maximum (FWHM) for different 2θ values according to the location of the peaks (version 4.9.0).

2.3. FESEM Analysis

FESEM is a magnification technique that allows magnification between ×10 and ×300,000, making the information about the elements present and their topographic data on the surface practically visible through an unlimited field depth. In this study, the MIRA3TESCAN-XMU FESEM was inserted into the high vacuum portion of the microscope through an exchange chamber and anchored to a moving stage. In the virtual FESEM, the object can be moved in the horizontal and vertical directions on the screen by pressing the arrows in the POSITION box. The detector types offered are the standard lens detector SE, the high-efficiency Everhart–Thornley detector (ETD) SE, and the angle-selective BSE detector. The in-lens detector SE is used to detect SE signals directly from the sample surface.

2.4. Electrochemical and Density Functional Theory (DFT) and Cyclic Voltammetry Study

The cyclic voltammetry (CV) technique was used to study the electrochemical properties of the synthesized dyes. The three electrodes were used to study the oxidation potential (Eox) in solution media. The counter and a Pt wire were chosen as counter and reference electrodes, respectively. The calibration routine in this technique was performed with the redox pair Fc and Fc+ [14,15]. Furthermore, the Gaussian software was utilized for density functional theory (DFT) calculations [16].
CV measurements of the organic dyes were carried out in acetonitrile. The oxidation potential (Eox) was measured using three small electrodes. A Ag quasi-reference electrode (QRE) was used as the reference. Platinum wires were used as the working and the counter electrodes. All electrode potentials were calibrated with respect to the ferrocene (Fc)/ferrocenium (Fc+) redox couplet. An acetonitrile solution (2 mL) of dyes containing tetrabutylammonium perchlorate (0.1 mol dm−3) and ferrocene (ca. 1 mmol dm−3) was prepared. The electrochemical measurements were performed at a scan rate of 100 mV s−1 [17].

2.5. Solar Cell Assembly

Nanocrystalline TiO2 paste was coated onto an FTO coated glass substrate. The dye was absorbed by dipping the coated glass for 18 h in an ethanolic solution of the organic dyes. Finally, the film was washed with an acetonitrile-ethanol 1:1 mixed solvent. Acenonitrile-ethylenecarbonate (v/v = 1:4) containing tetrabutyl ammonium iodide (0.5 mol dm−3) was used as electrolyte. Spray coating technique was employed to form GO, MoS2, GO/MoS2 nanohybrid, and GO/MoS2 nanocomposite thin films as the counter electrode. The dye-adsorbed photoelectrode, counter electrode, and the electrolyte solution were assembled into a sealed sandwich type solar cell. An action spectrum was measured under monochromatic light with a constant photon number (5 × 1015 photon cm−2 s−1) [18]. J-V characteristics were measured under illumination with AM 1.5 simulated sunlight (100 mW cm−2) through a shading mask (5.0 mm × 4 mm) by a Bunko-Keiki CEP-2000 system.

