Next Article in Journal
Highly Transparent and Zirconia-Enhanced Sol-Gel Hybrid Coating on Polycarbonate Substrates for Self-Cleaning Applications
Previous Article in Journal
Feature Selection in Machine Learning for Perovskite Materials Design and Discovery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of Nanostructured Sn/Ti Oxide Hybrid Films with Terpineol/PEG-Based Nanofluids: Perovskite Solar Cell Applications

1
Mechanical Engineering Department, Bradley University, 1501 West Bradley Avenue, Peoria, IL 61625, USA
2
Industrial and Manufacturing Engineering and Technology Department, Bradley University, 1501 West Bradley Avenue, Peoria, IL 61625, USA
3
Graduate School of Natural Science and Technology, Gifu University, Yanagido 1-1, Gifu 501-1193, Japan
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(8), 3136; https://doi.org/10.3390/ma16083136
Submission received: 14 March 2023 / Revised: 13 April 2023 / Accepted: 13 April 2023 / Published: 16 April 2023

Abstract

:
Tin oxide (SnO2) and titanium dioxide (TiO2) are recognized as attractive energy materials applicable for lead halide perovskite solar cells (PSCs). Sintering is one of the effective strategies for improving the carrier transport of semiconductor nanomaterials. Using the alternative metal-oxide-based ETL, nanoparticles are often used in a way that they are dispersed in a precursor liquid prior to their thin-film deposition. Currently, the creation of PSCs using nanostructured Sn/Ti oxide thin-film ETL is one of the topical issues for the development of high-efficiency PSCs. Here, we demonstrate the preparation of terpineol/PEG-based fluid containing both tin and titanium compounds that can be utilized for the formation of a hybrid Sn/Ti oxide ETL on a conductive substrate (F-doped SnO2 glass substrate: FTO). We also pay attention to the structural analysis of the Sn/Ti metal oxide formation at the nanoscale using a high-resolution transmission electron microscope (HR-TEM). The variation of the nanofluid composition, i.e., the concentration of tin and titanium sources, was examined to obtain a uniform transparent thin film by spin-coating and sintering processes. The maximum power conversion efficiency was obtained for the concentration condition of [SnCl2·2H2O]/[titanium tetraisopropoxide (TTIP)] = 25:75 in the terpineol/PEG-based precursor solution. Our method for preparing the ETL nanomaterials provides useful guidance for the creation of high-performance PSCs using the sintering method.

