
As a cheaper alternative to the industrial Pt electrode used in quantum-sensitized solar cells, the electrophoresis process is employed to create the low-cost FTO/PbS cathode. For structural cubic and sizes ranging from 40 nm to 200 nm, structure and morphology were investigated using high-resolution scanning electron microscopy and X-ray diffraction. The conversion efficiency of solar cells is significantly impacted by the calcination temperatures of cathodes at 100 ℃, 150 ℃, 200 ℃, and 300 ℃ under vacuum. The FTO/PbS cathode electrode was therefore calcined at 150 ℃ with a maximum efficiency of 3.938%. This happens as a result of the complete fusion of PbS nanoparticles with crystal at 150 ℃, which reduces resistance and increases electron lifetime compared to other temperature combinations.
Citation: Ha Thanh Tung, Ho Kim Dan, Dang Huu Phuc. Effect of the calcination temperature of the FTO/PbS cathode on the performance of a quantum dot-sensitized solar cell[J]. AIMS Materials Science, 2023, 10(3): 426-436. doi: 10.3934/matersci.2023023
[1] | Chandni Lotwala, Hai-Feng Ji . Electrochemistry on nanopillared electrodes. AIMS Materials Science, 2017, 4(2): 292-301. doi: 10.3934/matersci.2017.2.292 |
[2] | Christian M. Julien, Alain Mauger . In situ Raman analyses of electrode materials for Li-ion batteries. AIMS Materials Science, 2018, 5(4): 650-698. doi: 10.3934/matersci.2018.4.650 |
[3] | Hiroki Nara, Keisuke Morita, Tokihiko Yokoshima, Daikichi Mukoyama, Toshiyuki Momma, Tetsuya Osaka . Electrochemical impedance spectroscopy analysis with a symmetric cell for LiCoO2 cathode degradation correlated with Co dissolution. AIMS Materials Science, 2016, 3(2): 448-459. doi: 10.3934/matersci.2016.2.448 |
[4] | Gen Long, Kenneth Sabalo, Raheeb Alsaidi, Michael Beattie, Blawal Chaudhry, Mohammad Khan, Juhayer Uddin, Mostafa Sadoqi . A study of PbS nanoparticle synthesis via sulfur powder. AIMS Materials Science, 2017, 4(2): 515-521. doi: 10.3934/matersci.2017.2.515 |
[5] | Dong Geun Lee, Hwan Chul Yoo, Eun-Ki Hong, Won-Ju Cho, Jong Tae Park . Device performances and instabilities of the engineered active layer with different film thickness and composition ratios in amorphous InGaZnO thin film transistors. AIMS Materials Science, 2020, 7(5): 596-607. doi: 10.3934/matersci.2020.5.596 |
[6] | Alexandre Lavrov, Benjamin Werner, Anna Stroisz, Thomas Monge Øia, Kamila Gawel, Malin Torsæter, Narjes Jafariesfad . Effect of electric field on push-out strength of cemented steel pipes. AIMS Materials Science, 2021, 8(3): 373-389. doi: 10.3934/matersci.2021024 |
[7] | Brittany L. Oliva-Chatelain, Andrew R. Barron . Experiments towards size and dopant control of germanium quantum dots for solar applications. AIMS Materials Science, 2016, 3(1): 1-21. doi: 10.3934/matersci.2016.1.1 |
[8] | Fahmi Astuti, Rima Feisy Azmi, Mohammad Arrafi Azhar, Fani Rahayu Hidayah Rayanisaputri, Muhammad Redo Ramadhan, Malik Anjelh Baqiya, Darminto . Employing Na2CO3 and NaCl as sources of sodium in NaFePO4 cathode: A comparative study on structure and electrochemical properties. AIMS Materials Science, 2024, 11(1): 102-113. doi: 10.3934/matersci.2024006 |
[9] | Pham Van Viet, Le Van Hieu, Cao Minh Thi . The directed preparation of TiO2 nanotubes film on FTO substrate via hydrothermal method for gas sensing application. AIMS Materials Science, 2016, 3(2): 460-469. doi: 10.3934/matersci.2016.2.460 |
[10] | Niklas Ehrlin, Christina Bjerkén, Martin Fisk . Cathodic hydrogen charging of Inconel 718. AIMS Materials Science, 2016, 3(4): 1350-1364. doi: 10.3934/matersci.2016.4.1350 |