3. Results and Discussion

In the Modified-Scherrer method [19,20], when Ln β is plotted versus Ln ( 1 cos θ ) and the least squares method is performed, the intercept gives Ln K λ L , from which a single value for L can be obtained. K λ C u k α 1 L = e ( intercept ) ; in the equation, K is the shape factor (K = 0.89), λ C u k α = 0.15405 nm, L is the crystal size, and the intercept is related to the linear equation. According to the XRD spectra (Figure 2), no amorphous phases were observed and the sharp crystallite peaks belonged to 2θ ~ 12.4 ° , except for dyes 7 and 8, which were due to substitutions of the R1, R2, and phenothiazine derivative. Furthermore, the crystallite size values ranged from 99 to 101 nm (Figure 3).
The morphology of graphene was studied using FESEM. All graphene shows wrinkled and flexible nanosheets with a thickness below 20 nm. As shown in Figure 4, graphene exhibits a multilayered structure. Graphene flakes have a relatively large surface area and interconnected three-dimensional graphene sheets [21]. Figure 5 shows an overview of MoS2/GO composites and the hybrid morphology prepared by simple mixing and the hydrothermal method using surfactants, respectively. As shown in Figure 5, no isolated MoS2 particles or flakes can be seen in the MoS2/GO composites. In contrast, in the hybrid sample, the MoS2 layers are well distributed on the graphene surface [22]. Figure 6 shows microflowers of aggregated MoS2 nanosheets in monolayer form. It shows spherical particles composed of many layers of MoS2 nanosheets below 20 nm [23].
In contrast to thick graphitic material, atomically thin 2D materials such as grapheme have attracted much research interest as a possible platform for atomic diffusion barriers and ionic tunneling layers [24]. Recently, there have been rapid theoretical and experimental developments of elemental 2D materials in electrocatalytic reactions, including the hydrogen evolution reaction (HER), oxygen evolution reaction (OER), oxygen reduction reaction (ORR), carbon dioxide reduction (CO2 RR), and N2 reduction reaction (NRR). The most recent development in nanomaterials, especially 2D materials, provides enormous opportunities for high-performance solar cells beyond conventional bulk materials. Notably, some elemental 2D materials with high carrier mobility and tunable band gaps are considered to be promising candidates for solar cells. Many theoretical studies have investigated the potential of elemental 2D materials in solar cells [25]. Nikam et al. synthesized MoS2 nanosheets on three-dimensional (3D) conductive MoO2 via a two-step chemical vapor deposition (CVD) reaction. The 3D MoO2 structure can create structural disorders in MoS2 nanosheets (3D MoS2/MoO2). The MoS2 nanosheets could protect the inner MoO2 core from the acidic electrolyte in the HER. The high activity of the as-synthesized 3D MoS2/MoO2 hybrid material in HER is attributed to the small onset overpotential of 142 mV and a large cathodic current density of 85 mA cm–2 [26]. In this article, a series of novel compounds based on graphene have been used to prepare the counter electrode in the DSSCs.
The CV method is used to measure the energy level of organic dyes on a laboratory scale. The information obtained by this method can be used to predict the behavior of dyes in the solar device [27,28]. The oxidation potential of peal in acetonitrile for dyes 1–8 was estimated to be +0.83, +0.85, +0.85, +0.82, +0.86, +0.85, +0.83, and +0.86 V vs. a normal hydrogen electrode (NHE), respectively. The LUMO level of the photosensitizers was obtained at −1.29, −1.33, −1.32, −1.30, −1.33, −1.34, −1.32, and −1.36 V versus NHE, respectively. The results confirm our ability to use dyes in DSSCs as photosensitizers (Figure 7). Silva et al. investigated the CV behavior of photosensitizers based on the tetrazole group. The results showed that the presence of strong donor electron groups concentrates the electron density of HOMO in addition to the effects of good adsorption on the semiconductor when di methyl formamide (DMF) solvent was used for this test [29]. The compounds used in this study dissolved well in acetonitrile, so this solvent was used for the CV test. However, the results showed that the compounds used were capable of transferring electrons to the semiconductor layer and accepting electrons from the electrolyte. Subsequently, the laboratory results were compared with the theoretical studies.
The catalytic activity of the electrode materials was examined by cyclic voltammetry (CV) measurements. The results showed that two pairs of redox peaks can be observed, with the negative pair belonging to redox reaction (1) and the positive pair belonging to redox reaction (2):
I 3 + 2 e 3 I
3 I 2 + 2 e 2 I 3
The shape of the measured CV curves is similar to those observed in Pt CE, indicating a similar electrochemical behavior to Pt CEs. For new materials, the current density of the CV curve increased significantly, indicating higher catalytic activity. The improvement of the electrocatalytic activity can be partially ascribed to the overall surface area that leads to more diffusion pathways for the electrolyte species [30,31].
All calculations of the derivatives were performed using the Gaussian 09 program package [32]. Quantum chemical calculations were performed and complete optimization of all structures was done using a DFT method coupled to the TD-DFT/Becke-three-Lee-Yang-Parr (B3LYP) hybrid/6-31g(d,p) level of theory [33]. No symmetry constraints were used in the optimizations. To verify the nature of the stationary geometries, a harmonic vibration frequency calculation was carried out to ensure that no imaginary frequencies were located. Taking into account DFT calculations, the distributions of the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO), as well as the orbitals of the transitions, were performed when the carbazole and/or phenothiazine derivatives were coupled to the phenyl ring. Figure 8 illustrates the HOMO, LUMO, and S0 of the organic photosensitizers. The amounts of HOMO ranged from −5.22 to −6.11 eV for the dyes in tandem. Dye 4 showed the highest value of the HOMO distribution (6.11 eV), and, in comparison, it is due to the chemical structure that the simultaneous substitution of 2-cyanoacrylic acid (C4H3NO2) and iodide (I) in the carbazole derivative increases the HOMO value. Moreover, HOMO shows that the introduction of groups at these sites has a remarkable effect on the orbital properties of the molecule, affecting both the energy band gap and the photovoltaic properties. In addition, LUMO values of −1.95, −2.56, −2.06, −2.67, −2.09, −2.68, −2.18, and −2.78 eV were calculated for dyes 1–8, respectively. The special substitution was performed in the LUMO region, and the simultaneous use of C4H3NO2 and I on the phenothiazine derivative showed that the LUMO shifted to a higher energy level of 2.78 eV. To study and predict the interactions with the solid surface, conceptual density function theory (CDFT), i.e., electronic chemical potential μ (Equation (3)), chemical hardness η (Equation (4)), and global electrophilicity ω (Equation (5)), has been widely used to study and explain the reactivity of organic dyes [33,34].
μ = E HOMO + E LUMO 2
η = E LUMO E LUMO
ω = μ 2 2 η
The calculated values of μ, η, and ω are listed in Table 1. According to the results in Table 1, the μ of dye 5 (−3.65 eV) is higher than that of the other dyes. Therefore, throughout the adsorption process, the electron transfer ratio is higher for dye 5 and has the ability to evade the orientation of the electrons. Consequently, dye 5 behaves as a nucleophile, while dye 4 reacts as an electrophile, as is confirmed by the higher HOMO value of dye 4. In terms of the ω index, dye 8 exhibits high electrophilicity values and therefore has a high reactivity of nucleophiles [35]. According to the η values, dye 8 is the softest among the dyes, while dyes 1 and 3 are the hardest. Therefore, dye 8 has a low excitation energy and the electron density is easy to change, while the large excitation energy or electron density of dyes 1 and 3 is very difficult to change [36].