1. Introduction

Thin solid films consisting of semiconductor nanomaterials have received increasing attention for their potential applications, such as optoelectronic devices [1,2,3]. Among them, perovskite solar cells (PSCs) built from multi-semiconductor layers including halide perovskite light absorbers are now attracting widespread interest, in view of not only the intriguing basic science of related materials but also their commercialization [4,5,6]. These devices are also expected to be integrated into versatile renewable energy systems [7]. In particular, with the need for alternative energy sources, materials research for PSC applications has gained tremendous progress in the improvement of the light-to-electricity conversion of PSCs, as well as the durability and diversity of devices with various transparent conductive substrates. One of the major device structures of PSCs realizing a decent power conversion efficiency includes the use of metal oxide nanoparticle layers that are deposited on top of conductive F-doped tin oxide (FTO) glass substrates [4]. These are known as mesoscopic-type or planar-type devices, for which lead halide perovskite is crystallized adjacent to the metal oxide layers. As for the bottom layer, the following two processes are alternatively employed: annealing-free deposition at temperatures as low as 100 °C toward the development of flexible PSCs and high-temperature sintering, generally treated at 450~500 °C using an electric oven. For the latter case, heating of the metal oxide precursor films can induce the fusion of the corresponding nanoparticles to ensure efficient electron transport in the device.
There have been several reports on the advantages of using multi-layer-type SnO2/TiO2 or Sn-doped TiO2 as an electron transport layer (ETL) in PSCs. As an example, Guo et al. found that TiO2/SnO2 nanocomposites prepared by a simple one-step process had high electrical conductivity and favorable energy band alignment at the interface of ETLs/lead halide perovskite layers [8]. They demonstrated that the inclusion of SnO2 nanoparticles in a compact TiO2 layer had a positive result in increasing power conversion efficiency (PCE). SnO2 nanoparticle powder was added into a TiO2 precursor solution to produce nanocomposites, followed by spin-coating and sintering at 500 °C. A maximum PCE of 16.8% was reported for CH3NH3PbI3-based solar cells, showing a superior PCE as opposed to that of TiO2-based PSCs.
Another strategy regarding a bilayer ETL was reported by Hu et al. [9]. They pointed out that cascading energy levels of integrated SnO2/TiO2 layers and TiO2-mediated defect passivation of SnO2 contributed to photo-excited energy loss. Here, the SnO2/TiO2 designed for the ETL was made of nanoparticles, in which colloidal SnO2 nanoparticles are commercial products based on solution processes. The deposition was carried out via subsequent spin-coating and annealing at 150 °C. As a result, a marked maximum light-to-electricity conversion efficiency of 20.50% was achieved when a multi-cation-type perovskite (FAPbI3)1−x(MAPbBr3)x was deposited on top of the ETL. It was found that the bilayer structure led to a significant improvement in PCE when it was compared to that of the SnO2-only control device (18.09%). Additionally, the TiO2 overcoat for the SnO2 layer made it possible to enhance the long-term stability of the device.
In regard to the passivation of ETLs, Abuhelaiqa et al. [10] pointed out that SnO2 films in bilayer SnO2/TiO2 ETLs are effective in the improvement of the device stability of PSCs. The degradation of perovskite film, especially the interface between the ETL and perovskite, was retarded, and PbI2 formation could also be suppressed. The additional SnO2 was shown to reduce charge recombination at the material interface, and improved optical durability was also reported [10].
A low-temperature process has also been examined to investigate the effect of Sn doping in TiO2 on perovskite solar cell performance [11]. The authors demonstrated that Sn(IV) can be introduced into the TiO2 lattice, which was possible by the addition of SnCl4·5H2O into an aqueous TiCl4 solution. Film preparation was carried out at a temperature of less than 100 °C. As opposed to the performance of pristine TiO2, they showed more efficient electron extraction and transport, giving rise to a higher short-circuit current density (Jsc). Their optimized solar cell employing Sn-doped TiO2 had a PCE of 17.2%. This was a 29.3% higher value than that of the non-doped TiO2 device. A higher charge carrier mobility of Sn-doped TiO2 compared to that of the non-doped type has also been reported by other authors [12]. They showed an optimized PCE of 15.4% by introducing 1.0 mol% Sn-doped TiO2 in their perovskite solar cell.
An ETL consisting of Sn-TiO2 was also processed on an FTO substrate by co-electrodeposition and changing the doping amount of Sn [13]. The authors mentioned that the deposition rate was faster than that of the pristine type because of less ohmic overpotential. The use of a 5% Sn-doped SnO2 ETL enabled the maximum PCE of 16.8% to be obtained, which was a 9.8% increase compared to that of the non-doped TiO2 device. They pointed out that improved carrier concentration, as well as the increased extraction capability for Sn doping, led to higher PCE.
Multi-component-type ETLs thus have significant potential to create high-performance PSCs. Further investigation into the relationships between Sn/Ti oxide 3D nanostructures and device performance would be a major issue to be addressed. The purpose of this research is focused on fully understanding the formation of micro-structured Sn/Ti oxide thin films. XRD and TEM measurements were carried out for the detailed analysis of Sn/Ti metal oxide samples, which were prepared by mixing Tin(II) chloride and Titanium(IV) isopropoxide in a suitable base liquid and additive. Notably, we found out that nanofluid preparation using a terpineol/polyethylene glycol (PEG) medium is a key process to obtaining a uniform thin film. In addition, the formation mechanism of Sn/Ti oxide nanoparticles involving the elimination of the base liquid and additive was clarified. The sintering effects of nanostructured semiconductor layers on device performance were discussed towards the development of lead halide PSCs showing high efficiencies.

2. Materials and Methods

2.1. Chemicals

Titanium tetraisoproxide (TTIP, >97.0%), N,N-dimethylformamide (DMF, dehydrated, >99.5%), dimethyl sulfoxide (DMSO, dehydrated, 99.0%), 2-Propanol (IPA, dehydrated, >99.7%), and acetonitrile (dehydrated, >99.5%) were purchased from Kanto Chemical Co., Inc. (Portland, OR, USA). Spiro-MeOTAD, chlorobenzene (CB, anhydrous, 99.8%), Terpineol (anhydrous), and 4-tert-butylpyridine (TBP, 98%) were obtained from Sigma-Aldrich, Co. (St, Louis, MI, USA). Methylamine hydrobromide (MABr, >98.0%), Cesium iodide (CsI, >99.0%), Formamidine hydroiodide (FAI, 99.99%), Lead(II) iodide (PbI2, 99.99%), Lithium bis(trifluoromethanesulfonly)imide (Li-TFSI, >98.0%), and Formamidine hydrobromide (FABr, 99.99%) were purchased from Tokyo Chemical Industry Co., Ltd. (Tokyo, Japan). Polyethylene glycol 400 (PEG400) and Tin(II) chloride dihydrate (SnCl2·2H2O, >97.0%) were from FUJIFILM Wako Pure Chemical Corporation, Richmond, VA, USA. All chemicals were used as purchased.

2.2. Preparation of Sn/Ti Oxide Thin Films

F-doped tin oxide glass substrates (11 Ω/□ Nippon Sheet Glass, Tokyo, Japan) were washed and cleaned with UV–ozone treatment. We first mixed 40 wt% (0.7 g) of ethanol with 40 wt% (0.7 g) of terpineol as a base liquid and then added 20 wt% (0.35 g) of PEG400 to prepare a base solvent.
Several precursor solutions changing the concentrations of metal sources were prepared as follows: To the mixture of base liquid and PEG, SnCl2·2H2O and TTIP were added. The ratios of the metal sources were [SnCl2]/[TTIP] = 100:0, 75:25, 50:50, 25:75, 0:100 (mol/mol). The concentrations of the metal sources in all those solutions were 10 wt%. For example, 0.39 g of SnCl2·2H2O was used for preparing the 1:0 condition. Typically, the 10 wt% solution containing metal sources was stirred for 30 min at ambient temperature, followed by spin-coating on top of the FTO substrate at 3000 rpm for 30 s. The substrates were immediately heated on a hot plate at 120 °C for 20 min.

