Next Article in Journal
Laser-Ablative Synthesis of Silicon–Iron Composite Nanoparticles for Theranostic Applications
Previous Article in Journal
Tailing Optical Pulling Force on a Metal–Dielectric Hybrid Dimer with Electromagnetic Coupling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation of Structural, Morphological, and Optical Properties of Novel Electrospun Mg-Doped TiO2 Nanofibers as an Electron Transport Material for Perovskite Solar Cells

1
Department of Physics, Gebze Technical University, Gebze 41400, Turkey
2
Department of Chemical Engineering, Gebze Technical University, Gebze 41400, Turkey
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(15), 2255; https://doi.org/10.3390/nano13152255
Submission received: 24 June 2023 / Revised: 27 July 2023 / Accepted: 31 July 2023 / Published: 5 August 2023
(This article belongs to the Section Solar Energy and Solar Cells)

Abstract

:
Perovskite solar cells (PSCs) are quickly becoming efficient solar cells due to the effective physicochemical properties of the absorber layer. This layer should ideally be placed between a stable hole transport material (HTM) layer and a conductive electron transport material (ETM) layer. These outer layers play a critical role in the current densities and cell voltages of solar cells. In this work, we successfully fabricated Mg-doped TiO2 nanofibers as ETM layers via electrospinning. This study systematically investigates the morphological and optical features of Mg-doped nanofibers as mesoporous ETM layers. The existence of the Mg element in the lattice was confirmed by XRD and XPS. These optical characterizations indicated that Mg doping widened the energy band gap and shifted the edge of the conduction band minimum upward, which enhanced the open circuit voltage (Voc) and short current density (Jsc). The electron-hole recombination rate was lowered, and separation efficiency increased with Mg doping. The results have demonstrated the possibility of improving the efficiency of PSCs with the use of Mg-doped nanofibers as an ETM layer.