Furthermore, the EH-L values were recorded as 3.85, 3.49, 3.85, 3.44, 3.13, 2.99, 3.21, and 3.01 eV for dyes 1–8, respectively. It is obvious that C4H3NO2 and I as substituents simultaneously on the phenothiazine derivative decrease the EH-L value. TD-DFT results were carried out with optimized S0 geometries and considering the solvation effects of dichloromethane using the C-PCM model [33,37]. TD-DFT uses different functions for the interpretation of absorption molecule spectra [38,39]. Moreover, DFT functions were used to measure the absorption behavior and compare it with experimental data. Taking the transitions into account, the theoretical UV-Vis spectrum was determined at the relativistic level using Gaussian software (Figure 9). The maximum wavelength values extracted from the calculations of DFT agreed well with the experimental maximum wavelengths of the dyes. It is clear that dyes 6 and 8 showed visible spectra in the range from 400 to 600 nm, because the LUMO values of these dyes were higher due to the lowest optical bandgap energy (evaluated from the edges of the theoretical UV-Vis of the dyes, E g opt = 1240 λ onset ). In an overall comparison of these groups, it was noticed that the HOMO values were higher when the carbazole derivative was chosen as the HOMO part rather than the phenothiazine derivative. It was interesting to note that using the phenothiazine derivative as the HOMO part of the absorption spectrum showed a better visible range as dyes 5, 6, 7, and 8 showed maximum wavelength values of 457, 466, 440, and 466 nm in the visible range.
We overlay with the CIE photopic spectral luminous efficiency function, which represents the average spectral sensitivity of human visual perception to light. The average visible light transmission (AVT) through the photoanode was calculated based on the following equation (Equation (6)) [40]:
AVT ( % ) = 390 880 T ( λ ) × V ( λ ) × S ( λ ) d λ 390 880 V ( λ ) × S ( λ ) d λ
where T(λ) is the transmission of the sample, V(λ) is the photopic spectral luminous efficiency function, and S(λ) is the AM 1.5 g solar spectrum [34]. AVTs (%) of dyes 1–8 are 83.7, 82.5, 82.6, 88.1, 81.6, 85.6, 88.6, and 87.9%, respectively. Thus, all dyes are very promising candidates as photosensitizers for highly transparent DSSCs. This trend in AVT values could be in contrast with the fact all dyes have a good molar attenuation coefficient (ε) in solution [40,41].
In this study, titanium dioxide was used as the semiconductor in the photoanode. Titanium dioxide is one of the best metal oxides for use in DSSCs, and its optimization and restructuring have been the subject of numerous studies [42]. However, one of the most important and influential components of the cost of a DSSC is the counter electrode. In this research, new compounds were used as the counter electrode and their performance was investigated in comparison with platinum. The organic dyes used as photosensitizers in DSSCs are based on indoline (dyes 1–8). Platinum was used as a counter electrode in the fabrication of solar cells with these compounds. The efficiencies obtained for dyes 1 to 8 and N719 in the structure of the platinum DSSCs are: 2.98, 3.69, 4.31, 5.87, 2.11, 2.95, 3.71, 3.98, and 9.48%. To remove the expensive platinum, graphene oxide (GO), MoS2, GO-MoS2 nanohybrid, and GO-MoS2 nanocomposite were investigated. The photovoltaic performance of each DSSC device with TiO2 is summarized in Table 2, Table 3, Table 4 and Table 5. Furthermore, the results show that the performance of the photosensitizers containing cyanoarylic acid is better than that of the compounds lacking this substitution. The level of photocurrent depends on the chemical structure of the dye and the ability to bind effectively with the semiconductor [43,44]. Consequently, a decrease in photocurrent indicates poor bonding between the dye and the nanosheet [45].
Graphene oxide has poor electrocatalytic properties that reduce the efficiency and FF values of DSSCs. For example, the efficiency of a dye 6 in a DSSC with a platinum counter electrode is about 5.87%, but this value is reduced to 64% by changing the counter electrode to graphene oxide, which is about 3.93%. Suriani et al. used graphene oxide and its hybrid with platinum to fabricate a solar cell. They used the spray coating technique to fabricate a thin film. Moreover, the efficiency of the DSSCs with graphene oxide as the counter electrode is about 23% of that of DSSCs with platinum as the counter electrode. It is important to note that the use of a hybrid of graphene oxide and platinum improves the efficiency of the DSSCs by 5% compared to the platinum-individual DSSCs [46,47]. In another study, Tamilselvi et al. investigated the effects of using graphene oxide and its hybrid with NiSe2. The results showed that the performance of the DSSCs was improved by about 57% in the presence of the graphene oxide-NiSe2 hybrid [48]. Due to the significant reduction of efficiency in the presence of graphene oxide as a counter electrode of DSSCs found in this and other studies, the use of a hybrid or composite of graphene oxide seems necessary. One of the proposed components to optimize the performance of graphene oxide in DSSCs is MoS2. Francis et al. fabricated a new configuration of MoS2 using a hydrothermal process. The morphological structure of the desired composition was confirmed by XRD and FESEM methods. Further, a spin coating method was used to fabricate the counter electrode of MoS2. The efficiency of the DSSCs fabricated with this counter electrode was about 4.50% [49]. Silambarasan et al. prepared a composite using a combination of graphene oxide and MoS2 as a counter electrode for DSSCs. The efficiencies of DSSCs containing platinum, MoS2, and the fabricated composite were 5.17, 2.01, and 4.65%, respectively. As can be seen, the efficiency of composite-based DSSCs is very close to that of platinum [50]. The use of MoS2 to fabricate a suitable hybrid or composite of graphene oxide is a new approach to reducing the cost of DSSC technology. In this study, a hybrid or composite of graphene oxide and MoS2 was prepared as a counter electrode for use in DSSCs. The results showed that the efficiency of DSSCs using hybrids and composites is similar to platinum up to 95%. In other words, these compounds can replace platinum.
Although PCE and AVT are often inversely related, as discussed, all photovoltaic technologies have potential applications. Progressive development warrants new compound figures of merit beyond PCE and AVT that allow for improvements to be referenced across various technologies. Light utilization efficiency (LUE) metric was calculated by Equation (7) [41].
LUE = PCE × AVT
This metric enables comparison between technologies against theoretical limits and also represents an overall system efficiency: for example, in a window or display application, this would be the combination of generated power efficiency and overall lighting efficiency (that is, the transmitted light per incident lighting power). A device with a high LUE could be applied over a window to provide power without blocking natural light from entering the room to reduce the need for artificial lighting during daylight hours [40]. The LUE value obtained for DSSCs was in the range of 41–44%.