2.3. Sintering Process

The substrates were placed in an electric furnace, KDF 300 Plus (DENKEN-HIGHDENTAL, Kyoto, Japan) for the sintering process. In regard to the sintering strategy, the temperature was increased from room temperature to 500 °C. The two different ramping times of 2 h and 30 min were examined. After the maximum temperature was kept at 500 °C, it was then cooled down naturally. The periods of maximum temperature were 2 h and 30 min, respectively (see Figure 1).

2.4. Evaluation of Sn/Ti Oxide Thin Films

Nanostructures of TiO2 films after nanofluid pool boiling were characterized by scanning electron microscopy (SEM) of HITACHI S-4800 (Hitachi, Ltd., Tokyo, Japan). The transmittance spectrum of the Sn/Ti oxide film was measured using PerkinElmer Lambda950 (PerkinElmer, Waltham, MA, USA). The perovskite solar cell was assembled using the modified method using a triple-cation type lead halide perovskite that has been demonstrated previously [14].
After the sintering process of ETL, UV–ozone treatment was conducted prior to spin-coating the halide perovskite precursor solution. Nitrogen was flowed into a glove box to keep the humidity at around 20 %. The substrates were placed in the glove box and heated to 80 °C on a hot plate. To prepare a perovskite precursor solution with a composition of Cs0.05MA0.1FA0.85PbI2.9Br0.1·0.05PbI2, MABr 0.14 mmol (0.0157 g), CsI 0.07 mmol (0.0182 g), FAI 1.19 mmol (0.204 g), and PbI2 1.45 mmol (0.666 g) were dissolved in the mixed solvent of DMF 80 vol% (0.8 mL) and DMSO 20 vol% (0.2 mL). It was then stirred on a hot plate in the dark at 80 °C for 2 h. The heated substrates at 80 °C were set on the spin-coater, and 100 μL of perovskite precursor solution was dropped onto the substrate. A two-step program of spin-coater (1000 rpm for 10 s and 6000 rpm for 30 s) was used. A 200 μL volume of CB was dropped with an electronic micropipette 25 s after the start of spin-coating. The substrate was annealed using a hot plate at 100 °C for 1 h, and it was allowed to cool naturally to room temperature. Then, 0.12 mmol (0.0148 g) of FABr was added to 4 mL of IPA to prepare a solution (30 mM), and it was spin-coated on top of the crystallized perovskite layer. A 100 μL volume of the prepared FABr solution was coated at 3000 rpm for 30 s. The substrate was heated on a hot plate at 80 °C for 10 min. After heating, the substrate was allowed to cool to room temperature.
To prepare a solution for the hole transporting layer (HTL), 0.037 mmol (0.045 g) of Spiro-MeOTAD was dissolved in 0.5 mL of CB. Next, 0.18 mmol (0.052 g) of Li-TFSI was dissolved in 0.1 mL of acetonitrile. A 10 μL volume of the acetonitrile solution and 17.75 μL of TBP were added to the solution containing Spiro-MeOTAD. The substrate was cooled to room temperature and spin-coated at 3000 rpm for 30 s. After 5 s from the start of spin-coating, 60 μL of the solution was dropped. The substrates were left for 24 h. Finally, a gold layer was deposited on top of Spiro-MeOTAD by using the thermal evaporation method [15].
Current–voltage curves were obtained under AM 1.5 simulated sunlight (100 mW/cm2) using a solar simulator (Yamashita Denso, YSS-80A, Tokyo, Japan) and potentiostat (Hokuto Denko HSV-110, Tokyo, Japan). The active area of the device was 0.09 cm2. The current–voltage curves were measured under reverse scanning conditions at a scan rate of 50 mV/s.

