Next Article in Journal
A Novel Remaining Useful Life Prediction Method for Hydrogen Fuel Cells Based on the Gated Recurrent Unit Neural Network
Next Article in Special Issue
Nanostructured Top Contact as an Alternative to Transparent Conductive Oxides in Tandem Perovskite/c-Si Solar Cells
Previous Article in Journal
Glucose Isomerase: Functions, Structures, and Applications
Previous Article in Special Issue
Evaluation of Active Layer Thickness Influence in Long-Term Stability and Degradation Mechanisms in CsFAPbIBr Perovskite Solar Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Performance-Enhancing Sulfur-Doped TiO2 Photoanodes for Perovskite Solar Cells

by
Muhazri Abd Mutalib
1,
Norasikin Ahmad Ludin
1,*,
Mohd Sukor Su’ait
1,
Matthew Davies
2,3,
Suhaila Sepeai
1,
Mohd Asri Mat Teridi
1,
Mohamad Firdaus Mohamad Noh
1 and
Mohd Adib Ibrahim
1
1
Solar Energy Research Institute (SERI), Universiti Kebangsaan Malaysia (UKM), Bangi 43600, Selangor, Malaysia
2
SPECIFIC IKC, Materials Research Centre, College of Engineering, Swansea University Bay Campus, Fabian Way Institution, Swansea SA1 8EN, UK
3
School of Chemistry and Physics, University of KwaZulu-Natal, Durban 4041, South Africa
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(1), 429; https://doi.org/10.3390/app12010429
Submission received: 30 October 2021 / Revised: 19 November 2021 / Accepted: 2 December 2021 / Published: 3 January 2022
(This article belongs to the Special Issue Novel Organic-Inorganic Photovoltaic Materials)

Abstract

:
High-performance electron transport layer (ETL) anode generally needs to form a uniform dense layer with suitable conduction band position and good electron transport properties. The TiO2 photoanode is primarily applied as the ETL because it is low-cost, has diverse thin-film preparation methods and has good chemical stability. However, pure TiO2 is not an ideal ETL because it lacks several important criteria, such as low conductivity and conduction band mismatch with compositional-tailored perovskite. Thus, TiO2 is an inefficient photo-anode or ETL for high-performance perovskite devices. In this study, sulfur as dopant in the TiO2 photo-anode thin film is used to fabricate solid-state planar perovskite solar cells in relatively high humidity (40–50%). The deposited S-doped thin film improves the power conversion efficiency (PCE) of the device to 6.0%, with the un-doped TiO2 producing a PCE of 5.1% in the best device. Improvement in PCE is due to lower recombination and higher photocurrent density, resulting in 18% increase in PCE (5.1–6.0%).

1. Introduction

Perovskite solar cells (PSCs) have been attracting great attention in the past decade due to their rise in power conversion efficiencies (PCE), with certified efficiencies now greater that 25% [1,2,3,4,5]. A typical PSC consists of multiple layers of solid thin films, which include an electron transport layer (ETL), perovskite absorber layer and hole transport layer (HTL). These layers are aligned in the heterojunction according to its distinct device configurations. Regardless of the device configurations, an optimised ETL can potentially improve the overall behaviour of the PSCs in terms of the efficiency, hysteresis management and reproducibility of the device [6,7]. The functions of an ETL are closely related to its photoelectrochemical and structural properties by effectively facilitating electron collection and transfer from the photon-absorbing perovskite thin film. Good indications of a high-performance ETL or photo-anode include a thin, dense and suitable conduction band and good electron transfer. These characteristics also aim to minimise the interfacial recombination, facilitate electron movement and manage charge accumulation. The TiO2 photo-anode is primarily applied as an ETL because it is low-cost, has numerous thin film preparation methods and has suitable chemical stability [8,9,10]. However, pure TiO2 is not an ideal ETL because it lacks several important criteria such as low conductivity and conduction band mismatch with compositional-tailored perovskite (e.g., triple-cation perovskite system). Hence, there is room for improvement in the performance of TiO2 as the ETL [1].
Metal doping of TiO2 has become a practical technique to modify the band position and improve the thin film conductivity to fabricate photo-anodes with high charge extraction capacity. Dopants, such as yttria, niobium, magnesium and tin, have been proven to enhance the overall PSC efficiency due to the enriched layer conductivity and high electron mobility [11,12,13,14]. Although many dopants have been used to modify the TiO2 photoanode for PSCs, the application of sulfur as dopant for these cells has not been reported yet. S-doped TiO2 has been demonstrated to enhance photocurrent, light absorbance and photocatalytic activity and modify the band gap energy in other applications, such as coatings and photo-catalysis [15,16]. These properties could improve the characteristics of the photo-anode thin films for perovskite solar devices.
In this study, sulfur was used for the first time as dopant in the TiO2 photo-anode thin film in solid-state planar heterojunction solar cells. Methylammonium lead iodide (CH3NH3PbI3) was used as the light absorber and spiro-OMeTAD as HTL. The effects of charge extraction and blocking of holes from the TiO2 mesoporous layer was eliminated by choosing a planar PSC architecture. A sol-gel solution containing sulfur and titanium salts was spin coated onto the FTO glass to form the S-doped TiO2 photo-anode thin film. The doping concentration of sulfur is tuned by controlling the sulfur salt content in the sol-gel solution.