4. Conclusions

In this research, two types of hybrids and composites were prepared from the combination of graphene oxide and MoS2 for use in the counter electrode of DSSCs by the hydrothermal method. All graphene shows wrinkled and flexible nanosheets with thickness below 20 nm. No isolated MoS2 particles or flakes can be seen in the MoS2/GO composites. In contrast, in the hybrid sample, the MoS2 layers are well distributed on the graphene surface. The LUMO level of the photosensitizers was obtained at −1.29, −1.33, −1.32, −1.30, −1.33, −1.34, −1.32, and −1.36 V versus NHE, respectively. However, the results showed that the compounds used were capable of transferring electrons to the semiconductor layer and accepting electrons from the electrolyte. The results show that the conductivity characteristics of hybrid and composite GO/MoS2 prepared from their components (GO and MoS2) are higher and can be a suitable option to replace platinum in DSSCs. Graphene oxide has poor electrocatalytic properties that reduce the efficiency and FF values of DSSCs. Eight dyes based on indoline and N719 as a standard dye were used as sensitizers in the DSSCs containing crystallite size values ranging from 99 to 101 nm. The AVTs (%) of dyes 1–8 are 83.7, 82.5, 82.6, 88.1, 81.6, 85.6, 88.6, and 87.9%, respectively. Thus, all dyes are very promising candidates as photosensitizers for highly transparent DSSCs. Finally, the DSSCs were prepared and evaluated using organic dye, a titanium dioxide photoelectrode, and five different counter electrodes (GO, MoS2, GO-MoS2 nano-hybrid, and GO-MoS2 nano-composite, Pt). The results showed that the efficiency of DSSCs using hybrids and composites is similar to platinum up to 95%. In other words, these compounds can replace platinum. The LUE value obtained for DSSCs was in the range of 41–44%.

Author Contributions

M.H.: Conceptualization, Methodology, Investigation, Writing—original draft. M.G.: Data curation and calculations. G.M.: synthesis and analysis. M.F.: FESEM and fabrication devices. A.P. and G.J.: Data curation. V.N.: XRD analysis. S.N.: Analysis, DFT calculations, and Writing and editing of the draft. All authors have read and agreed to the published version of the manuscript.

Funding

This project received funding from European Social Fund (project No 09.3.3-LMT-K-712-23-0156) under the grant agreement with the Research Council of Lithuania (LMTLT).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could influence the work reported in this paper.