3. Results and Discussions

The fluid containing precursor materials of tin and titanium ions plays a vital role in determining the resultant film characteristics, such as uniformity of the obtained nanoparticle layers and density of nanocrystals in the film. We attempted to establish a sintering protocol of Sn/Ti oxide thin films applicable for electron transport layers in PSCs. In principle, base liquids and/or additive compounds in the fluid are required to optimize both the concentration of the metal ion sources and the packing structure of formed nanoparticles during heat treatment. By increasing the temperature, hydrolysis and polycondensation of those metal ions start to promote the oxide nanoparticle formation, where the homogeneous reactions should be adjusted in the precursor film. In addition to the use of both ethanol and terpineol to dissolve metal ions, we added a PEG polymer in order to disperse the obtained nanoparticles in the film. PEG is well-known to stabilize metal oxide nanoparticles preventing agglomeration, such as TiO2. The sizes of those nanoparticles can be reduced by steric repulsion.
The preliminary experiments allowed us to use the mixed base liquid, i.e., ethanol/terpineol, because the resultant fluid using only ethanol did not give a uniform film after the sintering treatment and instead some cloudy spots were observed in the film. We anticipated that terpineol would be a key chemical, as has been disclosed in the nanoparticle sintering research on dye-sensitized solar cells [16]. In this device, TiO2 nanoparticle films, which is the host material of light-absorbing dye molecules, can be prepared by a similar sintering approach with a precursor solution comprising suitable base liquids and additives.
The concentration dependence of tin and titanium ions on the film quality was systematically examined, and typical photos of the processed films after sintering are depicted in Figure 2. White agglomerated particles formed when terpineol was not used as a base liquid (Figure 2d–f), whereas transparent compact thin films with an approximate thickness of 100 nm as mentioned later were obtained for the ethanol/terpineol base liquid (Figure 2a–c). It appears that the terpineol with the higher boiling point (~100 °C) than ethanol assisted the homogenous reaction as a solvent, leading to the formation of well-dispersed nanoparticle films. It is likely that PEG also worked as a base liquid similar to the other solvents in a way that the inhomogeneous reactions were retarded due to the higher boiling temperature (>200 °C) than those of ethanol and terpineol.
X-ray diffraction patterns (XRD) of the Sn/Ti oxide samples sintered at 500 °C were analyzed to structurally identify crystalline compounds. Figure 1 depicts the two sintering strategies changing the ramping time, (a) 2 h and (b) 30 min, and periods of maximum temperature, (a) 2 h and (b) 30 min. As a result of heat treatment, XRD patterns exhibiting several broad peaks were obtained, which are presented in Figure 3.
The observed peaks for the measurements (Sn/Ti = 100:0, 75:25, and 50:50) were basically assigned to be those of rutile-phase SnO2. As indicated by the database of SnO2 (JCPDS:01-077-0452) and TiO2 (JCPDS: 01-076-0324), those peaks with similar diffraction angles, such as (110) peaks, hindered detailed analysis on crystallite sizes. Using the peaks corresponding to the (101) plane of SnO2, crystallite sizes were estimated from the Scherrer equation (1). D, K, λ, and θ indicate the crystallite size of SnO2, Scherrer constant (= 0.90), X-ray wavelength (= 1.54 Å), and Bragg angle, respectively.
D = Kλ/βcosθ
For 30 min sintering samples, the estimated crystallite sizes were 19 nm, 3 nm, and 6 nm for 100:0, 75:25, and 50:50, respectively. For 2 h sintering, the crystallite sizes were 4 nm, 3 nm, and 3 nm for 100:0, 75:25, and 50:50, respectively. Additionally, it turned out that TiO2 was not well crystallized for all those conditions. The more efficient crystallization at lower temperatures for SnO2 probably suppressed the TiO2 crystallization during the heat treatment. The time-dependent analysis via heat treatment suggested that crystal growth was not enhanced with the prolonged heating.
In contrast, TiO2 crystallization was observed both for Sn/Ti = 25:75 and 0:100. For the smaller amount of Sn source (Sn/Ti = 25:75), it is likely that one of the components of the base liquid, i.e., PEG, retards the crystal growth of SnO2, allowing to produce the Rutile TiO2 with estimated crystal sizes of less than 2 nm. Since SnO2 has a tendency to form a rutile crystal face, this may lead to the Rutile TiO2 formation, whereas for 0:100, only Anatase TiO2 grew as a result of the heat treatment.
To investigate the mechanism of film formation of Sn/Ti oxide hybrid films, a transmission electron microscope (TEM) was utilized in order to clarify the microstructure of nanoparticles at a single nanometer scale associated with the formation of mixed oxide layers from titanium and tin compounds. Figure 4 and Figure 5 present several TEM images obtained with different resolutions (Sn/Ti = 50:50), where the images from different positions are also depicted to show the observed uniformity of the samples. The samples prepared by two different sintering strategies (as shown in Figure 1) were measured, and the results are separately illustrated in Figure 4 and Figure 5. As can be seen in Figure 4a,b, it turned out that single-nanometer-scale particles with the approximate dimension of 3~5 nm were formed during heat treatment. The HR-TEM images in Figure 4c indicate lattice fringes exhibiting the (110) and (101) planes of rutile crystal-phase SnO2 and the (110) plane of rutile TiO2. These d-spacings were 0.33 nm, 0.26 nm, and 0.31 nm, respectively. In fact, the calculated crystallite size was consistent with that of the primary nanocrystal size observed in the images that were measured at different positions (Figure 4d). Irregular-shaped nanocrystals as a result of their fusion were more striking for the longer sintering (the period of maximum temperature: 2 h) than those of 30 min sintering (Figure 5c,d). In other words, more densely packed films were obtained for the 2 h sintering. Similarly, it was found from HR-TEM images of Figure 5c that lattice fringes were assigned to the (110) of rutile SnO2 and the (101) plane of rutile TiO2. These d-spacings were 0.33 nm and 0.25 nm, respectively.
In contrast, in the Sn/Ti oxide hybrid film prepared with Sn/Ti = 25:75 condition (Figure 6), the TEM measurement showed very weak crystallinity. This made it difficult to estimate the primary particle sizes because of the aggregation of the particles.
Figure 7 illustrates a device structure of PSCs employing Sn/Ti oxide films as the electron transport layer (ETL) evaluated in this work. In principle, the light absorption of the perovskite layer sandwiched between an ETL and the hole transporting layer (HTL) enables charge separation in PSCs. The photocarriers (e and h+) can be moved to each electrode to achieve photoenergy conversion. Here, our focus was centered on the ratio of Sn/Ti in the precursor solutions of terpineol/PEG base liquids that can optimize the light-to-electricity conversion efficiency of the device.
Figure 8 shows the I–V curve of the best-performing cell measured under one sunlight irradiation (100 mW/cm2). Additionally, Table 1 summarizes the I–V parameters. A maximum power conversion efficiency (PCE) of 18.1% was obtained for the solar cell fabricated using a sintering strategy (ramping time: 30 min, period of maximum temperature: 30 min) when the Sn/Ti was 25:75. Figure 9 presents the distributions of I–V parameters obtained from each solar cell consisting of the 10 different ETLs. The data clearly show that there is a correlation between Jsc and PCE. In other words, the enhanced Jsc led to a marked improvement in PCEs. Considering that the samples (Sn/Ti = 100:0, 75:25, 50:50) had higher crystallinity compared to those of Sn/Ti = 25:75, it is most likely that the more effective electron transport took place for the 25:75 film-based solar cells with low-crystallinity ETLs. In this regard, Dorman et al. reported the increased charge mobility of high-temperature processed Sn-doped TiO2, whereas charge carrier recombination can be decreased in solar cells [11]. Since the ionic radius of Sn(IV) is 69 p.m., which is larger than Ti(IV) (61 p.m.) [17], it is thought that Sn-doped TiO2 lowered the crystallinity, and the ETL had positive effects on the improved PCEs.
As shown in Figure 10a, the lead halide perovskite had a grain boundary even for the best-performing device. The approximate thickness of the ETL was 100 nm (Figure 10b). The cross-sectional SEM image in Figure 10b enabled us to observe the smooth surface of the thin film. In addition, the transmittance spectrum of the Sn/Ti oxide film in Figure 10c showed that the incident light can ensure efficient absorption in the perovskite layer in the device. The optimization of the deposition condition of the perovskite precursor solution and ETL thickness will further improve efficiencies by the smaller grain boundary formation of perovskite and the efficient electron generation and transport at the material interface. Most notably, a simple one-time deposition process of the Sn/Ti ETL using the mixed ethanol/terpineol/PEG base liquid made it possible to realize the respectable PCE.