1. Introduction

The broad use of technology and the increase in world population have caused extensive energy demands and environmental awareness in developed and developing countries. Solar energy has a significant capability for future energy that can aid in reducing fossil fuel use and growing environmental concerns such as global warming [1]. In recent years, perovskite solar cells (PSCs) have attracted intensive attention because of their ever-increasing power conversion efficiency (PCE), low-cost material constituents, and simple solution fabrication processes [2,3]. The conventional configuration of perovskite solar cells has five layers: (i) transparent conducting oxide (TCO), which can be made of fluorine-doped tin oxide (FTO) or indium tin oxide (ITO), (ii) the planar or mesoporous electron transporting material (ETM), (iii) the light-absorbing material made of perovskite materials, (iv) the hole transporting material (HTM), and (v) the electrode (Figure 1a). The perovskite layer absorbs the incident of light, generating electron-hole pairs, and subsequently, the electrons are transported to the ETM layer, while the holes are transported to the HTM layer. These charge carriers are finally collected by electrodes forming a photovoltage, which depends on the gap between the conduction band minimum (CBM) of ETM and the valence band maximum (VBM) of HTM (Figure 1b) [4]. Therefore, tuning the energy band gaps and the band edge positions of ETM and HTM layers is of critical importance to the performance of PSCs.
The ETM layer has two main roles in the electron transport mechanism; the first role is to receive the photoexcited electrons from the perovskite absorber and transport the collected charges to the conductive layer, and the second role is to block the generated holes [4]. Titanium dioxide (TiO2) is widely used as an ETM due to its favorable band gap positioning, stability, and low cost in perovskite solar cells [5]. To replace TiO2 as an ETL layer in PSCs, scientists and researchers are investigating other types of inorganic materials, such as metal oxides like ZnO, SnO2, WO3, In2O3, CuO, and Cu2O [6,7]. ZnO, which is one of the best candidates due to a number of reasons, is not the optimal material to replace TiO2 as the ETL in PSC devices. For example, ZnO is naturally hygroscopic, which could lead to perovskite disintegration in ambient air [8]. PSCs based on low-temperature processed ZnO as an ETL were additionally environmentally fragile because any hydroxyl residues that remained on the ETL surface could hasten the breakdown of the perovskite structure [9].
Two types of structures can be used in the architecture of an ETM layer; planar (bl-TiO2) or mesoporous (mp-TiO2) layers [6]. In a planar structure, TiO2 is deposited on FTO film as a compact layer. With the implementation of this geometry, Granados et al. obtained 6.29% of PCE for an FTO/bl-TiO2/mp-TiO2/Perovskite/P3HT/Au cell structure [10]. However, it is crucial to fabricate a mesoporous structure to achieve high-efficiency perovskite solar cells. Having a larger surface area offers the benefit of facilitating a more significant interaction between the ETM and the perovskite active layer, thereby promoting faster charge transfer kinetics. In modern mesoporous perovskite solar cells (mp-PSCs), TiO2 nanoparticles are often employed as nanostructured scaffolds. Because of their labyrinthine shape and the presence of grain boundaries that interfere with the electron transport channels within mp-TiO2 nanoparticles, the mesoporous layer that uses nanoparticles might have difficulty in loading perovskite [11]. To solve this problem, numerous one-dimensional TiO2 nanostructures have been used as ETLs in PSCs, including nanorods, nanowires, nanotubes, and nanocones [12]. Among the mesoporous fabrication methods, electrospinning is a simple, inexpensive technique by which an ETM layer can be fabricated [13]. Dharani et al. used electrospun TiO2 fibers as an ETM layer leading to a device efficiency as high as 9.8% [14].
As an ETM layer, TiO2 doped with metal ions such as Mg, Li, Y, and Nb have been utilized to increase the efficiency of the system to over 19% as a result of improving the extraction and/or injection of carriers [15,16,17,18]. Amalathas et al. studied Li-doped mesoporous (mp-TiO2) and obtained a 17.59% power conversion efficiency (PCE) [15]. Roose and colleagues studied mesoporous (mp-TiO2) Nd-doped electrodes for perovskite-based solar cells and obtained an 18.2% PCE [19]. A PCE of 19.8% was achieved with Mg-doped TiO2 electrodes in perovskite solar cells [20]. The Mg-doped TiO2, which had a larger band gap and lower resistance compared to the non-doped TiO2, has important advantages [21]. Firstly, a higher band gap showed better optical transmission properties, allowing a higher number of photons to excite the perovskite material leading to elevating the short current density (Jsc) [22]. Secondly, Mg doping lifted up the CBM and deflated the VBM of the ETM layer. This enlargement provided PSCs with higher values of open circuit voltage (Voc) and fill factor (FF) [23]. Third, the shift in CBM provided a better energy-level alignment at the interface, which could be quantitatively stated by the conduction band offset (CBO): the difference between the energy levels of the CBM of the ETL and that of the perovskite layer [24]. This alignment suppressed the electron-hole recombination and improved both the Voc and Jsc when employed in PSCs.
Mg-doped TiO2 nanofibers as an ETM layer fabricated by the electrospinning method have not been investigated yet. There are few reports available for Mg-doped TiO2 nanoparticles to be carried out through a complex sol–gel process, which is expensive and slow and was deposited using the spin coating method [20,21].
To the best of our knowledge, this is the first study that successfully shows the fabrication of Mg-doped TiO2 nanofibers that can provide a larger interfacial contact area to support film deposition as a mesoporous ETM layer. In addition, a more homogeneous structure was obtained by doping TiO2 nanofibers with Mg ions before the electrospinning process. The effect of the calcination temperature and time on the crystal structure, morphology, and optical properties of the nanofibers was also investigated. After the calcination of fibers at 500 °C, XRD analyses were carried out to reveal the crystal structure and UV-Vis absorption was utilized to determine the band gap of each sample.

2. Materials and Methods

2.1. Materials

Titanium tetraisopropoxide (TTIP), polyvinylpyrrolidone (PVP; Mw 1,300,000), ethanol, and acetic acid were obtained from Sigma Aldrich. Magnesium chloride (MgCl2) was obtained from Merck. TTIP and MgCl2 were used as precursors. Ethanol and distilled water were used as a solvent. FTO glass (10 Ω per square) with the dimensions 25 mm × 75 mm × 2.2 mm was obtained from Teknoma Technological Materials Industrial and Trading Inc.