References

  1. Omar, A.; Ali, M.S.; Abd Rahim, N. Electron transport properties analysis of titanium dioxide dye-sensitized solar cells (TiO2-DSSCs) based natural dyes using electrochemical impedance spectroscopy concept: A review. Sol. Energy 2020, 207, 1088–1121. [Google Scholar] [CrossRef]
  2. Duarte, D.A.; Massi, M.; Da Silva Sobrinho, A.S. Development of dye-sensitized solar cells with sputtered N-Doped TiO2 thin films: From modeling the growth mechanism of the films to fabrication of the solar cells. Int. J. Photoenergy 2014, 2014, 839757. [Google Scholar] [CrossRef]
  3. Korshunov, V.M.; Chmovzh, T.N.; Chkhetiani, G.R.; Taydakov, I.V.; Rakitin, O.A. New D–A–D luminophores of the [1,2,5]thiadiazolo[3,4-d]pyridazine series. Mendeleev Commun. 2022, 32, 371–373. [Google Scholar] [CrossRef]
  4. Badawi, A.; Althobaiti, M.G.; Alharthi, S.S.; Al-Baradi, A.M. Tailoring the optical properties of CdO nanostructures via barium doping for optical windows applications. Phys. Lett. A 2021, 411, 127553. [Google Scholar] [CrossRef]
  5. Ghomashi, R.; Rabiei, M.; Ghomashi, S.; Reza Massah, A.; Kolahdoozan, M.; Hosseinnezhad, M.; Ebrahimi-Kahrizsangi, R.; Palevicius, A.; Nasiri, S.; Janusas, G. Synthesis and investigation of the theoretical and experimental optical properties of some novel azo pyrazole sulfonamide hybrids. Mater. Lett. 2022, 317, 132132. [Google Scholar] [CrossRef]
  6. Ishikawa, K.; Garskaite, E.; Kareiva, A. Sol-gel synthesis of calcium phosphate-based biomaterials-A review of environmentally benign, simple, and effective synthesis routes. J. Sol-Gel Sci. Technol. 2020, 94, 551–572. [Google Scholar] [CrossRef]
  7. Muñoz-García, A.B.; Benesperi, I.; Boschloo, G.; Concepcion, J.J.; Delcamp, J.H.; Gibson, E.A.; Meyer, G.J.; Pavone, M.; Pettersson, H.; Hagfeldt, A.; et al. Dye-sensitized solar cells strike back. Chem. Soc. Rev. 2021, 50, 12450–12550. [Google Scholar] [CrossRef] [PubMed]
  8. Hosseinnezhad, M.; Gharanjig, K.; Ghahari, M.; Nasiri, S. Investigation of Photo-electrode and Counter Electrode Effect on DSSCs Based on Indoline Dyes. Prog. Color. Color. Coat. 2023, in press. [Google Scholar]
  9. Samson Adedoja, O.; Sadiku, E.R.; Hamam, Y. Prospects of Hybrid Conjugated Polymers Loaded Graphene in Electrochemical Energy Storage Applications. J. Inorg. Organomet. Polym. Mater. 2023, 1, 3. [Google Scholar]
  10. Baskaran, P.; Nisha, K.D.; Harish, S.; Ikeda, H.; Archana, J.; Navaneethan, M. Enhanced catalytic performance of Cu2ZnSnS4/MoS2 nanocomposites based counter electrode for Pt-free dye-sensitized solar cells. J. Alloys Compd. 2022, 894, 162166. [Google Scholar] [CrossRef]
  11. Sarkar, A.; Bera, S.; Chakraborty, A.K. CoNi2S4-reduced graphene oxide nanohybrid: An excellent counter electrode for Pt-free DSSC. Sol. Energy 2020, 208, 139–149. [Google Scholar] [CrossRef]
  12. Hosseinnezhad, M.; Moradian, S.; Gharanjig, K.; Afshar Taromi, F. Synthesis and characterisation of eight organic dyes for dye sensitised solar cells. Mater. Technol. 2014, 29, 112–117. [Google Scholar] [CrossRef]
  13. Mobarhan, G.; Zolriasatein, A.; Ghahari, M.; Jalili, M.; Rostami, M. The enhancement of wear properties of compressor oil using MoS2 nano-additives. Adv. Powder Technol. 2022, 33, 103648. [Google Scholar] [CrossRef]
  14. Devadiga, D.; Selvakumar, M.; Shetty, P.; Santosh, M.S. Recent progress in dye sensitized solar cell materials and photo-supercapacitors: A review. J. Power Sources 2021, 493, 229698. [Google Scholar] [CrossRef]
  15. Rouhani, S.; Hosseinnezhad, M.; Sohrab, N.; Gharanjig, K.; Salem, A.; Ranjbar, Z. Investigation of the Effect of rGO/TiO2 on Photovoltaic Performance of DSSCs Devices. Prog. Color. Color. Coat. 2022, 15, 123–131. [Google Scholar]
  16. Yang, J.; Yu, X.; Li, Y.; Cheng, G.; Yi, Z.; Zhang, Z.; Chi, F.; Liu, L. A Novel Dye-Sensitized Solar Cell Structure Based on Metal Photoanode without FTO/ITO. Micromachines 2022, 13, 122. [Google Scholar] [CrossRef]
  17. Hosseinnezhad, M.; Ghahari, M.; Mobarhan, G.; Rouhani, S.; Fathi, M. Towards low cost and green photovoltaic devices: Using natural photosensitizers and MoS2/Graphene oxide composite counter electrodes. Opt. Mater. 2023, 139, 113775. [Google Scholar] [CrossRef]
  18. Wu, Z.S.; Song, X.C.; Liu, Y.D.; Zhang, J.; Wang, H.S.; Chen, Z.J.; Liu, S.; Weng, Q.; An, Z.W.; Guo, W.J. New organic dyes with varied arylamine donors as effective co-sensitizers for ruthenium complex N719 in dye sensitized solar cells. J. Power Sources 2020, 451, 227776. [Google Scholar] [CrossRef]
  19. Rabiei, M.; Palevicius, A.; Monshi, A.; Nasiri, S.; Vilkauskas, A.; Janusas, G. Comparing Methods for Calculating Nano Crystal Size of Natural Hydroxyapatite Using X-ray Diffraction. Nanomaterials 2020, 10, 1627. [Google Scholar] [CrossRef]
  20. Nasiri, S.; Rabiei, M.; Shaki, H.; Hosseinnezhad, M.; Kalyani, K.; Palevicius, A.; Vilkauskas, A.; Janusas, G.; Nutalapati, V.; Kment, S.; et al. What is TADF (thermally activated delayed fluorescence) compared to the mechanisms of FL (fluorescence), PH (phosphorescence), and TTA (triplet–triplet annihilation) based on a novel naphthalimide sulfonylphenyl derivative as a host? J. Photochem. Photobiol. A Chem. 2024, 447, 115289. [Google Scholar] [CrossRef]