4. Conclusions

Nanoparticle-based uniform Sn/Ti metal oxide films were successfully obtained via sintering through optimization of the composition of precursor solutions containing titanium and tin salts, which were dissolved in an ethanol/terpineol/PEG base liquid. It was suggested that PEG also had a key role in improving the homogeneity of the film during sintering.
We demonstrated that the maximum power conversion efficiency was obtained for the concentration condition of [SnCl2·2H2O]/[titanium tetraisopropoxide (TTIP)] = 25:75 in the terpineol/PEG-based precursor solution. The combination of XRD and high-resolution TEM analysis was a straightforward approach to understanding the microstructures related to the crystallinity and fusion of nanoparticles depending on the sintering strategies. It appears that the lower crystallinity for the 25:75 Sn/Ti oxide films has significant potential that boosts the power conversion efficiency of the device. Such fundamental information on semiconductor Sn/Ti oxide hybrid layers will be beneficial for the further development of cost-effective high-performance perovskite solar cells [18].

Author Contributions

Conceptualization, S.V. and K.M.; methodology, S.V. and K.M.; experiments and analysis, S.V., N.H., S.H., T.S. and K.M., writing—original draft preparation, S.V., V.K.B., S.J., P.K.B. and K.M.; writing—review and editing, S.V. and K.M.; project administration; S.V. and K.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Caterpillar Fellowship, Research Excellence Program in Bradley University, and Koshiyama Research Grant, Japan.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bairy, R.; Vijeth, H.; Kulkarni, S.D.; Murari, M.-S.; Bhat, U.-K. Improvement of third-order NLO properties of vacuum deposited Cd1-xPbxS nanostructured thin films for optoelectronic device applications. Mater. Res. Bull. 2023, 161, 112146. [Google Scholar] [CrossRef]
  2. Suwannakham, N.; Tubtimtae, A. Synthesis of aluminum telluride thin films via using electrodeposition method for solar absorber applications. Mater. Lett. 2023, 336, 133920. [Google Scholar] [CrossRef]
  3. Yu, Z.; Wei, X.; Zheng, Y.; Hui, H.; Bian, M.; Dhole, S.; Seo, J.-H.; Sun, Y.Y.; Jia, Q.; Zhang, S.; et al. Chalcogenide perovskite BaZrS3 thin-film electronic and optoelectronic devices by low temperature processing. Nano Energy 2021, 85, 105959. [Google Scholar] [CrossRef]
  4. Jena, A.-K.; Kulkarni, A.; Miyasaka, T. Halide perovskite photovoltaics: Background, status, and future prospects. Chem. Rev. 2019, 119, 3036–3103. [Google Scholar] [CrossRef] [PubMed]
  5. Zhao, Y.; Ma, F.; Qu, Z.; Yu, S.; Shen, T.; Deng, H.-X.; Chu, X.; Peng, X.; Yuan, Y.; Zhang, X.; et al. Inactive (PbI2)2RbCl stabilizes perovskite films for efficient solar cells. Science 2022, 377, 531–534. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, P.; Li, M.; Chen, W.-C. A Perspective on Perovskite Solar Cells: Emergence, Progress, and Commercialization. Front. Chem. 2022, 10, 802890. [Google Scholar] [CrossRef] [PubMed]
  7. Zhang, X.; Song, W.-L.; Tu, J.; Wang, J.; Wang, M.; Jiao, S. A Review of Integrated Systems Based on Perovskite Solar Cells and Energy Storage Units: Fundamental, Progresses, Challenges, and Perspectives. Adv. Sci. 2021, 8, 2100552. [Google Scholar] [CrossRef] [PubMed]
  8. Guo, H.; Zhang, H.; Yang, J.; Chen, H.; Li, Y.; Wang, L.; Niu, X. TiO2/SnO2 Nanocomposites as Electron Transporting Layer for Efficiency Enhancement in Planar CH3NH3PbI3-Based Perovskite Solar Cells. ACS Appl. Energy Mater. 2018, 1, 6936–6944. [Google Scholar] [CrossRef]
  9. Hu, M.; Zhang, L.; She, S.; Wu, J.; Zhou, X.; Li, X.; Wang, D.; Miao, J.; Mi, G.; Chen, H.; et al. Electron Transporting Bilayer of SnO2 and TiO2 Nanocolloid Enables Highly Efficient Planar Perovskite Solar Cells. Solar RRL 2019, 4, 1900331. [Google Scholar] [CrossRef]
  10. Abuhelaiqa, M.; Shibayama, N.; Gao, X.-X.; Kanda, H.; Nazeeruddin, M.K. SnO2/TiO2 Electron Transporting Bilayers: A Route to Light Stable Perovskite Solar Cells. ACS Appl. Energy Mater. 2021, 4, 3424–3430. [Google Scholar] [CrossRef]
  11. Cai, Q.; Zhang, Y.; Liang, C.; Li, P.; Gu, H.; Liu, X.; Wang, J.; Shentu, Z.; Fan, J.; Shao, G. Enhancing efficiency of planar structure perovskite solar cells using Sn-doped TiO2 as electron transport layer at low temperature. Electrochim. Acta 2018, 261, 227–235. [Google Scholar] [CrossRef]
  12. Liao, Y.-H.; Chang, Y.-H.; Lin, T.-H.; Chan, S.-H.; Lee, K.-M.; Hsu, K.-H.; Hsu, J.-F.; Wu, M.-C. Boosting the power conversion efficiency of perovskite solar cells based on Sn doped TiO2 electron extraction layer via modification the TiO2 phase junction. Sol. Energy 2020, 205, 390–398. [Google Scholar] [CrossRef]
  13. Su, T.-S.; Wei, T.-C. Co-Electrodeposition of Sn-Doped TiO2 ElectronTransporting Layer for Perovskite Solar Cells Phys. Status Solidi A 2020, 217, 1900491. [Google Scholar] [CrossRef]