2. Materials and Methods

Titanium diisopropoxide bis(acetylacetonate) (75 wt% in isopropanol), thiourea (≥99.0%), absolute ethanol (99.95%), acetylacetone (≥99%), lead iodide (PbI2, 99%), lithium bis(trifluoromethanesulfonyl)imide (Li-TFSI, 99%), 4-tert-butylpyridine (tBP, 96%), N,N-dimethylformamide (DMF, 99.8%), 2,20,7,70-tetrakis-(N,N-dip-methoxyphenylamine)-9,90-spirobifluorene (spiro-OMeTAD, SHT-263, Solarpur), tris(2-(1H-pyrazol-1-yl)-4-tert-butylpyridine) cobalt(III) tri[bis(trifluoromethane)sulfonimide] (FK209 Co(III) TFSI, 98%) and anhydrous acetonitrile (99.8%) were obtained from Sigma Aldrich and used as received. Prior to the deposition of photoanode layer, FTO-coated glass (FTO, TEC15, 15 Ω/sq, Greatcell Solar Materials) was first etched to design the desired pattern of electrode using zinc powder and 3M HCL. Next, the FTO substrate is cleaned using ultrasonic bath for 10 min, in the following order: soap (2% Hellmanex in water), acetone then propanol (99.5%), to remove the organic residue on the surface of FTO-glass substrate. In this modified sol-gel method, S-doped TiO2 photoanode is prepared by first mixing 1.2 mL acetyl acetone, 7 mL TTIP, 35 mL absolute ethanol and calculated amounts of thiourea as sulfur source. To evaluate to effect of sulfur doping concentration, the amount of thiourea is prepared with 5%, 10% and 15% mol thiourea compared to TTIP. The precursor solution was then magnetically stirred for 180 min to produce a clear and homogeneous solution. A blank precursor solution was also prepared without the thiourea as control. The deposition of photoanode was carried out using spin coating method. 3 layers of TiO2 photoanode was spin-coated at 3000 rpm with drying process at 120 °C for 10 min after each layer deposition. Afterwards, the TiO2 photoanode was sintered at 450 °C to complete the deposition. The nomination of all sample relates to the sulfur doping content of the initial precursor solution.
To investigate the optical properties of the photoanode thin film, room temperature absorbance measurement (300 to 700 nm) is recorded using a UV-Vis spectrophotometer (Lambda 35, PerkinElmer). The XRD diffractogram of the photoanode thin film were analysed using X-ray diffractometer model Bruker D8 Advance (2θ, 10–70°). The FTIR spectrum of TiO2 photoanodes were evaluated using FTIR spectrometer (Shimadzu—840 s). Fluorescence spectroscopy was used to assess the charge recombination behaviours of TiO2 photoanodes. The fluorescence spectroscopy analysis was carried out using a Perkin Elmer Luminescence spectrometer (LS 55) with a thin film holder accessory, at room temperature and 300 nm as excitation wavelength. Morphological study of the TiO2 photoanodes were carried out using field-emission scanning electron microscopy (FESEM) were analysed using FE-SEM SUPRA VP55 equipped with an energy dispersive X-ray (EDX) spectroscopy for elemental detection and mapping. The photoelectrochemical (PEC) properties of the TiO2 photoanodes were determined via linear sweep voltammetry (LSV) using an electrochemical workstation (Metrohm Autolab), simulated AM 1.5 at a calibrated intensity 100 mWcm−2 at NTP conditions. TiO2 samples as the working electrode is immersed in Na2SO4 solution (0.5 M), Pt as counter electrode and Ag/AgCl as the reference electrode. The scanning rate is kept at 20 mVs−1 (−0.4 to 1.4 V vs. Ag/AgCl).
The perovskite solar devices were prepared by depositing the perovskite absorber layer, hole transporting layer and silver contact, consecutively. The perovskite absorber layer is deposited using a 2-step method where in the first layer, 100 µL of PbI2 solution consisting of PbI2 powder (507 mg) in DMF (1 mL) and tBP (100 µL) is spin coated on the TiO2 photoanode at 3000 rpm for 30 s then annealed at 100°C for 10 min. Previous study have emphasised the addition of tBP to enhance the hydrophobicity of PbI2 and the deposited perovskite absorber [17]. Then, 250 µL of MAI solution consisting of 35 mg MAI powder in 1 mL isopropanol at 4000 rpm for 30 s then annealed at 100 °C for 30 min. After allowing the temperature to cool down to room temperature, hole transport layer solution is then spin coated at 4000 rpm for 30 s. The hole transport layer solution consisted of spiro-OMeTAD powder (72.3 mg), chlorobenzene (1 mL), 4-tert butylpyridine (28.8 μL), Li-TFSI solution (17.5 μL, 520 mg/mL in acetonitrile) and FK209 solution (29.0 μL, 520 mg/mL in acetonitrile). All of the perovskite solar devices layer solutions were prepared and spin coated in RH40-50%. To finish the perovskite solar devices, Ag counter electrode was thermal evaporated onto the hole transport layer under high vacuum atmosphere with 0.07 cm2 active area. The photocurrent-voltage (I-V) characteristics were measured by applying a reverse scan at a rate of 0.1 V/s in a Keithley 2400 source meter under AM 1.5 G solar illumination. The average values of PSC devices were measured over 10 samples.