  21. Liu, C.; Bai, Y.; Zhao, Y.; Yao, H.; Pang, H. MoS2/graphene composites: Fabrication and electrochemical energy storage. Energy Storage Mater. 2020, 33, 470–502. [Google Scholar] [CrossRef]
  22. Wang, Z.; Ma, L.; Chen, W.; Huang, G.; Chen, D.; Wang, L.; Lee, J.Y. Facile synthesis of MoS2/graphene composites: Effects of different cationic surfactants on microstructures and electrochemical properties of reversible lithium storage. RSC Adv. 2013, 3, 21675–21684. [Google Scholar] [CrossRef]
  23. Barzegar, M.; Berahman, M.; Zad, A.I. Sensing behavior of flower-shaped MoS2 nanoflakes: Case study with methanol and xylene. Beilstein J. Nanotechnol. 2018, 9, 608–615. [Google Scholar] [CrossRef]
  24. Nikam, R.D.; Kwak, M.; Lee, J.; Rajput, K.G.; Hwang, H. Controlled Ionic Tunneling in Lithium Nanoionic Synaptic Transistor through Atomically Thin Graphene Layer for Neuromorphic Computing. Adv. Electron. Mater. 2020, 6, 1901100. [Google Scholar] [CrossRef]
  25. Wu, Z.; Qi, J.; Wang, W.; Zeng, Z.; He, Q. Emerging elemental two-dimensional materials for energy applications. J. Mater. Chem. A 2021, 9, 18793–18817. [Google Scholar] [CrossRef]
  26. Nikam, R.D.; Lu, A.Y.; Sonawane, P.A.; Kumar, U.R.; Yadav, K.; Li, L.J.; Chen, Y.T. Three-Dimensional Heterostructures of MoS2 Nanosheets on Conducting MoO2 as an Efficient Electrocatalyst to Enhance Hydrogen Evolution Reaction. ACS Appl. Mater. Interfaces 2015, 7, 23328–23335. [Google Scholar] [CrossRef]
  27. Huang, R.Y.; Tsai, W.H.; Wen, J.J.; Chang, Y.J.; Chow, T.J. Spiro[fluorene-9,9′-phenanthren]-10′-one as auxiliary acceptor of D-A-π-A dyes for dye-sensitized solar cells under one sun and indoor light. J. Power Sources 2020, 458, 228063. [Google Scholar] [CrossRef]
  28. Movahedi, J.; Haratizadeh, H.; Falah, N.; Hosseinnezhad, M. Investigation of effect of thiophene-2-acetic acid as an electron anchoring group for a photovoltaic device. Opto-Electron. Rev. 2019, 27, 334–338. [Google Scholar] [CrossRef]
  29. da Silva, L.; Sánchez, M.; Freeman, H.S. New tetrazole based dyes as efficient co-sensitizers for dsscs: Structure-properties relationship. Org. Electron. 2020, 87, 105964. [Google Scholar] [CrossRef]
  30. Bu, I.Y.-y. Hydrothermal production of low-cost bismuth sulfide/reduced graphene oxide nanocomposite as counter electrode for DSSCs. Optik 2020, 217, 164868. [Google Scholar] [CrossRef]
  31. Suriani, A.B.; Muqoyyanah Mohamed, A.; Mamat, M.H.; Hashim, N.; Isa, I.M.; Malek, M.F.; Kairi, M.I.; Mohamed, A.R.; Ahmad, M.K. Improving the photovoltaic performance of DSSCs using a combination of mixed-phase TiO2 nanostructure photoanode and agglomerated free reduced graphene oxide counter electrode assisted with hyperbranched surfactant. Optik 2018, 158, 522–534. [Google Scholar] [CrossRef]
  32. Gaussian 09 Citation. Available online: https://gaussian.com/ (accessed on 1 November 2016).
  33. Nasiri, S.; Dashti, A.; Hosseinnezhad, M.; Rabiei, M.; Palevicius, A.; Doustmohammadi, A.; Janusas, G. Mechanochromic and thermally activated delayed fluorescence dyes obtained from D–A–D′ type, consisted of xanthen and carbazole derivatives as an emitter layer in organic light emitting diodes. Chem. Eng. J. 2022, 430, 131877. [Google Scholar] [CrossRef]
  34. Lefebvre, C.; Khartabil, H.; Boisson, J.C.; Contreras-García, J.; Piquemal, J.P.; Hénon, E. The Independent Gradient Model: A New Approach for Probing Strong and Weak Interactions in Molecules from Wave Function Calculations. ChemPhysChem 2018, 19, 724–735. [Google Scholar] [CrossRef] [PubMed]
  35. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef]
  36. El Azouzi, L.; El Hadki, H.; El Hadki, A.; El Alouani, M.; Mabrouki, J.; Tazi, R.; Komiha, N.; Zrineh, A.; Kabbaj, O.K. A computational investigation on the adsorption of amoxycillin on graphene oxide nanosheet. Int. J. Environ. Anal. Chem. 2022. [Google Scholar] [CrossRef]
  37. Pearson, R.G. Chemical hardness and density functional theory. J. Chem. Sci. 2005, 117, 369–377. [Google Scholar] [CrossRef]
  38. Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. Energies, structures, and electronic properties of molecules in solution with the C-PCM solvation model. J. Comput. Chem. 2003, 24, 669–681. [Google Scholar] [CrossRef]
  39. Franchi, D.; Bartolini, M.; D’Amico, F.; Calamante, M.; Zani, L.; Reginato, G.; Mordini, A.; Dessì, A. Exploring Different Designs in Thieno[3,4-b]pyrazine-Based Dyes to Enhance Divergent Optical Properties in Dye-Sensitized Solar Cells. Processes 2023, 11, 1542. [Google Scholar] [CrossRef]
  40. Traverse, C.J.; Pandey, R.; Barr, M.C.; Lunt, R.R. Emergence of highly transparent photovoltaics for distributed applications. Nat. Energy 2017. [Google Scholar] [CrossRef]
  41. Meguellati, K.; Ladame, S.; Spichty, M. A conceptually improved TD-DFT approach for predicting the maximum absorption wavelength of cyanine dyes. Dye. Pigment. 2011, 90, 114–118. [Google Scholar] [CrossRef]
  42. Kumar, B.A.; Vetrivelan, V.; Ramalingam, G.; Manikandan, A.; Viswanathan, S.; Boomi, P.; Ravi, G. Computational studies and experimental fabrication of DSSC device assembly on 2D-layered TiO2 and MoS2@TiO2 nanomaterials. Phys. B Condens. Matter 2022, 633, 413770. [Google Scholar] [CrossRef]
  43. Li, F.; Jiang, X.; Zhao, J.; Zhang, S. Graphene oxide: A promising nanomaterial for energy and environmental applications. Nano Energy 2015, 16, 488–515. [Google Scholar] [CrossRef]