  14. Suo, J.; Yang, B.; Mosconi, E.; Choi, H.-S.; Kim, Y.; Sakeeruddin, S.M.; Angelis, F.-D.; Grätzel, M.; Kim, H.-S.; Hagfeldt, A. Surface Reconstruction Engineering with Synergistic Effect of Mixed-Salt Passivation Treatment toward Efficient and Stable Perovskite Solar Cells. Adv. Funct. Mater. 2021, 31, 2102902. [Google Scholar] [CrossRef]
  15. Manseki, K.; Yamasaki, M.; Yoshida, M.; Sugiura, T.; Vafaei, S. Optimization of Thermal Evaporation Process of Gold Deposition for Perovskite Solar Cells. In Proceedings of the Thermal and Fluids Engineering Summer Conference, New Orleans, LA, USA, 5–8 April 2020; pp. 739–745. [Google Scholar]
  16. Ito, S.; Murakami, T.-N.; Comte, P.; Liska, P.; Grätzel, C.; Nazeeruddin, M.K.; Grätzel, M. Fabrication of thin film dye sensitized solar cells with solar to electric power conversion efficiency over 10%. Thin Solid Films 2008, 516, 4613–4619. [Google Scholar] [CrossRef]
  17. Duan, Y.; Fu, N.; Liu, Q.; Fang, Y.; Zhou, X.; Zhang, J.; Lin, Y. Sn-doped TiO2 photoanode for dye-sensitized solar cells. J. Phys. Chem. C 2012, 116, 8888–8893. [Google Scholar] [CrossRef]
  18. Čulík, P.; Brooks, K.; Momblona, C.; Adams, M.; Maréchal, F.; Dyson, P.-J.; Nazeeruddin, M.-K. Design and Cost Analysis of 100 MW Perovskite Solar Panel Manufacturing Process in Different Locations. ACS Energy Lett. 2022, 7, 3039–3044. [Google Scholar] [CrossRef]
Figure 1. Sintering strategies of the Sn/Ti oxide thin-film samples in air. The periods of maximum temperature at 500 °C were 2 h for (a) and 30 min for (b), respectively. From 500 °C, the samples were cooled down to room temperature naturally.
Figure 1. Sintering strategies of the Sn/Ti oxide thin-film samples in air. The periods of maximum temperature at 500 °C were 2 h for (a) and 30 min for (b), respectively. From 500 °C, the samples were cooled down to room temperature naturally.
Materials 16 03136 g001
Figure 2. Photos of several Sn/Ti oxides-deposited FTO glass substrates after heat treatment at 500 °C. Samples (ac) were prepared by using a precursor solution of tin and titanium compounds dissolved in a mixed ethanol/terpineol base liquid. The ratios of tin and titanium ions (Sn/Ti) in the precursor were 25:75 for (a), 50:50 for (b), and 100:0 for (c), respectively. Samples (df) were prepared via the same sintering strategy as those of (ac). An ethanol base liquid without terpineol was employed for the preparation of samples (df). Similarly, the ratios of Sn/Ti in the precursor were 25:75 for (d), 50:50 for (e), and 100:0 for (f), respectively. For samples (ac), the right parts highlighted in a dotted yellow line were not covered with Sn/Ti oxide thin films for fabricating perovskite solar cells. The three samples (df) showed the formation of agglomerated Sn/Ti oxides on top of the FTO substrates, whereas uniform films formed for (ac) through the optimization of the composition of precursor solutions.
Figure 2. Photos of several Sn/Ti oxides-deposited FTO glass substrates after heat treatment at 500 °C. Samples (ac) were prepared by using a precursor solution of tin and titanium compounds dissolved in a mixed ethanol/terpineol base liquid. The ratios of tin and titanium ions (Sn/Ti) in the precursor were 25:75 for (a), 50:50 for (b), and 100:0 for (c), respectively. Samples (df) were prepared via the same sintering strategy as those of (ac). An ethanol base liquid without terpineol was employed for the preparation of samples (df). Similarly, the ratios of Sn/Ti in the precursor were 25:75 for (d), 50:50 for (e), and 100:0 for (f), respectively. For samples (ac), the right parts highlighted in a dotted yellow line were not covered with Sn/Ti oxide thin films for fabricating perovskite solar cells. The three samples (df) showed the formation of agglomerated Sn/Ti oxides on top of the FTO substrates, whereas uniform films formed for (ac) through the optimization of the composition of precursor solutions.
Materials 16 03136 g002
Figure 3. XRD pattern of the sintered Sn/Ti oxide samples at 500 °C in air. The ratios of Sn/Ti in the precursor were 100:0, 75:25, 50:50, 25:75, and 0:100. The database of SnO2 (blue, JCPDS: 01-077-0452), rutile TiO2 (red, JCPDS: 01-076-0324), and anatase TiO2 (blue, JCPDS: 01-075-2550) is also presented. The period of maximum temperature in the sintering was 30 min and 2 h.
Figure 3. XRD pattern of the sintered Sn/Ti oxide samples at 500 °C in air. The ratios of Sn/Ti in the precursor were 100:0, 75:25, 50:50, 25:75, and 0:100. The database of SnO2 (blue, JCPDS: 01-077-0452), rutile TiO2 (red, JCPDS: 01-076-0324), and anatase TiO2 (blue, JCPDS: 01-075-2550) is also presented. The period of maximum temperature in the sintering was 30 min and 2 h.
Materials 16 03136 g003