As a cheaper alternative to the industrial Pt electrode used in quantum-sensitized solar cells, the electrophoresis process is employed to create the low-cost FTO/PbS cathode. For structural cubic and sizes ranging from 40 nm to 200 nm, structure and morphology were investigated using high-resolution scanning electron microscopy and X-ray diffraction. The conversion efficiency of solar cells is significantly impacted by the calcination temperatures of cathodes at 100 ℃, 150 ℃, 200 ℃, and 300 ℃ under vacuum. The FTO/PbS cathode electrode was therefore calcined at 150 ℃ with a maximum efficiency of 3.938%. This happens as a result of the complete fusion of PbS nanoparticles with crystal at 150 ℃, which reduces resistance and increases electron lifetime compared to other temperature combinations.
Inorganic semiconductor quantum dots (QDs), which produce electrons for use in solar cells and have highly powerful light absorption properties, were developed as such materials in recent years. CdS, CdSe, PbS, PbSe, and InP have all been used in quantum dot sensitized solar cells (QDSSCs) [1,2]. The QDs have several advantages over dye molecules, including the ability to modify the particle size to change the bandgap energy [3], a higher optical absorption coefficient than dye molecules [4], and the formation of many more exciton pairs when photons are absorbed. Currently, QDSSCs have a lower conversion efficiency than dye-sensitized solar cells (DSSCs) [5,6,7,8,9].
The amazing studies [9,10,11] are all carried out on single quantum dots, which hardly absorb sunlight and do not fully exploit the visible region. More specifically, the bulk materials' absorption wavelengths for CdS and CdSe are respectively 550 and 705 nm. Both absorption wavelengths are considerably shorter than the aforementioned values at the scale of the QDs. The performance improvement was rather slight when the binding linker was employed after that. Extending the peak of the absorption spectrum well into the visible region while lowering recombination and dark current are required to improve QDSSC performance. Numerous groups have investigated the issue of expanding the absorption ranger by co-quantum dots, and the yield is significantly higher than that of devices using single quantum dot [12,13,14,15].
The counter electrode (CE), which is essential to boosting the conversion energy of devices because it receives electrons from the external circuit via a redox process at the surface of the electrolyte and the CE, was the subject of research in addition to the photoanodic electrode. The CE must consequently have a large surface area, high conductivity, and high electrochemical activity in order for the electrolyte's redox reactions to occur more quickly and be chemically stable at a low cost [16].
In a quest to replace the traditional CE Pt, researchers have spent years looking for CEs that might accomplish the aforementioned objectives. The ability of metal sulfide compounds such Mo2S [17], NiS [18], FeS [19], CuS [20], and Cu2S [12] to satisfy the aforementioned criteria and replace CE Pt has been demonstrated by researchers. Compared to the other materials in this category, copper sulfide (CuS, Cu2S) is used in more studies because it has a band gap energy of 1.1–1.4 eV, a high electrochemical activity, and is stable in a polysulfide electrolyte [21,22,23]. According to the duration, temperature, and environment, thin copper (brass) films with a thickness of several micrometers are immersed in a polysulfide electrolyte to produce either a CuS thin film or a Cu2S thin film. The copper brass, however, will continue to degrade if immersed in the electrolyte for a long time. This will increase the resistance of the film and cause a drop in the fill factor and open circuit potential, which will reduce the efficiency of QDSSCs [21,22,23].