  44. Yu, X.Y.; Feng, Y.; Guan, B.; Lou, X.W.D.; Paik, U. Carbon coated porous nickel phosphides nanoplates for highly efficient oxygen evolution reaction. Energy Environ. Sci. 2016, 9, 1246–1250. [Google Scholar] [CrossRef]
  45. Sanjay, P.; Deepa, K.; Madhavan, J.; Senthil, S. Optical, spectral and photovoltaic characterization of natural dyes extracted from leaves of Peltophorum pterocarpum and Acalypha amentacea used as sensitizers for ZnO based dye sensitized solar cells. Opt. Mater. 2018, 83, 192–199. [Google Scholar] [CrossRef]
  46. Zoei, M.S.; Sadeghi, M.H.; Salehi, M. Effect of Grinding Parameters on the Fracture Toughness of WC–10Co–4Cr Coating Deposited by HVOF. Prog. Color. Color. Coat. 2019, 12, 231–239. [Google Scholar]
  47. Suriani, A.B.; Fatiatun Mohamed, A.; Muqoyyanah Hashim, N.; Rosmi, M.S.; Mamat, M.H.; Malek, M.F.; Salifairus, M.J.; Abdul Khalil, H.P.S. Reduced graphene oxide/platinum hybrid counter electrode assisted by custom-made triple-tail surfactant and zinc oxide/titanium dioxide bilayer nanocomposite photoanode for enhancement of DSSCs photovoltaic performance. Optik 2018, 161, 70–83. [Google Scholar] [CrossRef]
  48. Tamilselvi, C.; Duraisamy, P.; Subathra, N. Graphene wrapped NiSe2 nanocomposite-based counter electrode for dye-sensitized solar cells (DSSCs). Diam. Relat. Mater. 2021, 116, 108396. [Google Scholar] [CrossRef]
  49. Francis, M.K.; Santhosh, N.; Govindaraj, R.; Ahmed, N.; Balaji, C. Bifacial DSSC fabricated using low-temperature processed 3D flower like MoS2—High conducting carbon composite counter electrodes. Mater. Today Commun. 2021, 27, 102208. [Google Scholar] [CrossRef]
  50. Silambarasan, K.; Harish, S.; Hara, K.; Archana, J.; Navaneethan, M. Ultrathin layered MoS2 and N-doped graphene quantum dots (N-GQDs) anchored reduced graphene oxide (rGO) nanocomposite-based counter electrode for dye-sensitized solar cells. Carbon 2021, 181, 107–117. [Google Scholar] [CrossRef]
Figure 1. Chemical structure of photosensitizers.
Figure 1. Chemical structure of photosensitizers.
Micromachines 14 02161 g001
Figure 2. X-ray diffraction spectra of dyes 1–8.
Figure 2. X-ray diffraction spectra of dyes 1–8.
Micromachines 14 02161 g002
Figure 3. Linear fit plots of the Modified-Scherrer equation and the obtained intercepts to calculate crystallite size of dyes 1–8.
Figure 3. Linear fit plots of the Modified-Scherrer equation and the obtained intercepts to calculate crystallite size of dyes 1–8.
Micromachines 14 02161 g003
Figure 4. FESEM of the graphene flake.
Figure 4. FESEM of the graphene flake.
Micromachines 14 02161 g004
Figure 5. FESEM image of the MoS2/GO composite (a) and hybrid (b).
Figure 5. FESEM image of the MoS2/GO composite (a) and hybrid (b).
Micromachines 14 02161 g005
Figure 6. (a) FESEM micrographs of flower-shaped MoS2 nanoflakes on different scales and (b) zoom area.
Figure 6. (a) FESEM micrographs of flower-shaped MoS2 nanoflakes on different scales and (b) zoom area.
Micromachines 14 02161 g006
Figure 7. CV results of photosensitizers.
Figure 7. CV results of photosensitizers.
Micromachines 14 02161 g007
Figure 8. The optimized molecules @ S0 and HOMO–LUMO distribution of the dyes.
Figure 8. The optimized molecules @ S0 and HOMO–LUMO distribution of the dyes.
Micromachines 14 02161 g008
Figure 9. Simulated absorption spectra at TD-DFT/B3LYP/6-31g(d,p) of dyes.
Figure 9. Simulated absorption spectra at TD-DFT/B3LYP/6-31g(d,p) of dyes.
Micromachines 14 02161 g009
Table 1. Calculated quantum chemical parameters of dyes.
Table 1. Calculated quantum chemical parameters of dyes.
Dye12345678
μ (eV)−3.87−4.30−3.98−4.39−3.65−4.17−3.78−4.28
η (eV)3.853.493.853.443.132.993.213.01
ω (eV)1.952.652.062.802.132.912.233.05
Table 2. Photovoltaic data of DSSCs containing GO as counter electrode.
Table 2. Photovoltaic data of DSSCs containing GO as counter electrode.
DyeJSC (mA·cm−2)VOC (V)FFη (%)
15.120.450.461.06
26.160.510.421.32
37.590.530.441.77
49.090.520.442.08
53.080.540.440.73
64.810.510.441.08
73.860.510.661.30
85.920.540.451.44
N71912.920.680.423.69
Table 3. Photovoltaic data of DSSCs containing MoS2 as counter electrode.
Table 3. Photovoltaic data of DSSCs containing MoS2 as counter electrode.
DyeJSC (mA·cm−2)VOC (V)FF (%)η (%)
17.290.570.632.62
29.040.640.603.47
310.280.620.644.08
412.620.640.645.17
54.390.650.641.83
69.010.550.452.23
77.730.650.653.27
88.160.660.653.50
N71916.680.840.628.68
Table 4. Photovoltaic data of DSSCs containing GO-MoS2 nano-composite as counter electrode.
Table 4. Photovoltaic data of DSSCs containing GO-MoS2 nano-composite as counter electrode.
DyeJSC (mA·cm−2)VOC (V)FF (%)η (%)
18.170.550.632.83
29.520.620.613.60
310.300.630.654.22
414.390.620.645.71
54.900.630.652.01
67.130.620.642.83
78.730.620.653.52
88.870.640.663.75
N71918.650.810.629.37
Table 5. Photovoltaic data of DSSCs containing GO-MoS2 nano-hybrid as counter electrode.
Table 5. Photovoltaic data of DSSCs containing GO-MoS2 nano-hybrid as counter electrode.
DyeJSC (mA·cm−2)VOC (V)FF (%)η (%)
17.710.570.632.77
29.400.620.613.558
39.980.650.644.15
413.670.640.645.60
54.710.650.641.96
66.690.640.640.74
78.320.640.653.46
88.640.660.653.71
N71917.400.850.629.17
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