3. Results

3.1. UV-Visible Spectroscopy

Given the boiling temperature of elemental sulfur at 444.6 °C, evaluating the behaviour of the S-doped TiO2 thin film towards the change in sintering temperature is interesting. Figure 1a shows the UV-vis absorption spectra of various TiO2 thin films sintered at 450 °C and 500 °C. The sintering temperatures were selected based on the temperature required to produce highly-crystalline TiO2 thin films, that is, at least 450 °C [18]. 550 °C is not suitable due to the formation of rutile TiO2 at this point, and this type of TiO2 is less photo-active than anatase TiO2 [19]. Firstly, all TiO2 samples exhibit the typical absorbance behaviour of TiO2 materials, in which the absorbance is low at higher wavelength and gradually increases at the UV region. Compared with the un-doped TiO2 photo-anode, the S-doped TiO2 displayed a small red shift of the absorption edge due to the new energy levels within the band gap caused by the doping of S. This phenomenon has been reported in a previous study [20,21]. Based on a previous study, the mechanism of TiO2 band gap reduction could be attributed to the interaction between the S dopant (d and p orbitals) with the TiO2 energy levels [22]. The bandgaps were reduced from 3.56 eV (un-doped TiO2 photoanode) to 3.41 eV (10% S-doped TiO2 photoanode) as estimated from the Tauc plots in Figure 1b. When the sintering temperature was increased to 500 °C, the absorption spectrum of the S-doped TiO2 photoanode was reduced. This phenomenon suggested the S dopant has diffused out of the thin film, which was also found in a previous study [15]. Thus, to produce S-doped TiO2 with anatase phase, 450 °C was selected as the sintering temperature.
The different doping levels of S were investigated using UV-vis spectroscopy. Figure 1b shows the absorption response of all S-doped TiO2 photo-anodes with various concentrations of S. With 5% S dopant, the absorption bands were improved compared with those of the un-doped TiO2 photo-anode. According to the plots, 10% S-doped TiO2 showed the highest light absorbance. Thus, 10% was selected as the optimum concentration for sulfur doping.

3.2. X-ray Diffraction (XRD) Analysis

Changes in the crystalline phase of the S-TiO2 photo-anode thin film were analysed using X-ray diffraction. The X-ray diffractograms of the un-doped and S-doped TiO2 photo-anode thin film sintered at 450 °C are shown in Figure 2. Each sample had defined TiO2 diffraction peaks, which were correlated with the characteristics of nano-crystalline anatase TiO2 peaks with several FTO diffraction peaks. According to a previous study, the characteristic peaks of anatase TiO2 are found at the angles of 27.0°, 38.2°, 55.1° and 62.0° corresponding to the (1 0 1), (1 0 4), (1 0 5) and (2 0 4) planes [23]. Rutile-phase TiO2 was not found in the sintered TiO2 photo-anode thin-film sample, which could be related to the sintering temperature of 450 °C. The sintering temperature is an important factor in producing the S-doped TiO2 photo-anode thin film because a higher temperature could lead to the formation of un-desired, less photo-active rutile phase and could reduce the incorporation of S (boiling point: 444.6 °C) in the thin-film lattice [15]. The XRD diffractogram of the S-doped TiO2 was compared with that of pure TiO2, and the two diffraction patterns were almost similar with no S characteristic peak. This observation indicated that the S ions had entered the lattice structure of TiO2 or the concentration of the S compounds were lower than the detection limit of the equipment, as observed in a previous study [24]. In addition to the diffraction pattern, the data in the XRD pattern could be used to calculate the average particle size by using the Scherrer equation as follows:
D hkl = 0.89 λ β   cos θ
where λ is the wavelength of the Cu Kα laser used (0.1541 nm), θ is the diffraction angle of the peak and β refers to the full width of the peak measured at half maximum intensity (FWHM) [25]. Thus, the average particle sizes of both samples were calculated according to Equation (1), resulting in 23.8 nm and 18.0 nm for the un-doped and S-doped TiO2 photo-anode, respectively. Thus, S-doping had a direct effect on the average particle size of the sintered TiO2 photo-anode. A previous study has reported that doping via substitution of a lattice atom with a dopant would reduce the grain size compared with the un-doped sample, as observed in this study [26].