2.2. Experimental

In the fabrication of the ETM layer, firstly, FTO substrates were cleaned ultrasonically in acetone, isopropyl alcohol, and ethanol for 15 min before being sequentially and finally dried under nitrogen medium (99.9% purity). The substrates were exposed to air plasma (at 300 mTorr pressure) for 15 min just before the deposition of the compact layer. For the compact layer, the precursor solution was prepared by mixing two solutions. The first solution was composed of 1.5 mL TTIP and 4.5 mL ethanol (≥99.9%), stirred at 1000 rpm for 15 min. The second solution was composed of 1 mL acetic acid (99.8–100.5%) and 3 mL ethanol, stirred at 1000 rpm for 15 min. The second solution was added to the first solution drop by drop and stirred for 30 min at 1000 rpm. Then, the mixed solution aged for one day. During aging, the solution was covered with aluminum foil and stored in a dark medium. Before the deposition step, the solution was diluted by absolute ethanol 1:1 v/v. The compact layer was deposited two times on FTO glass by spin coating for 30 s at 3000 rpm and was annealed at 120 °C for 10 min. After the preparation of the compact layer, which provided better adhesion of the fibers when formed on the surface, the electrospinning process was carried out.
Figure 2 shows the schematic illustration of the TiO2 nanofiber fabrication process. Step I shows the polymer solution preparation. For non-doped TiO2 nanofibers, 0.25 g of titanium tetraisopropoxide was hydrolyzed with a mixture of 1.4 mL ethanol and 0.6 mL acetic acid and aged for 48 h. The PVP (9 wt.%) was dissolved in 2.75 mL ethanol for three hours. Then, the TiO2 solution was added drop by drop to the PVP solution and stirred for half an hour. For the Mg-doped TiO2 fibers, PVP was dissolved in 2.75 mL of ethanol, and then MgCl2 was added to the PVP solution at various concentrations (0.1% and 0.5%). Then, the TiO2 solution, as prepared before, was added drop by drop to the MgCl2-PVP solution. The mixed precursor solution was stirred for half an hour at room temperature to attain the sufficient viscosity required for electrospinning.
For the electrospinning process, the solutions were loaded into a syringe with a 25-gauge (25 G) needle, and the distance between the needle and collector was adjusted to 10–15 cm. A direct current was applied at 10–20 kV, and the feed rate of the precursor solutions was adjusted to 0.5 mL/h, controlled by using a syringe pump. The humidity of the air was around 50%. Aluminum foil was used for nanofiber collection and for optical and microscopic measurements. Fiber diameters were measured from SEM images using the ImageJ analysis program. Finally, all fibers were calcined at 500 °C [25,26] for 2 h in a Protherm PTF 12/40/250 tube furnace for the transition from the amorphous structure to the crystal structure.

3. Results and Discussion

3.1. SEM Characterization

The surface morphology of TiO2-PVP nanofiber samples before calcination was investigated to optimize the electrospinning operating conditions. Figure 3 shows SEM images of electrospun TiO2-PVP nanofibers when fabricated by a solution of ethanol/acetic acid/PVP/Ti(OiPr)4 at a feed rate of 0.5 mL/h. Uniform fibers with different diameters were achieved at various voltage fields and collector distances. The fiber diameters decreased with an increase in the applied electric field and increased with an extension in the collector distance. The obtained diameters are consistent with the literature [27].
The solutions of ethanol/acetic acid/PVP/MgCl2/Ti(OiPr)4 including 0.1% (wt.) MgCl2 and 0.5% (wt.) MgCl2 were used to produce Mg-doped TiO2 fibers. It is known that after the calcination process, a shrinkage in the nanofiber diameter can be observed [14]; therefore, fabricating higher diameter nanofibers compared to TiO2-PVP nanofibers was crucial. After our preliminary trials, the nanofibers were deposited on a glass substrate at a potential of 20 kV, a distance of 15 cm, and a feed rate of 0.5 mL/h. Then, nanofibers were calcinated at 500 °C for 2 h on the glass substrate for SEM characterization and in a crucible for XRD measurements.
Figure 4 depicts the SEM images of the TiO2-PVP nanofibers with MgCl2 and Mg-doped TiO2 nanofibers after calcination. Table 1 summarizes the nanofiber’s diameters and shrinkage percentages. The mean fiber diameters for samples A and C without calcination were 249 ± 68 nm and 262 ± 88 nm, and the sample area was 2365 nm2 and 972 nm2, respectively. The higher amount of Mg-doped in sample C led to a slight increase in the fiber diameter. On the other hand, after the calcination process for samples B and D, the fibers shrank, and a significant decrease in the diameter was observed. The diameters of samples B and D were measured as 80 ± 30 nm and 134 ± 32 nm, and the sample area 306 nm2 and 499 nm2, respectively. The shrinkage percentages of samples with 0.1% and 0.5% MgCl2 were 68% and 50%, respectively. However, it was concluded that a higher amount of Mg restricted the shrinkage percentages. This effect could be attributed to the insertion of the dopants into the TiO2 lattice, which might cause lattice and volume expansion [28].
The EDX elemental mapping used to determine the homogeneous distribution of the Mg element is shown in Figure 5. Before and after the calcination process, Mg element mapping showed a uniform distribution in the nanofibers. These images indicate that Mg doping was achieved throughout the nanofibers. The cross-section view of the nanofibers after calcination can be seen in Figure 5c.