PbS, which has the most consistent electrochemical activity among the other materials, has been investigated, fabricated, and employed in QDSSCs with a high efficiency of 3% [24] by Tachan's group and colleagues. The results show the material's potential. However, PbS has not been studied using a variety of fabrication methods, nor has it been investigated how the crystallization temperature of PbS film affects the electrochemical properties of the electrode and the conversion efficiency of solar cells. An FTO/PbS electrode's production is influenced by a number of variables, including the ratio of precursors, temperature, and calcination environment. In the earlier article, we were able to successfully raise the precursor ratio. Therefore, in this publication, we focus on studying CE crystallization at the calcination temperature. Choosing the right annealing temperature is crucial for good crystallization since the electrical resistance of the CEs electrode effects crystallization. With better crystallization, the resistance of the crystal decreases. This is also the justification behind our choice to research and develop FTO/PbS cathode electrodes for the production of solar cells using electrophoresis at varied temperatures.
The substrate for the cathode electrode film is a thin sheet of FTO (fluorine-doped tin oxide), a conductor with a low resistance of roughly 7 Ω/cm2. The two diameters of commercial TiO2 paste—400 nm as a light reflector and 20 nm as a transmission layer—are employed for the anode electrode. A full solar cell is created by joining two electrodes together using Meltonix 1170-25 (Surlyn). Using Na2S2O3·5H2O and Pb(NO3)2 purchased from Sigma's company in Germany, we perform current and potential density measurements, electrochemical impedance measurements, and use colloidal silver as the path of the front and back of the solar cell. All of the aforementioned compounds are 99.9% pure.
The photoanode electrode is fabricated according to the process described in the group's previously published paper [25].
Commercial conductive glass (FTO, TEC7, Dyesol) was cut into several plates with an area of 1.2 cm × 2 cm each. The FTO plate is drilled with 2 small holes, 1 mm in diameter, used for injecting electrolyte solution. These bases are treated as shown in Figure 1.
The electrolyte solution used in this study is a polysulfide with an oxidation-reduction pair S2/Sn2– which was synthesized by dissolving 1.2 g of Na2S with 0.064 g of sulfur powder and 0.149 g of KCl in 10 mL of aqueous methanol solution with a ratio of 7:3 to obtain a yellow color.
The cathode electrode of the PbS film was coated on the FTO conductive substrate by electrophoresis in a mixed solution of 2.0 mM Pb(NO3)2 and 1.0 mM NaS2O3 and dissolved in 100 mL of distilled water and methanol; adjust pH to 2.65. A voltage of 100 mV was used for 1 h for all samples; we fixed the concentration and pH and only changed the electrode heating temperature at 100 ℃, 150 ℃, 200 ℃, and 300 ℃ for 1 h in vacuum. Finally, we assemble the solar cell according to the structure from Figure 2.
To study the properties of the cathode electrode, we used a high resolution surface scanning electromagnetic microscope at scanning potentials 15 kV and 1 kV (FESEM, SU820) for samples at different resolutions to analyze Geometric surface analysis of electrodes, Philips X-ray diffraction (XRD) spectroscopy, (PANalytical X'Pert) with CuKα radiation to determine the structure of PbS crystals on the FTO film, allowing size calculation particles and lattice parameters. After assembling the battery, we used the Keithley 2400 instrument and the SOLARENA sunlight simulation system to determine the current density and potential curves, thereby determining the conversion efficiency of the solar cells. Besides, in order to better understand the electron transfer processes through the cathode, we use electrochemical impedance spectroscopy to measure and determine the value of the resistances in the battery, including the resistances across the cathode surface with electrolyte to evaluate cathode activity.
The FTO/PbS cathode electrode is depicted in high resolution scanning electron microscopy images in Figure 3 at various resolution levels. In general, the film's surface is quite porous, making it ideal for the electrode's electrochemical activity. The porous surface leaves a lot of empty spaces, making it easier for the polysulfide electrolyte solution to come into direct contact with the PbS nanoparticles and carry out the electrochemical reaction. The electrolyte system's electron exchange. The PbS nanoparticles were found to range in size from 40 nm to 200 nm, and they were extremely uniformly shaped like square or rectangular blocks. It is shown that the electrophoresis-created film is extremely stable.