Figure 4. TEM images of the Sn/Ti oxide sample treated at 500 °C in air. The ratio of Sn/Ti in the precursor solution was 50:50. The period of maximum temperature in the sintering was 2 h. (a,b) Lower-resolution images of the Sn/Ti oxide sample. Inserted image in (a) presents the Selected Area Diffraction (SAD) pattern corresponding to the observed area of interest highlighted with a dotted orange circle. (c,d) High-resolution TEM images (HR-TEM) measured at two different positions. Figure (c) shows a d-spacing for the observed lattice fringes and lattice planes corresponding to (110) and (101) of SnO2, and (110) of TiO2.
Figure 4. TEM images of the Sn/Ti oxide sample treated at 500 °C in air. The ratio of Sn/Ti in the precursor solution was 50:50. The period of maximum temperature in the sintering was 2 h. (a,b) Lower-resolution images of the Sn/Ti oxide sample. Inserted image in (a) presents the Selected Area Diffraction (SAD) pattern corresponding to the observed area of interest highlighted with a dotted orange circle. (c,d) High-resolution TEM images (HR-TEM) measured at two different positions. Figure (c) shows a d-spacing for the observed lattice fringes and lattice planes corresponding to (110) and (101) of SnO2, and (110) of TiO2.
Materials 16 03136 g004aMaterials 16 03136 g004bMaterials 16 03136 g004c
Figure 5. TEM images of the Sn/Ti oxide sample treated at 500 °C in air. The ratio of Sn/Ti in the precursor solution was 50:50. The period of maximum temperature in the sintering was 30 min. (a) Lower-resolution image of the Sn/Ti oxide sample. Inserted image in (a) presents the SAD pattern. (b,c) HR-TEM images measured at two different positions. (b) shows a d-spacing for the observed lattice fringes and lattice planes corresponding to (110) of SnO2 and (101) of TiO2.
Figure 5. TEM images of the Sn/Ti oxide sample treated at 500 °C in air. The ratio of Sn/Ti in the precursor solution was 50:50. The period of maximum temperature in the sintering was 30 min. (a) Lower-resolution image of the Sn/Ti oxide sample. Inserted image in (a) presents the SAD pattern. (b,c) HR-TEM images measured at two different positions. (b) shows a d-spacing for the observed lattice fringes and lattice planes corresponding to (110) of SnO2 and (101) of TiO2.
Materials 16 03136 g005aMaterials 16 03136 g005b
Figure 6. A TEM image of the Sn/Ti oxide sample treated at 500 °C in air (the ratio of Sn/Ti in the precursor solution was 25:75). The period of maximum temperature in the sintering was 30 min.
Figure 6. A TEM image of the Sn/Ti oxide sample treated at 500 °C in air (the ratio of Sn/Ti in the precursor solution was 25:75). The period of maximum temperature in the sintering was 30 min.
Materials 16 03136 g006
Figure 7. A device structure of the lead halide PSCs tested in this work. The ETL consists of a Sn/Ti oxide thin film (single layer) deposited on top of the FTO glass.
Figure 7. A device structure of the lead halide PSCs tested in this work. The ETL consists of a Sn/Ti oxide thin film (single layer) deposited on top of the FTO glass.
Materials 16 03136 g007
Figure 8. I–V curves of the best-performing devices by Sn/Ti oxide ETLs measured under simulated sunlight. The period of maximum temperature in the sintering of ETL was 30 min.
Figure 8. I–V curves of the best-performing devices by Sn/Ti oxide ETLs measured under simulated sunlight. The period of maximum temperature in the sintering of ETL was 30 min.
Materials 16 03136 g008
Figure 9. Distributions of I–V parameters of (a) Voc, (b) Jsc, (c) FF, and (d) PCE. All the data were obtained from each solar cell fabricated using the 10 different ETLs.
Figure 9. Distributions of I–V parameters of (a) Voc, (b) Jsc, (c) FF, and (d) PCE. All the data were obtained from each solar cell fabricated using the 10 different ETLs.
Materials 16 03136 g009aMaterials 16 03136 g009b
Figure 10. Cross-sectional SEM images of the best-performing device: (a) low-magnification image and (b) high-magnification image presenting the ETL part; (c) transmittance spectrum of the ETL thin film, which is the same sample as (b).
Figure 10. Cross-sectional SEM images of the best-performing device: (a) low-magnification image and (b) high-magnification image presenting the ETL part; (c) transmittance spectrum of the ETL thin film, which is the same sample as (b).
Materials 16 03136 g010aMaterials 16 03136 g010b
Table 1. I–V parameters obtained for the best-performing perovskite solar cell fabricated using the 10 ETLs with different ratios of Sn/Ti and sintering times.
Table 1. I–V parameters obtained for the best-performing perovskite solar cell fabricated using the 10 ETLs with different ratios of Sn/Ti and sintering times.
Sn:TiTimeVoc (mV)Jsc (mA/cm2)FFPCE (%)
100:030 min643 25.47 0.571 9.3
2 h646 23.12 0.508 7.6
75:2530 min939 15.04 0.620 8.8
2h922 13.90 0.548 7.0
50:5030min997 18.29 0.686 12.5
2 h1006 25.77 0.469 12.2
25:7530 min1063 25.32 0.671 18.1
2 h1057 25.90 0.657 18.0
0:10030 min1078 24.66 0.628 16.7
2 h1068 25.10 0.601 16.1
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