3.2. XRD and XPS Analysis

TiO2 is a polymorphous material and has three crystalline phases, namely anatase, brookite, and rutile [29]. A stronger surface interaction and simpler charge transfer were also indicated by the fact that the bond lengths in a rutile TiO2-perovskite system were shorter than those in an anatase TiO2-perovskite system [30]. XRD techniques were used to examine the crystal structure and phase evolution of the electrospun TiO2 nanofibers (Rigaku Slab Cu k-alpha 1.58 A Xray). For the phase evolution, Figure 6a,b show the XRD patterns for non-doped TiO2 nanofibers for different calcination times, and Figure 6c,d show 0.1% and 0.5% Mg-doped nanofibers, respectively. Mg and MgO peaks are indicated [31]. The as-spun nanofibers were amorphous. Well-crystallized anatase-rutile TiO2 and Mg-doped TiO2 nanofibers were obtained by heating the as-spun fibers at 500 °C. It was observed that at 2 h calcination time, the rutile phase occurred more clearly. In Figure 6c,d, the Mg peak was observed at 44° (012), as consistent with the literature [32]. However, the doping of Mg drove the direction of rutile to anatase as expected, and the rutile phase intensity was reduced dramatically. Because the Mg2+ ions were incorporated in the anatase lattice, this resulted in Mg-doped TiO2 nanofibers with a higher conduction band.
We could determine the crystal size (D) using the Scherer relation D = k × λ/β × (cosθ), where k is the constant (0.9), the X-ray wavelength (0.154 nm) is β, and the FWHM value is the reflection angle. The crystal size of undoped TiO2 and MgCl2-doped TiO2 films is displayed in Table 2.
As shown in Table 2, when the amount of doping increased, there was a small reduction in the crystallite size of the crystal structure. This was because some of the Mg dopant ions did not proceed into the TiO2 lattice and instead remained on the surface, creating grain boundaries. This could impact the growth of the crystal structure, resulting in a smaller crystal size as the doping concentration increased. However, if the amount of doping increased too much, it could inhibit crystal growth and increase the number of grain boundaries. This could have a negative impact on the performance of the material as a photo-anode [33]. In addition, the smaller crystal sizes reduced these defects and generally resulted in a larger surface area and higher porosity, which could improve the contact between the ETL and the perovskite layer, leading to improved device performance.
In Figure 7, there is an increase in the binding energy (BE) in Mg-doped TiO2 relative to non-doped TiO2. For the XPS fittings, adj.R-squared values ranged between 0.95 and 0.98. This shift showed the interaction of the Ti element with the dope material. The charge transfer through Mg doping could result in a slight shift of BE without the formation of different phases (MgO) [22]. The Ti 2p1/2 and Ti 2p3/2 binding energy of Mg-doped TiO2 was higher than that of non-doped TiO2, and this meant that due to the stronger metallicity and lower valence state, maybe Ti4+ could donate electrons easier than Mg2+. For MgO, the 458.7 eV curve shows 3/2 Ti0, and 460.7 eV shows ½ Ti0 (Figure 7e,f).

3.3. Optical Band Gap and UPS Characterizations

The optical properties of non-doped TiO2 and Mg-doped TiO2 nanofibers were characterized by a UV-Vis spectrophotometer. The optical responses of these samples are displayed in Figure 8a. The compact layer was used as a base measurement. The results show that samples with calcinated nanofibers had a higher absorbance in UV and visible regions. This resulted from the interaction of light with a higher surface area of nanofibers. Among samples with nanofibers, the absorbance of Mg-doped TiO2 nanofibers was less than that of non-doped TiO2 nanofibers from 320 to 900 nm. On the other hand, for the wavelength below 320 nm, the opposite behavior was observed. This absorption capability is reported in the literature [34]. It is a result of Mg doping that shifts the absorption edge from a visible to a UV region. It can be stated that the substitution of Mg in Ti sites and the formation of Mg-Ti mixed oxides could enlarge the band gap [33]. Therefore, Mg doping enhances the absorption of light by the perovskite layer with the improved Jsc when the incident light passes through the ETM layer. The absorption onset appeared to be blue-shifted with the increasing concentration of Mg. However, the light absorption of 0.5% MgCl2 doped films deteriorated, which could be attributed to the crystal size.
The optical band gaps were derived from the Tauc plot [35]. The Tauc relation is ( a h v 1 / 2 ) = A ( h v E g ) , where a is the absorption coefficient, A is the absorption constant, E g is the band gap and h v is the photon energy; h is Planck’s constant, v is the frequency of light, and ½ is for the indirect band gap energy. The measured band gaps of non-doped TiO2, 0.1% Mg-doped TiO2, and 0.5% Mg-doped TiO2 nanofiber films are presented in Figure 8b. The non-doped TiO2 nanofiber’s band gap was 3.20 eV, and Mg-TiO2 showed a band gap value of 3.27 eV, which is consistent with the value reported in the literature [36]. When the amount of Mg doping was increased to 0.5%, a small increase in the band gap from 3.27 eV to 3.28 eV was obtained.
The changes in absorption and band gap tuning that occurred could be linked to the creation of trapping states and intermediate sources that can transport electrons. The addition of Mg as a dopant resulted in the formation of oxygen vacancies and localized trap states, which helped transport electrons. Therefore, the vacancies generated by these dopant atoms were highly effective for adjusting the material’s optoelectronic properties. The introduction of Mg as a dopant changes the structure of the material by rearranging the atoms into their respective lattice positions. As a result, the low bandgap energy of the Ti nanofibers is likely due to the larger average crystal size. When Mg is added to TiO2 nanofiber films, the crystal size decreases, and the band gap shifts. This increment in the band gap between the conduction and valence bands was anticipated to improve the injection of charges from the perovskite layer to the ETL.
In order to achieve the relatively accurate position of energy bands, UPS characterization was performed. Figure 9a depicts the UPS spectra obtained using He I radiation. The VBM value is calculated by the sum of the work function and the peak value at the extended valance spectra. The work function is derived from subtracting the cut-off banding energy (the right-hand side of the spectra) from the photon energy (21.21 eV). The work function of non-doped TiO2 can be calculated as 5.19 eV, that of doped TiO2 with 0.1% Mg as 5.05 eV, and that of doped TiO2 with 0.5% Mg as 5.21 eV. The extended valance spectra are displayed on the left-hand side of the spectra. The non-doped TiO2 film, the 0.1% Mg-doped TiO2 film, and the 0.5% Mg-doped TiO2 film, respectively, had peaks centered at 1.52, 1.67, and 1.51 eV. The VBM value was calculated by the sum of the work function and the peak value at the extended valance spectra. Therefore, the VBM positions of non-doped TiO2, doped TiO2 with 0.1% Mg, and doped TiO2 with 0.5% Mg were 6.71 eV, 6.72 eV, and 6.72 eV, respectively. The wider band gap that occurred in Mg-doped TiO2 nanofibers was due to an ascending shift in CBM [22].