By using electrophoresis to create a strong adhesion film, the FTO/PbS electrode is created. To stop the formation of PbO, the film is crystallized in a vacuum. As a result, we first need to ascertain the structure of PbS in order to use it as electrodes. The FTO/PbS electrode's ray diffraction spectrum in a vacuum at 150 ℃ is shown in Figure 4, together with the PbS film's and the FTO substrate's distinctive diffraction peaks. The lattice planes with the Miller indices (111), (200), (220), (311), and (222) that correspond to the diffraction peaks at positions 26.1º, 30.2º, 43.19º, 51.61º, and 52.5º fit well with the standard JCPDS card No. 05-0592 of the face-centered cubic structure. Additionally, this outcome is entirely consistent with the findings of the research groups Varma [26], Yang [27], and Shyju [28]. The PbS face-centered cubic's distinctive peak is the plane (200). We estimate the average particle size of the PbS at 150 ℃ to be about 48.9511 nm using Scherer's Eq 1 (see Table 1), which is in agreement with the findings of the FESEM pictures. Additionally, we noticed the characteristic peaks of the FTO conductive film at diffraction angles of 26.64º, 33.85º, and 37.9º [29].
D=0.9λβcosθ | (1) |
Where β is the half-width of the characteristic peak of PbS and D is the grain diameter.
d-space (Å) | 2θ (º) | (hkl) | D (nm) |
3.4 | 26.1 | (111) | 48.9511 |
2.95 | 30.2 | (200) | |
2.09 | 43.19 | (220) | |
1.77 | 51.61 | (311) | |
2.74 | 52.5 | (222) |
After examining the electrode's structural and morphological characteristics, we attach the cathode electrode and the anode using surlyn that has been heated to more than 100 ℃. The electrolyte is pumped and measured at the end. The parameters, short circuit current density (Jsc), open circuit potential (Voc), fill factor, and conversion efficiency of QDSSCs are determined using current and voltage density curves. Figure 5 displays the characteristic curves, and Table 2 displays the calculated values of the parameter. Figure 5 shows that the electrode heated to 150 ℃ has the highest current density (18.18 mA/cm2) and the biggest open circuit voltage (0.572 V), resulting in a very high conversion efficiency of 3.938%. The conversion efficiency of this study is higher than that of studies by Tachan and co-workers (3%) and Zhang et al. (1.75%) [30], and it is in line with the findings of Yang and co-workers (3.91%) [27]. The achieved efficiency is substantially lower for the FTO/PbS cathode electrode heated at 100 ℃, 200 ℃, and 300 ℃, respectively: 2.579%, 1.929%, and 2.338%. As a result, it is clear that the temperature needed to create PbS crystals on the FTO substrate is crucial. This temperature has a significant impact on the electrochemical characteristics of the cathode electrode, which in turn impacts the current density in cells and the functionality of QDSSCs.