3.3. Fourier-Transform Infrared (FTIR) Spectroscopy

The FTIR spectrum can be an effective tool to detect the functional groups in the TiO2 thin films. Figure 3a shows the FTIR spectra of the S-doped and un-doped TiO2 thin films. The broad absorption peak in the region between 1650 cm−1 and 3450 cm−1 could be associated to the O-H vibrations of the absorbed water molecules on the TiO2 thin film surface (Figure 3a) [24,27]. Most importantly, the drop in the intensity in this region for the S-doped TiO2 thin film would probably be attributed to the hydrophobicity of S. Close inspection of Figure 3b shows that the S distinct peak was absent in the un-doped TiO2 sample but found in the S-doped TiO2 sample. In a previous study, the distinct peak at 1086 cm−1 and 1242cm−1 corresponded to the characteristic of bidentate SO 4 2 co-ordination with metals, such as Ti4+, and the formation of Ti-O-S bonds, respectively [22]. The presence of thiourea was negligible due to the absence of the C=O bond absorption bands. Thus, the FTIR spectra had proven the formation of S-doped TiO2 thin film by using the precursor solution and sintering at 450 °C to form the Ti-O-S bonds.

3.4. Fluorescence Spectroscopy

Fluorescence emission spectroscopy is important to examine the ability of charge carrier trapping, migration and transfer and study the behaviour of electron/hole pairs in the semiconductor particles, such as TiO2 [28]. To elucidate the effects of S-doping on the recombination of the photo-generated electron/hole pairs produced in the TiO2 thin film, the fluorescence spectra were examined for the S-doped and un-doped TiO2. Figure 4 illustrates the intensity of the fluorescence emission from the electron/hole recombination in the sample. The S-doped TiO2 thin film emits lower fluorescence intensity, which strongly suggested a lower radiative charge recombination and thus higher photo-catalytic efficiency [28]. Additionally, the produced fluorescence spectra showed similar patterns of peaks with only distinct intensities, which suggests that the addition of S in the TiO2 lattice did not change the TiO2 photo-catalytic mechanism.

3.5. Field Emission Scanning Electron Microscope (FESEM) Analysis

The effect of sulfur doping on the microstructure of the TiO2 thin film photo-anode was assessed using the FESEM micrograph. Figure 5a,b indicate the FESEM morphological images of the un-doped and S-doped TiO2 photo-anode thin films sintered at 450 °C, respectively. High-magnification images of the TiO2 thin film revealed the nano-crystallisation of the deposited thin film, as suggested in the XRD diffractograms in Figure 2. Additionally, the thin film showed a dense and uniform morphology of homogenous granular and spherical grain. As mentioned in a previous study, the incorporation of S in the TiO2 lattice would not contribute to a major change in the morphology of the thin film [22]. The EDX elemental spectra were recorded for the un-doped and S-doped TiO2 thin films to observe the elements, as shown in Figure 5c,d, respectively. The EDX spectra for the S-doped TiO2 thin film featured a minor peak for S at approximately 2.3 keV [24]. Additionally, the Ti peaks were observed at approximately 0.2 keV and a strong peak at 4.5 keV, which were attributed to the surface and bulk TiO2, respectively [29]. Figure 5e,f show the EDX elemental mapping of Ti, O and S in the un-doped and S-doped TiO2 photo-anode thin films, respectively. The mapping images suggested the highly distributed behaviour of S ions throughout the S-doped TiO2 thin film and the absence of S ions in the un-doped TiO2 sample. Additionally, the presence of carbon in the S-doped TiO2 photo-anode thin film most likely due to the usage of thiourea in the preparation of the thin film, as also described in previous study [30].