3.4. Photoluminescence (PL) Spectra

Photoluminescence (PL) is the spontaneous emission of light from a material following optical excitation. It is a powerful technique to investigate the recombination rate of photoexcited electron-hole pairs in semiconductor materials. The PL spectrum is a result of the irradiative recombination of the photoexcited electron and hole; thus, a lower PL intensity indicates a lower electron-hole recombination rate. Figure 10 indicates that the intensity of Mg-doped TiO2 nanofiber samples was lower than those of non-doped TiO2. This behavior implies that for the Mg-doped sample, the electron-hole recombination rate was much lower, and separation efficiency was much higher.
The suppression of electron-hole recombination is critical for the energy-level alignment at the interface of TiO2 and perovskite layers. When the CBM of the TiO2 is lower than that of the perovskite (cliff structure), interface recombination becomes dominant, and the Voc declines. When the CBM of the TiO2 is higher than that of the perovskite (spike structure), the interfacial recombination is largely cut off, and an increase in the Voc of the solar cells is achieved [37]. With Mg doping, the cliff structure is minimized, which suppresses the electron-hole recombination and increases the efficiency of the perovskite cells. Figure 9b shows that the CBM of Mg-doped TiO2 samples increased with diminishing the cliff structure. Therefore, the decrease in the PL intensity was a result of this CBM tuning.

4. Conclusions

In this work, TiO2 nanofibers with different Mg doping concentrations were successfully fabricated using the electrospinning technique for the first time as the ETM layer. The nanofibers became crystalline after calcination at 500 °C for 2 h. From XRD patterns, anatase and rutile phases of TiO2 nanofibers were obtained. According to SEM analysis, nanofiber diameters lessened during calcination. XRD results confirmed Mg doping in anatase TiO2 with a structural change in terms of increasing crystallinity due to Mg doping. The presence of the Mg dopant in the TiO2 was also confirmed by the shift in Ti2p peaks in the XPS spectrum. The nanofiber samples, compared with the compact layer samples, showed a higher absorbance in UV and visible regions due to the interaction of light with an enhanced surface area. The absorption coefficient of Mg-doped TiO2 nanofibers became less than that of non-doped TiO2 nanofibers in the visible spectrum. Mg doping shifted the absorption edge from the visible to the UV region. Therefore, when incident light passed through the ETM layer, this shift increased the perovskite layer’s capacity to absorb light with an improved Jsc. Adding Mg into the ETM layer not only increased the energy band gap, which contributed to the improvement of the Voc but also elevated the position of the CBM of TiO2, which lowered the electron-hole recombination rate and high Jsc based on PL measurements. The TiO2 mesoporous layer’s porosity is a crucial component in the development of perovskite crystal, which is crucial for charge carrier extraction and light harvesting effectiveness. The TiO2 mesoporous layer’s porosity was obtained at around 50%. In future work, we aim to fabricate the perovskite solar cell to investigate the performance of the cell in which perovskite infiltrates through the Mg-doped TiO2 ETM layer.