Electrodes | JSC (mA/cm2) | VOC (V) | Fill factor (FF) | Efficiency (%) | Rct1 (Ω) | Rct2 (Ω) | τn (ms) |
FTO/PbS-100 ℃ | 12.29 | 0.571 | 0.373 | 2.579 | 211.5 | 58.4 | 79.6 |
FTO/PbS-150 ℃ | 18.18 | 0.572 | 0.382 | 3.938 | 10.6 | 36.8 | 199 |
FTO/PbS-200 ℃ | 9.72 | 0.493 | 0.41 | 1.929 | 81.7 | 193 | 201.5 |
FTO/PbS-300 ℃ | 13.39 | 0.529 | 0.331 | 2.338 | 711.2 | 221 | 101.8 |
The speed of charge carriers in solar cells and the chemical composition of cathode electrodes are investigated using electrochemical impedance spectroscopy. We obtained the experimental EIS spectrum from the EIS spectrum in Figure 6 after measuring the solar cells, and we then used the specialized software Autolab to fit and calculate the resistance values of QDSSCs, such as: Rct1 is the resistance of the excited electrons that are transferred through the cathode and the cathode surface with polysulfide electrolyte, and Rct2 is the resistance of the excited electrons when moving through the photoanode to the external circuit. The EIS of QDSSCs, which were created using cathode electrodes heated at various temperatures in a vacuum atmosphere, is shown in Figure 6. Figure 6 illustrates this relationship between semi-circle size and resistance values; the cathode heated to 150 ℃ has the lowest Rct1 resistance (10.6 Ω), followed by cathodes heated to 100 ℃ (211 Ω), 200 ℃ (81.7 Ω), and 300 ℃ (711.2 Ω). This outcome can be explained by the fact that PbS layer on FTO crystallized more precisely at 150 ℃ than it did at higher heating temperatures. Additionally, all of the films were calcined in a vacuum to prevent Pb from oxidizing and becoming PbO. The outcomes of the XRD demonstrate this. Therefore, it can be said that the electrons moving through the PbS film and across its surface with the S2–/Sn2– electrolyte do so at a faster rate the smaller the resistance value of Rct1. The efficiency of the cells for the cathode heated at 150 ℃ is the highest due to its highest current density, and this result is entirely consistent with the results of the current and potential curve measurements of QDSSCs. This result is also entirely consistent with the findings of Zhang et al., who examined the PbS electrode film formation time [26] and compared it to that of the Pt electrode [30].
The bode spectrum, which is depicted in Figure 7, is derived from electrochemical impedance spectroscopy and is used to determine how well excited electrons may recombine and how long electrons in the PbS film's conduction band live. The calculation results for electrodes heated at 100 ℃, 150 ℃, 200 ℃, and 300 ℃ for 1 h in a vacuum atmosphere are shown in Table 2. This is related to the minimum frequency by the formula τmax=12πfmin. According to Table 2, when two cathode electrodes are heated to 150 ℃, the lifetime of QDSSCs is the longest (199 ms and 201.5 ms) at 200 ℃, and it is subsequently half at 300 ℃ (101.8 ms). The electrode is heated to 100 ℃ for a brief period of time (79.6 ms) since PbS crystals didn't crystallize perfectly. The measured current density is larger than that of other electrodes because the longer the lifespan, the less likely it is that the electron will recombine with the hole in the material's valence band. This result is in perfect accordance with the findings of current and potential density (J-V) curve measurements as well as X-ray diffraction (PbS crystals crystallize most efficiently at 150 ℃).
Electrophoresis is used to create the FTO/PbS cathode electrode at a voltage of 100 mV, which changes depending on the calcination temperature. Due to its reduced cost and superior performance, the film is utilized to replace commercial Pt electrodes. The PbS nanofilm was effectively created on the FTO conductive substrate as a result, and it possesses a face-centered cubic structure, excellent crystallinity, and somewhat uniform particle size, according to FESEM and XRD. We note the highest current density and efficiency performance at 18.18 mA/cm2 and 3.938% with an FTO/PbS cathode electrode at 150 ℃ when compared to other heating temperatures based on the findings of measuring current density and potential. What's more, this outcome was better than that of other research teams. The electrochemical impedance spectrum, commonly known as the bode spectrum, can also be used to explain the performance's outcomes. According to the findings, the PbS film was calcined at 150 ℃ for flawless crystallization, which resulted in fewer crystal lattice defects and recombination processes, as well as the least Rct1 resistance (10.6 Ω) and the longest lifetime (199 ms).
The authors declare that they have no conflict of interest.