3.6. Linear Sweep Voltammetry (LSV) and Photocurrent–Voltage (J–V) Curve Studies

The photo-current response of the undoped and S-doped TiO2 ETLs fabricated by spin-coating and sintered at 450 °C were evaluated using linear sweep voltammetry. The measurements were performed in the dark and under visible light-irradiated environment. The samples were soaked in 0.5 M Na2SO4 solution at a scan rate of 20 mV/s, and the results are displayed in Figure 6a. At 1.0 V, the photo-current densities of the un-doped and S-doped samples were 0.0611 mA/cm2 and 0.0695 mA/cm2 (13.75% increase), respectively. Hence, the incorporation of S in the TiO2 lattice increased the number of charge carriers (electron/hole pairs) in the photo-catalytic reaction of the thin film [23]. A previous study has reported similar responses, in which an increase in the photo-current to a certain extent was observed when the S content was increased, and a higher loading would be detrimental towards the photo-current readings due to the charge recombination of the electron-hole pair [31]. Thus, the incorporation of S as dopant could serve as a beneficial component to increase the photo-current in the thin film.
Perovskite solar devices with different photo-anodes were fabricated in high-humidity environment as described in the experimental procedure. Figure 6b (inset) shows the schematic diagram of the fabricated planar structured PSC in a relatively high humidity (40–50%). The photocurrent–voltage (J–V) characteristic curves were observed under 100 mW cm−2 (1 sun) illumination. Figure 6b shows the J–V curve of the heterojunction perovskite solar devices based on the FTO/S-doped TiO2 or TiO2/CH3NH3PbI3/spiro-MeOTAD/Ag for the best device. The un-doped TiO2 sample (reference device) produced a short-circuit density (JSC) of 13.1 mA cm−2, open circuit voltage of 0.982mV, fill factor of 39.88% and PCE of 5.1%. By contrast, the S-doped perovskite device was improved with device performance of JSC of 13.9 mA cm−2, open circuit voltage of 0.997 mV, fill factor of 43.18% and PCE of 6.0%. The values of best device and device average over 10 made samples are summarised in Table 1. The device equipped with S-doped TiO2 had better JSC, FF and PCE compared with the un-doped TiO2, but the VOC was relatively the same. The improvement in the JSC of the S-doped perovskite device was attributed to the increase in the electron collection capacity of the charge carrier and photo-current density, as described by Figure 6a, as reported in a previous study [1]. Additionally, Table 1 shows that the charge transfer resistance (RCT) in the S-doped TiO2 perovskite device was lower than that of the un-doped TiO2 device. The lower resistance could facilitate higher electron transfer in the perovskite device, resulting in the higher overall PCE.

4. Conclusions

In summary, the application of S-doped TiO2 photo-anode as the ETL for planar PSC had been demonstrated. The optimized sintering temperature was 450 °C. The optimized sulfur dopant was 10% because a higher loading induced a drop in the absorbance capacity of the photo-anode. The S as dopant was proved to be present in the TiO2 photo-anode layer after the deposition of S containing the precursor solution. The S-doped TiO2 photo-anode was superior to the un-doped TiO2 photo-anode in terms of higher absorbance, photo-catalytic activity and photo-current density (13.75% increase). Perovskite solar devices with S-doped TiO2 also showed better efficiency (18% increase) fabricated under relative humidity of 40–50%. The increase in the electron collection capacity of the carrier charge and photo-current density enhanced the PCE at 6.0%. Additional studies on the relation between hydrophobicity of ETL and stability of the PSC device should be performed. This work may facilitate the tailoring of the TiO2 photo-anode to match the energy levels of an ETL with the perovskite absorber layer.