Vafaei, S.; Boddu, V.K.; Jala, S.; Bezawada, P.K.; Hattori, N.; Higashi, S.; Sugiura, T.; Manseki, K. Preparation of Nanostructured Sn/Ti Oxide Hybrid Films with Terpineol/PEG-Based Nanofluids: Perovskite Solar Cell Applications. Materials 2023, 16, 3136. https://doi.org/10.3390/ma16083136

AMA Style

Vafaei S, Boddu VK, Jala S, Bezawada PK, Hattori N, Higashi S, Sugiura T, Manseki K. Preparation of Nanostructured Sn/Ti Oxide Hybrid Films with Terpineol/PEG-Based Nanofluids: Perovskite Solar Cell Applications. Materials. 2023; 16(8):3136. https://doi.org/10.3390/ma16083136

Chicago/Turabian Style

Vafaei, Saeid, Vamsi Krishna Boddu, Stephen Jala, Pavan Kumar Bezawada, Nagisa Hattori, Seiho Higashi, Takashi Sugiura, and Kazuhiro Manseki. 2023. "Preparation of Nanostructured Sn/Ti Oxide Hybrid Films with Terpineol/PEG-Based Nanofluids: Perovskite Solar Cell Applications" Materials 16, no. 8: 3136. https://doi.org/10.3390/ma16083136

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