[1] |
Zaban AMOI, Mićić OI, Gregg BA, et al. (1998) Photosensitization of nanoporous TiO2 electrodes with InP quantum dots. Langmuir 14: 3153–3156. https://doi.org/10.1021/la9713863 doi: 10.1021/la9713863
![]() |
[2] |
Yu P, Zhu K, Norman AG, et al. (2006) Nanocrystalline TiO2 solar cells sensitized with InAs quantum dots. J Phys Chem B 110: 25451–25454. https://doi.org/10.1021/jp064817b doi: 10.1021/jp064817b
![]() |
[3] |
Yu WW, Wang YA, Peng X (2003) Formation and stability of size-, shape-, and structure-controlled CdTe nanocrystals: ligand effects on monomers and nanocrystals. Chem Mater 15: 4300–4308. https://doi.org/10.1021/cm034729t doi: 10.1021/cm034729t
![]() |
[4] |
Beard MC (2011) Multiple exciton generation in semiconductor quantum dots. J Phys Chem Lett 2: 1282–1288. https://doi.org/10.1021/jz200166y doi: 10.1021/jz200166y
![]() |
[5] |
Fang J, Wu J, Lu X, et al. (1997) Sensitization of nanocrystalline TiO2 electrode with quantum sized CdSe and ZnTCPc molecules. Chem Phys Lett 270: 145-151. https://doi.org/10.1016/S0009-2614(97)00333-3 doi: 10.1016/S0009-2614(97)00333-3
![]() |
[6] |
Lee W, Kwak WC, Min SK, et al. (2008) Spectral broadening in quantum dots-sensitized photoelectrochemical solar cells based on CdSe and Mg-doped CdSe nanocrystals. Electrochem Commun 10: 1699–1702. https://doi.org/10.1016/j.elecom.2008.08.025 doi: 10.1016/j.elecom.2008.08.025
![]() |
[7] |
Liu D, Kamat PV (1993) Photoelectrochemical behavior of thin cadmium selenide and coupled titania/cadmium selenide semiconductor films. J Physical Chemistry 97: 10769-10773. https://doi.org/10.1021/j100143a041 doi: 10.1021/j100143a041
![]() |
[8] |
Peter LM, Riley DJ, Tull EJ, et al. (2002) Photosensitization of nanocrystalline TiO2 by self-assembled layers of CdS quantum dots. Chem Commun 10: 1030–1031. https://doi.org/10.1039/b201661c doi: 10.1039/b201661c
![]() |
[9] |
Vogel R, Pohl K, Weller H (1990) Sensitization of highly porous, polycrystalline TiO2 electrodes by quantum sized CdS. Chem Phys Lett 174: 241–246. https://doi.org/10.1016/0009-2614(90)85339-E doi: 10.1016/0009-2614(90)85339-E
![]() |
[10] |
Mora-Seró I, Giménez S, Moehl T, et al. (2008) Factors determining the photovoltaic performance of a CdSe quantum dot sensitized solar cell: the role of the linker molecule and of the counter electrode. Nanotechnology 19: 424007. https://doi.org/10.1088/0957-4484/19/42/424007 doi: 10.1088/0957-4484/19/42/424007
![]() |
[11] |
Shen YJ, Lee YL (2008) Assembly of CdS quantum dots onto mesoscopic TiO2 films for quantum dot-sensitized solar cell applications. Nanotechnology 19: 045602. https://doi.org/10.1088/0957-4484/19/04/045602 doi: 10.1088/0957-4484/19/04/045602
![]() |
[12] |
Lee YL, Lo YS (2009) Highly efficient quantum-dot-sensitized solar cell based on co‐sensitization of CdS/CdSe. Adv Funct Mater 19: 604–609. https://doi.org/10.1002/adfm.200800940 doi: 10.1002/adfm.200800940
![]() |
[13] |