Author Contributions

Conceptualization, O.Y. and M.E.O.; methodology, O.Y. and M.E.O.; formal analysis, K.E., O.Y. and M.E.O.; investigation, K.E., O.Y. and M.E.O.; data curation, K.E., O.Y. and M.E.O.; validation, K.E., O.Y. and M.E.O.; writing—original draft preparation, K.E.; writing—review and editing, O.Y. and M.E.O.; supervision, O.Y. and M.E.O. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by The Scientific and Technological Research Council of Türkiye (TUBITAK) with project no. 120M233.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank Mevlut Karabulut for his help in acquiring and interpretating PL data.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Halder, B.; Banik, P.; Almohamad, H.; Al Dughairi, A.A.; Al-Mutiry, M.; Al Shahrani, H.F.; Abdo, H.G. Land Suitability Investigation for Solar Power Plant Using GIS, AHP and Multi-Criteria Decision Approach: A Case of Megacity Kolkata, West Bengal, India. Sustainability 2022, 14, 11276. [Google Scholar] [CrossRef]
  2. Rong, Y.; Hu, Y.; Mei, A.; Tan, H.; Saidaminov, M.I.; Seok, S.I.; McGehee, M.D.; Sargent, E.H.; Han, H. Challenges for commercializing perovskite solar cells. Science 2018, 361, eaat8235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Pang, S.; Zhou, Y.; Wang, Z.; Yang, M.; Krause, A.R.; Zhou, Z.; Zhu, K.; Padture, N.P.; Cui, G. Transformative Evolution of Organolead Triiodide Perovskite Thin Films from Strong Room-Temperature Solid-Gas Interaction between HPbI3-CH3NH2 Precursor Pair. J. Am. Chem. Soc. 2016, 138, 750–753. [Google Scholar] [CrossRef]
  4. Tong, X.; Lin, F.; Wu, J.; Wang, Z.M. High performance perovskite solar cells. Adv. Sci. 2015, 3, 7867–7918. [Google Scholar] [CrossRef] [PubMed]
  5. Arshad, Z.; Shakir, S.; Khoja, A.H.; Javed, A.H.; Anwar, M.; Rehman, A.; Javaid, R.; Qazi, U.Y.; Farrukh, S. Performance Analysis of Calcium-Doped Titania (TiO2) as an Effective Electron Transport Layer (ETL) for Perovskite Solar Cells. Energies 2022, 15, 1408. [Google Scholar] [CrossRef]
  6. Mohamad Noh, M.F.; Teh, C.H.; Daik, R.; Lim, E.L.; Yap, C.C.; Ibrahim, M.A.; Ahmad Ludin, N.; Bin Mohd Yusoff, A.R.; Jang, J.; Mat Teridi, M.A. The architecture of the electron transport layer for a perovskite solar cell. J. Mater. Chem. C 2018, 6, 682–712. [Google Scholar] [CrossRef]
  7. Zuo, C.; Ding, L. Solution-Processed Cu2O and CuO as Hole Transport Materials for Efficient Perovskite Solar Cells. Small 2015, 11, 5528–5532. [Google Scholar] [CrossRef]
  8. Heo, J.H.; Lee, M.H.; Han, H.J.; Patil, B.R.; Yu, J.S.; Im, S.H. Highly efficient low temperature solution processable planar type CH3NH3PbI3 perovskite flexible solar cells. J. Mater. Chem. A 2016, 4, 1572–1578. [Google Scholar] [CrossRef]
  9. Song, J.; Bian, J.; Zheng, E.; Wang, X.F.; Tian, W.; Miyasaka, T. Efficient and environmentally stable perovskite solar cells based on ZnO electron collection layer. Chem. Lett. 2015, 44, 610–612. [Google Scholar] [CrossRef] [Green Version]
  10. Hernández-Granados, A.; Corpus-Mendoza, A.N.; Moreno-Romero, P.M.; Rodríguez-Castañeda, C.A.; Pascoe-Sussoni, J.E.; Castelo-González, O.A.; Menchaca-Campos, E.C.; Escorcia-García, J.; Hu, H. Optically uniform thin films of mesoporous TiO2 for perovskite solar cell applications. Opt. Mater. 2019, 88, 695–703. [Google Scholar] [CrossRef]
  11. Quy, H.V.; Truyen, D.H.; Kim, S.; Bark, C.W. Facile Synthesis of spherical TiO2 hollow nanospheres with a diameter of 150 nm for high-performance mesoporous perovskite solar cells. Materials 2021, 14, 629. [Google Scholar] [CrossRef]