Author Contributions

Conceptualization, M.A.M. and N.A.L.; methodology, M.A.M.; software, M.A.M. and N.A.L.; validation, N.A.L., M.D. and M.A.I.; formal analysis, M.A.M.; investigation, M.A.M.; resources, M.F.M.N., M.A.M.T. and N.A.L.; data curation, M.A.M.; writing—original draft preparation, M.A.M. and N.A.L.; writing—review and editing, N.A.L., M.S.S., M.D. and S.S.; visualization, M.A.M., N.A.L. and M.D.; supervision, N.A.L.; project administration, N.A.L.; funding acquisition, N.A.L. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the financial support from the Universiti Kebangsaan Malaysia (UKM) under the Dana Impak Perdana (DIP-2019-025). The first author also wants to acknowledge the UKM for the PhD scholarship under the Skim Zamalah Yayasan Canselor.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors also appreciate the technical and management support from the Centre for Research and Instrumentation (CRIM), UKM.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Su, T.S.; Wei, T.C. Co-Electrodeposition of Sn-Doped TiO2 Electron-Transporting Layer for Perovskite Solar Cells. Phys. Status Solidi Appl. Mater. Sci. 2020, 217, 1900491. [Google Scholar] [CrossRef]
  2. Abd Mutalib, M.; Ahmad Ludin, N.; Nik Ruzalman, N.A.A.; Barrioz, V.; Sepeai, S.; Mat Teridi, M.A.; Su’ait, M.S.; Ibrahim, M.A.; Sopian, K. Progress towards highly stable and lead-free perovskite solar cells. Mater. Renew. Sustain. Energy 2018, 7, 7. [Google Scholar] [CrossRef] [Green Version]
  3. Mohamad Noh, M.F.; Arzaee, N.A.; Nawas Mumthas, I.N.; Mohamed, N.A.; Mohd Nasir, S.N.F.; Safaei, J.; Rashid bin Mohd Yusoff, A.; Nazeeruddin, M.K.; Mat Teridi, M.A. High-humidity processed perovskite solar cells. J. Mater. Chem. A 2020, 8, 10481–10518. [Google Scholar] [CrossRef]
  4. He, R.; Ren, S.; Chen, C.; Yi, Z.; Luo, Y.; Lai, H.; Wang, W.; Zeng, G.; Hao, X.; Wang, Y.; et al. Wide-bandgap organic–inorganic hybrid and all-inorganic perovskite solar cells and their application in all-perovskite tandem solar cells. Energy Environ. Sci. 2021, 14, 5723–5759. [Google Scholar] [CrossRef]
  5. Ahsan Saeed, M.; Hyeon Kim, S.; Baek, K.; Hyun, J.K.; Youn Lee, S.; Won Shim, J. PEDOT:PSS: CuNW-based transparent composite electrodes for high-performance and flexible organic photovoltaics under indoor lighting. Appl. Surf. Sci. 2021, 567, 150852. [Google Scholar] [CrossRef]
  6. Mahmood, K.; Sarwar, S.; Mehran, M.T. Current status of electron transport layers in perovskite solar cells: Materials and properties. RSC Adv. 2017, 7, 17044–17062. [Google Scholar] [CrossRef] [Green Version]
  7. Firdaus Mohamad Noh, M.; Hoong Teh, C.; Daik, R.; Liang Lim, E.; Chin Yap, C.; Adib, M.; Ahmad Ludin, N.; Rashid bin Mohd Yusoff, A.; Jang, J.; Asri Mat Teridi, M.; et al. The architecture of the electron transport layer for a perovskite solar cell. J. Mater. Chem. C 2017, 6, 682–712. [Google Scholar] [CrossRef]
  8. Giordano, F.; Abate, A.; Correa Baena, J.P.; Saliba, M.; Matsui, T.; Im, S.H.; Zakeeruddin, S.M.; Nazeeruddin, M.K.; Hagfeldt, A.; Graetzel, M. Enhanced electronic properties in mesoporous TiO2 via lithium doping for high-efficiency perovskite solar cells. Nat. Commun. 2016, 7, 10379. [Google Scholar] [CrossRef] [PubMed]
  9. Abd Mutalib, M.; Aziz, F.; Fauzi, A.; Norharyati, W.; Salleh, W.; Yusof, N.; Jaafar, J.; Soga, T.; Zainizan, M.; Ahmad, N. Towards high performance perovskite solar cells: A review of morphological control and HTM development. Appl. Mater. Today 2018, 13, 69–82. [Google Scholar] [CrossRef]
  10. You, Y.-J.; Saeed, M.A.; Shafian, S.; Kim, J.; Hyeon Kim, S.; Kim, S.H.; Kim, K.; Shim, J.W. Energy recycling under ambient illumination for internet-of-things using metal/oxide/metal-based colorful organic photovoltaics. Nanotechnology 2021, 32, 465401. [Google Scholar] [CrossRef]
  11. Zhou, H.; Chen, Q.; Li, G.; Luo, S.; Song, T.B.; Duan, H.S.; Hong, Z.; You, J.; Liu, Y.; Yang, Y. Interface engineering of highly efficient perovskite solar cells. Science 2014, 345, 542–546. [Google Scholar] [CrossRef]