Hosseinnezhad, M.; Ghahari, M.; Mobarhan, G.; Fathi, M.; Palevicius, A.; Nutalapati, V.; Janusas, G.; Nasiri, S. New Insights into Improving the Photovoltaic Performance of Dye-Sensitized Solar Cells by Removing Platinum from the Counter Electrode Using a Graphene-MoS2 Composite or Hybrid. Micromachines 2023, 14, 2161. https://doi.org/10.3390/mi14122161

AMA Style

Hosseinnezhad M, Ghahari M, Mobarhan G, Fathi M, Palevicius A, Nutalapati V, Janusas G, Nasiri S. New Insights into Improving the Photovoltaic Performance of Dye-Sensitized Solar Cells by Removing Platinum from the Counter Electrode Using a Graphene-MoS2 Composite or Hybrid. Micromachines. 2023; 14(12):2161. https://doi.org/10.3390/mi14122161

Chicago/Turabian Style

Hosseinnezhad, Mozhgan, Mehdi Ghahari, Ghazal Mobarhan, Mohsen Fathi, Arvydas Palevicius, Venkatramaiah Nutalapati, Giedrius Janusas, and Sohrab Nasiri. 2023. "New Insights into Improving the Photovoltaic Performance of Dye-Sensitized Solar Cells by Removing Platinum from the Counter Electrode Using a Graphene-MoS2 Composite or Hybrid" Micromachines 14, no. 12: 2161. https://doi.org/10.3390/mi14122161

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