  12. Yun, J.H.; Wang, L.; Amal, R.; Ng, Y.H. One-dimensional TiO2 nanostructured photoanodes: From dye-sensitised solar cells to perovskite solar cells. Energies 2016, 9, 1030. [Google Scholar] [CrossRef] [Green Version]
  13. Kim, W.T.; Park, D.C.; Yang, W.H.; Cho, C.H.; Choi, W.Y. Effects of electrospinning parameters on the microstructure of pvp/tio2 nanofibers. Nanomaterials 2021, 11, 1616. [Google Scholar] [CrossRef] [PubMed]
  14. Dharani, S.; Mulmudi, H.K.; Yantara, N.; Thu Trang, P.T.; Park, N.G.; Graetzel, M.; Mhaisalkar, S.; Mathews, N.; Boix, P.P. High efficiency electrospun TiO2 nanofiber based hybrid organic-inorganic perovskite solar cell. Nanoscale 2014, 6, 1675–1679. [Google Scholar] [CrossRef]
  15. Peter Amalathas, A.; Landová, L.; Conrad, B.; Holovský, J. Concentration-Dependent Impact of Alkali Li Metal Doped Mesoporous TiO2 Electron Transport Layer on the Performance of CH3NH3PbI3 Perovskite Solar Cells. J. Phys. Chem. C 2019, 123, 19376–19384. [Google Scholar] [CrossRef]
  16. 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 highlyefficient perovskite solar cellsaccumbens. Science 2014, 345, 535–542. [Google Scholar]
  17. Song, J.; Zheng, E.; Liu, L.; Wang, X.F.; Chen, G.; Tian, W.; Miyasaka, T. Magnesium-doped Zinc Oxide as Electron Selective Contact Layers for Efficient Perovskite Solar Cells. ChemSusChem 2016, 9, 2640–2647. [Google Scholar] [CrossRef]
  18. 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] [Green Version]
  19. Roose, B.; Gödel, K.C.; Pathak, S.; Sadhanala, A.; Baena, J.P.C.; Wilts, B.D.; Snaith, H.J.; Wiesner, U.; Grätzel, M.; Steiner, U.; et al. Enhanced efficiency and stability of perovskite solar cells through Nd-doping of mesostructured TiO2. Adv. Energy Mater. 2016, 6, 1501868. [Google Scholar] [CrossRef]
  20. 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]
  21. Arshad, Z.; Khoja, A.H.; Shakir, S.; Afzal, A.; Mujtaba, M.A.; Soudagar, M.E.M.; Fayaz, H.; Saleel C, A.; Farukh, S.; Saeed, M. Magnesium doped TiO2as an efficient electron transport layer in perovskite solar cells. Case Stud. Therm. Eng. 2021, 26, 101101. [Google Scholar] [CrossRef]
  22. Rafieh, A.I.; Ekanayake, P.; Wakamiya, A.; Nakajima, H.; Lim, C.M. Enhanced performance of CH3NH3PbI3-based perovskite solar cells by tuning the electrical and structural properties of mesoporous TiO2 layer via Al and Mg doping. Sol. Energy 2019, 177, 374–381. [Google Scholar] [CrossRef]
  23. Yang, W.S.; Noh, J.H.; Jeon, N.J.; Kim, Y.C.; Ryu, S.; Seo, J.; Seok, S. Il High-performance photovoltaic perovskite layers fabricated through intramolecular exchange. Science 2015, 348, 1234–1237. [Google Scholar] [CrossRef] [PubMed]
  24. Ding, C.; Zhang, Y.; Liu, F.; Kitabatake, Y.; Hayase, S.; Toyoda, T.; Yoshino, K.; Minemoto, T.; Katayama, K.; Shen, Q. Effect of the conduction band offset on interfacial recombination behavior of the planar perovskite solar cells. Nano Energy 2018, 53, 17–26. [Google Scholar] [CrossRef]
  25. Cheng, Y.; Huang, W.; Zhang, Y.; Zhu, L.; Liu, Y.; Fan, X.; Cao, X. Preparation of TiO2 hollow nanofibers by electrospining combined with sol-gel process. CrystEngComm 2010, 12, 2256–2260. [Google Scholar] [CrossRef]
  26. Fujihara, K.; Kumar, A.; Jose, R.; Ramakrishna, S.; Uchida, S. Spray deposition of electrospun TiO2 nanorods for dye-sensitized solar cell. Nanotechnology 2007, 18, 365709. [Google Scholar] [CrossRef]
  27. Jinchu, I.; Krishnan, B.; Sreekala, C.O.; Balakrishnan, N.; Sajeev, U.S.; Sreelatha, K.S. Escalating the performance of perovskite solar cell via electrospun TiO2 nanofibers. Int. Conf. Electr. Electron. Optim. Tech. ICEEOT 2016, 2016, 4158–4160. [Google Scholar] [CrossRef]