  12. Lü, X.; Mou, X.; Wu, J.; Zhang, D.; Zhang, L.; Huang, F.; Xu, F.; Huang, S. Improved-performance dye-sensitized solar cells using Nb-doped TiO2 electrodes: Efficient electron injection and transfer. Adv. Funct. Mater. 2010, 20, 509–515. [Google Scholar] [CrossRef]
  13. Zhang, H.; Shi, J.; Xu, X.; Zhu, L.; Luo, Y.; Li, D.; Meng, Q. Mg-doped TiO2 boosts the efficiency of planar perovskite solar cells to exceed 19%. J. Mater. Chem. A 2016, 4, 15383–15389. [Google Scholar] [CrossRef]
  14. 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]
  15. Arman, S.Y.; Omidvar, H.; Tabaian, S.H.; Sajjadnejad, M.; Fouladvand, S.; Afshar, S. Evaluation of nanostructured S-doped TiO2 thin films and their photoelectrochemical application as photoanode for corrosion protection of 304 stainless steel. Surf. Coat. Technol. 2014, 251, 162–169. [Google Scholar] [CrossRef]
  16. Niu, Y.; Xing, M.; Tian, B.; Zhang, J. Improving the visible light photocatalytic activity of nano-sized titanium dioxide via the synergistic effects between sulfur doping and sulfation. Appl. Catal. B Environ. 2012, 115, 253–260. [Google Scholar] [CrossRef]
  17. Dong, G.; Ha, J.; Yang, Y.; Qiu, L.; Fan, R.; Zhang, W.; Bai, L.; Gao, W.; Fu, M. 4-Tert butylpyridine induced MAPbI 3 film quality enhancement for improving the photovoltaic performance of perovskite solar cells with two-step deposition route. Appl. Surf. Sci. 2019, 484, 637–645. [Google Scholar] [CrossRef]
  18. Lee, H.; Hwang, D.; Jo, S.M.; Kim, D.; Seo, Y.; Kim, D.Y. Low-temperature fabrication of TiO2 electrodes for flexible dye-sensitized solar cells using an electrospray process. ACS Appl. Mater. Interfaces 2012, 4, 3308–3315. [Google Scholar] [CrossRef] [PubMed]
  19. Gao, S.A.; Xian, A.P.; Cao, L.H.; Xie, R.C.; Shang, J.K. Influence of calcining temperature on photoresponse of TiO2 film under nitrogen and oxygen in room temperature. Sens. Actuators B Chem. 2008, 134, 718–726. [Google Scholar] [CrossRef]
  20. Wu, M.C.; Chan, S.H.; Lee, K.M.; Chen, S.H.; Jao, M.H.; Chen, Y.F.; Su, W.F. Enhancing the efficiency of perovskite solar cells using mesoscopic zinc-doped TiO2 as the electron extraction layer through band alignment. J. Mater. Chem. A 2018, 6, 16920–16931. [Google Scholar] [CrossRef]
  21. Loan, T.T.; Huong, V.H.; Tham, V.T.; Long, N.N. Effect of zinc doping on the bandgap and photoluminescence of Zn2+-doped TiO2 nanowires. Phys. B Condens. Matter 2018, 532, 210–215. [Google Scholar] [CrossRef]
  22. Hamadanian, M.; Reisi-Vanani, A.; Behpour, M.; Esmaeily, A.S. Synthesis and characterization of Fe,S-codoped TiO2 nanoparticles: Application in degradation of organic water pollutants. Desalination 2011, 281, 319–324. [Google Scholar] [CrossRef]
  23. Lei, J.; Li, X.; Li, W.; Sun, F.; Lu, D.; Yi, J. Arrayed porous iron-doped TiO2 as photoelectrocatalyst with controllable pore size. Int. J. Hydrogen Energy 2011, 36, 8167–8172. [Google Scholar] [CrossRef]
  24. Hamadanian, M.; Reisi-Vanani, A.; Majedi, A. Preparation and characterization of S-doped TiO2 nanoparticles, effect of calcination temperature and evaluation of photocatalytic activity. Mater. Chem. Phys. 2009, 116, 376–382. [Google Scholar] [CrossRef]
  25. Wang, M.C.; Lin, H.J.; Yang, T.S. Characteristics and optical properties of iron ion (Fe3+)-doped titanium oxide thin films prepared by a sol-gel spin coating. J. Alloys Compd. 2009, 473, 394–400. [Google Scholar] [CrossRef]
  26. Kment, S.; Kmentova, H.; Kluson, P.; Krysa, J.; Hubicka, Z.; Cirkva, V.; Gregora, I.; Solcova, O.; Jastrabik, L. Notes on the photo-induced characteristics of transition metal-doped and undoped titanium dioxide thin films. J. Colloid Interface Sci. 2010, 348, 198–205. [Google Scholar] [CrossRef] [PubMed]
  27. León, A.; Reuquen, P.; Garín, C.; Segura, R.; Vargas, P.; Zapata, P.; Orihuela, P.A. FTIR and raman characterization of TiO2 nanoparticles coated with polyethylene glycol as carrier for 2-methoxyestradiol. Appl. Sci. 2017, 7, 49. [Google Scholar] [CrossRef]
  28. Basheer, C. Application of titanium dioxide-graphene composite material for photocatalytic degradation of alkylphenols. J. Chem. 2013, 2013, 456586. [Google Scholar] [CrossRef]