Zhang Q, Zhang Y, Huang S, et al. (2010) Application of carbon counterelectrode on CdS quantum dot-sensitized solar cells (QDSSCs). Electrochem Commun 12: 327–330. https://doi.org/10.1016/j.elecom.2009.12.032 doi: 10.1016/j.elecom.2009.12.032
![]() |
[14] |
Sudhagar P, Jung JH, Park S, et al. (2009) The performance of coupled (CdS: CdSe) quantum dot-sensitized TiO2 nanofibrous solar cells. Electrochem Commun 11: 2220–2224. https://doi.org/10.1016/j.elecom.2009.09.035 doi: 10.1016/j.elecom.2009.09.035
![]() |
[15] |
Yu Z, Zhang Q, Qin D, et al. (2010) Highly efficient quasi-solid-state quantum-dot-sensitized solar cell based on hydrogel electrolytes. Electrochem Commun 12: 1776–1779. https://doi.org/10.1016/j.elecom.2010.10.022 doi: 10.1016/j.elecom.2010.10.022
![]() |
[16] |
Wang S, Tian J (2016) Recent advances in counter electrodes of quantum dot-sensitized solar cells. RSC Adv 6: 90082–90099. https://doi.org/10.1039/C6RA19226B doi: 10.1039/C6RA19226B
![]() |
[17] |
Kamaja CK, Devarapalli RR, Dave Y, et al. (2016) Synthesis of novel Cu2S nanohusks as high performance counter electrode for CdS/CdSe sensitized solar cell. J Power Sources 315: 277–283. https://doi.org/10.1016/j.jpowsour.2016.03.027 doi: 10.1016/j.jpowsour.2016.03.027
![]() |
[18] |
Kim HJ, Kim DJ, Rao SS, et al. (2014) Highly efficient solution processed nanorice structured NiS counter electrode for quantum dot sensitized solar cells. Electrochim Acta 127: 427–432. https://doi.org/10.1016/j.electacta.2014.02.019 doi: 10.1016/j.electacta.2014.02.019
![]() |
[19] |
Chen H, Zhu L, Liu H, et al. (2014) Efficient iron sulfide counter electrode for quantum dots-sensitized solar cells. J Power Sources 245: 406–410. https://doi.org/10.1016/j.jpowsour.2013.06.004 doi: 10.1016/j.jpowsour.2013.06.004
![]() |
[20] |
Raj CJ, Prabakar K, Savariraj AD, et al. (2013) Surface reinforced platinum counter electrode for quantum dots sensitized solar cells. Electrochim Acta 103: 231–236. https://doi.org/10.1016/j.electacta.2013.04.016 doi: 10.1016/j.electacta.2013.04.016
![]() |
[21] |
Selopal GS, Zhao H, Tong X, et al. (2017) Highly stable colloidal "giant" quantum dots sensitized solar cells. Adv Funct Mater 27: 1701468. https://doi.org/10.1002/adfm.201701468 doi: 10.1002/adfm.201701468
![]() |
[22] |
Jiao S, Shen Q, Mora-Sero I, et al. (2015) Band engineering in core/shell ZnTe/CdSe for photovoltage and efficiency enhancement in exciplex quantum dot sensitized solar cells. ACS Nano 9: 908–915. https://doi.org/10.1021/nn506638n doi: 10.1021/nn506638n
![]() |
[23] |
Kim JY, Yang J, Yu JH, et al. (2015) Highly efficient copper-indium-selenide quantum dot solar cells: suppression of carrier recombination by controlled ZnS overlayers. ACS Nano 9: 11286–11295. https://doi.org/10.1021/acsnano.5b04917 doi: 10.1021/acsnano.5b04917
![]() |
[24] |
Tachan Z, Shalom M, Hod I, et al. (2011) PbS as a highly catalytic counter electrode for polysulfide-based quantum dot solar cells. J Phys Chem C 115: 6162–6166. https://doi.org/10.1021/jp112010m doi: 10.1021/jp112010m
![]() |
[25] |
Nguyen TP, Ha TT, Nguyen TT, et al. (2018) Effect of Cu2+ ions doped on the photovoltaic features of CdSe quantum dot sensitized solar cells. Electrochim Acta 282: 16–23. https://doi.org/10.1016/j.electacta.2018.06.046 doi: 10.1016/j.electacta.2018.06.046
![]() |
[26] |
Thulasi-Varma CV, Rao SS, Ikkurthi KD, et al. (2015) Enhanced photovoltaic performance and morphological control of the PbS counter electrode grown on functionalized self-assembled nanocrystals for quantum-dot sensitized solar cells via cost-effective chemical bath deposition. J Mater Chem C 3: 10195–10206. https://doi.org/10.1039/C5TC01988E doi: 10.1039/C5TC01988E
![]() |
[27] |
Yang Y, Zhu L, Sun H, et al. (2012) Composite counter electrode based on nanoparticulate PbS and carbon black: towards quantum dot-sensitized solar cells with both high efficiency and stability. ACS Appl Mater Interfaces 4: 6162–6168. https://doi.org/10.1021/am301787q doi: 10.1021/am301787q
![]() |
[28] |
Shyju TS, Anandhi S, Sivakumar R, et al. (2014) Studies on lead sulfide (pbs) semiconducting thin films deposited from nanoparticles and its nlo application. Int J Nanosci 13: 1450001. https://doi.org/10.1142/S0219581X1450001X doi: 10.1142/S0219581X1450001X
![]() |
[29] |
Turgut G, Koçyiğit A, Sönmez E (2015) Influences of Pr and Ta doping concentration on the characteristic features of FTO thin film deposited by spray pyrolysis. Chinese Phys B 24: 107301. https://doi.org/10.1088/1674-1056/24/10/107301 doi: 10.1088/1674-1056/24/10/107301
![]() |
[30] |
Zhang JB, Zhao FY, Tang GS, et al. (2013) Influence of highly efficient PbS counter electrode on photovoltaic performance of CdSe quantum dots-sensitized solar cells. J Solid State Electr 17: 2909–2915. https://doi.org/10.1007/s10008-013-2210-4 doi: 10.1007/s10008-013-2210-4
![]() |
d-space (Å) | 2θ (º) | (hkl) | D (nm) |
3.4 | 26.1 | (111) | 48.9511 |
2.95 | 30.2 | (200) | |
2.09 | 43.19 | (220) | |
1.77 | 51.61 | (311) | |
2.74 | 52.5 | (222) |
Electrodes | JSC (mA/cm2) | VOC (V) | Fill factor (FF) | Efficiency (%) | Rct1 (Ω) | Rct2 (Ω) | τn (ms) |
FTO/PbS-100 ℃ | 12.29 | 0.571 | 0.373 | 2.579 | 211.5 | 58.4 | 79.6 |
FTO/PbS-150 ℃ | 18.18 | 0.572 | 0.382 | 3.938 | 10.6 | 36.8 | 199 |
FTO/PbS-200 ℃ | 9.72 | 0.493 | 0.41 | 1.929 | 81.7 | 193 | 201.5 |
FTO/PbS-300 ℃ | 13.39 | 0.529 | 0.331 | 2.338 | 711.2 | 221 | 101.8 |
d-space (Å) | 2θ (º) | (hkl) | D (nm) |
3.4 | 26.1 | (111) | 48.9511 |
2.95 | 30.2 | (200) | |
2.09 | 43.19 | (220) | |
1.77 | 51.61 | (311) | |
2.74 | 52.5 | (222) |
Electrodes | JSC (mA/cm2) | VOC (V) | Fill factor (FF) | Efficiency (%) | Rct1 (Ω) | Rct2 (Ω) | τn (ms) |
FTO/PbS-100 ℃ | 12.29 | 0.571 | 0.373 | 2.579 | 211.5 | 58.4 | 79.6 |
FTO/PbS-150 ℃ | 18.18 | 0.572 | 0.382 | 3.938 | 10.6 | 36.8 | 199 |
FTO/PbS-200 ℃ | 9.72 | 0.493 | 0.41 | 1.929 | 81.7 | 193 | 201.5 |
FTO/PbS-300 ℃ | 13.39 | 0.529 | 0.331 | 2.338 | 711.2 | 221 | 101.8 |