  29. Nagaveni, K.; Hegde, M.S.; Madras, G. Structure and photocatalytic activity of Ti1−xMxO2±δ (M = W, V, Ce, Zr, Fe, and Cu) synthesized by solution combustion method. J. Phys. Chem. B 2004, 108, 20204–20212. [Google Scholar] [CrossRef]
  30. Piątkowska, A.; Janus, M.; Szymański, K.; Mozia, S. C-, N- and S-Doped TiO2 Photocatalysts: A Review. Catalysts 2021, 11, 144. [Google Scholar] [CrossRef]
  31. Kumari, S.; Chaudhary, Y.S.; Agnihotry, S.A.; Tripathi, C.; Verma, A.; Chauhan, D.; Shrivastav, R.; Dass, S.; Satsangi, V.R. A photoelectrochemical study of nanostructured Cd-doped titanium oxide. Int. J. Hydrogen Energy 2007, 32, 1299–1302. [Google Scholar] [CrossRef]
Figure 1. UV-Vis absorption spectra and Tauc plots (inset plot) of various TiO2 photo-anodes (a) sintered at different temperatures and (b) with different loadings of S dopant.
Figure 1. UV-Vis absorption spectra and Tauc plots (inset plot) of various TiO2 photo-anodes (a) sintered at different temperatures and (b) with different loadings of S dopant.
Applsci 12 00429 g001
Figure 2. XRD diffractograms of the un-doped and S-doped TiO2 photo-anode thin films.
Figure 2. XRD diffractograms of the un-doped and S-doped TiO2 photo-anode thin films.
Applsci 12 00429 g002
Figure 3. (a) FTIR spectrum and (b) sulfur characteristic absorption peaks of undoped and S-doped TiO2 ETL sintered at 450 °C.
Figure 3. (a) FTIR spectrum and (b) sulfur characteristic absorption peaks of undoped and S-doped TiO2 ETL sintered at 450 °C.
Applsci 12 00429 g003
Figure 4. Fluorescence emission spectra of undoped and S-doped TiO2 photoanode thin film.
Figure 4. Fluorescence emission spectra of undoped and S-doped TiO2 photoanode thin film.
Applsci 12 00429 g004
Figure 5. FESEM morphological images of the EDX elemental spectra and elemental mapping of Ti, O and S of the un-doped (a,c,e) and (b,d,f) S-doped TiO2.
Figure 5. FESEM morphological images of the EDX elemental spectra and elemental mapping of Ti, O and S of the un-doped (a,c,e) and (b,d,f) S-doped TiO2.
Applsci 12 00429 g005
Figure 6. (a) LSV responses of the un-doped and S-doped TiO2 photo-anode in the dark and visible light-irradiated environment and (b) photo-current–voltage curve of the perovskite solar devices based on the S-doped TiO2 and TiO2 measured under standard AM1.5G illumination and the planar PSC structure (inset).
Figure 6. (a) LSV responses of the un-doped and S-doped TiO2 photo-anode in the dark and visible light-irradiated environment and (b) photo-current–voltage curve of the perovskite solar devices based on the S-doped TiO2 and TiO2 measured under standard AM1.5G illumination and the planar PSC structure (inset).
Applsci 12 00429 g006
Table 1. Photovoltaic parameters derived from J-V measurements for perovskite solar cells device using S-doped TiO2 and undoped TiO2 with a 0.07 cm2 active area.
Table 1. Photovoltaic parameters derived from J-V measurements for perovskite solar cells device using S-doped TiO2 and undoped TiO2 with a 0.07 cm2 active area.
Jsc (mA/cm2)VOC (V)FF (%)PCE (%)Rct (Ω)
TiO2Best13.300.9839.885.1018.72
Average13.18 ± 0.080.97 ± 0.0138.70 ± 0.774.98 ± 0.1320.68 ± 1.54
S-doped TiO2Best13.90.99743.186.014.26
Average13.65 ± 0.170.99 ± 0.00342.19 ± 0.835.76 ± 0.1914.68 ± 1.05
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Abd Mutalib, M.; Ahmad Ludin, N.; Su’ait, M.S.; Davies, M.; Sepeai, S.; Mat Teridi, M.A.; Mohamad Noh, M.F.; Ibrahim, M.A. Performance-Enhancing Sulfur-Doped TiO2 Photoanodes for Perovskite Solar Cells. Appl. Sci. 2022, 12, 429. https://doi.org/10.3390/app12010429

AMA Style

Abd Mutalib M, Ahmad Ludin N, Su’ait MS, Davies M, Sepeai S, Mat Teridi MA, Mohamad Noh MF, Ibrahim MA. Performance-Enhancing Sulfur-Doped TiO2 Photoanodes for Perovskite Solar Cells. Applied Sciences. 2022; 12(1):429. https://doi.org/10.3390/app12010429

Chicago/Turabian Style

Abd Mutalib, Muhazri, Norasikin Ahmad Ludin, Mohd Sukor Su’ait, Matthew Davies, Suhaila Sepeai, Mohd Asri Mat Teridi, Mohamad Firdaus Mohamad Noh, and Mohd Adib Ibrahim. 2022. "Performance-Enhancing Sulfur-Doped TiO2 Photoanodes for Perovskite Solar Cells" Applied Sciences 12, no. 1: 429. https://doi.org/10.3390/app12010429

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