  28. Çalışır, M.D. Fabrication of nanostructured metal oxide materials and their use in energy and environmental applications (No:513142001 Doctoral dissertation), Istanbul Technical University, Repository. 2020. Available online: https://polen.itu.edu.tr/bitstreams/804af29d-2142-4955-ac1f-47e57dfddcdb/download (accessed on 30 July 2023).
  29. Arbiol, J.; Cerdà, J.; Dezanneau, G.; Cirera, A.; Peiró, F.; Cornet, A.; Morante, J.R. Effects of Nb doping on the TiO2 anatase-to-rutile phase transition. J. Appl. Phys. 2002, 92, 853–861. [Google Scholar] [CrossRef]
  30. Geng, W.; Tong, C.J.; Liu, J.; Zhu, W.; Lau, W.M.; Liu, L.M. Structures and Electronic Properties of Different CH3NH3PbI3/TiO2 Interface: A First-Principles Study. Sci. Rep. 2016, 6, 20131. [Google Scholar] [CrossRef] [Green Version]
  31. Abrinaei, F. Nonlinear optical response of Mg/MgO structures prepared by laser ablation method. J. Eur. Opt. Soc. 2017, 13, 15. [Google Scholar] [CrossRef] [Green Version]
  32. Shivaraju, H.P.; Midhun, G.; Anil Kumar, K.M.; Pallavi, S.; Pallavi, N.; Behzad, S. Degradation of selected industrial dyes using Mg-doped TiO2 polyscales under natural sun light as an alternative driving energy. Appl. Water Sci. 2017, 7, 3937–3948. [Google Scholar] [CrossRef] [Green Version]
  33. Shakir, S.; Abd-ur-Rehman, H.M.; Yunus, K.; Iwamoto, M.; Periasamy, V. Fabrication of un-doped and magnesium doped TiO2 films by aerosol assisted chemical vapor deposition for dye sensitized solar cells. J. Alloys Compd. 2018, 737, 740–747. [Google Scholar] [CrossRef]
  34. Manju, J.; Jawhar, S.M.J. Synthesis of magnesium-doped TiO2 photoelectrodes for dye-sensitized solar cell applications by solvothermal microwave irradiation method. J. Mater. Res. 2018, 33, 1534–1542. [Google Scholar] [CrossRef]
  35. Jabeen, N.; Zaidi, A.; Hussain, A.; Hassan, N.U.; Ali, J.; Ahmed, F.; Khan, M.U.; Iqbal, N.; Elnasr, T.A.S.; Helal, M.H. Single- and Multilayered Perovskite Thin Films for Photovoltaic Applications. Nanomaterials 2022, 12, 3208. [Google Scholar] [CrossRef]
  36. Zhang, C.; Chen, S.; Mo, L.E.; Huang, Y.; Tian, H.; Hu, L.; Huo, Z.; Dai, S.; Kong, F.; Pan, X. Charge recombination and band-edge shift in the dye-sensitized Mg2+-doped TiO2 solar cells. J. Phys. Chem. C 2011, 115, 16418–16424. [Google Scholar] [CrossRef]
  37. Sabbah, H. Numerical Simulation of 30% Efficient Lead-Free Perovskite CsSnGeI3-Based Solar Cells. Materials 2022, 15, 3229. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) The general configuration and (b) Energy level diagram of Perovskite solar cells.
Figure 1. (a) The general configuration and (b) Energy level diagram of Perovskite solar cells.
Nanomaterials 13 02255 g001
Figure 2. Fabrication steps of electrospun TiO2 nanofibers: (I) Polymer solution preparation, (II) Electrospinning process and (III) Calcination process.
Figure 2. Fabrication steps of electrospun TiO2 nanofibers: (I) Polymer solution preparation, (II) Electrospinning process and (III) Calcination process.
Nanomaterials 13 02255 g002
Figure 3. SEM images of the electrospun TiO2-PVP nanofibers (a) Voltage: 10 kV, Distance: 10 cm, Diameter: 85 nm (b) Voltage: 20 kV, Distance: 10 cm, Diameter: 70 nm (c) Voltage: 10 kV, Distance: 15 cm, Diameter: 108 nm (d) Voltage: 20 kV, Distance: 15 cm, Diameter: 74 nm.
Figure 3. SEM images of the electrospun TiO2-PVP nanofibers (a) Voltage: 10 kV, Distance: 10 cm, Diameter: 85 nm (b) Voltage: 20 kV, Distance: 10 cm, Diameter: 70 nm (c) Voltage: 10 kV, Distance: 15 cm, Diameter: 108 nm (d) Voltage: 20 kV, Distance: 15 cm, Diameter: 74 nm.
Nanomaterials 13 02255 g003
Figure 4. SEM micrographs of the electrospun nanofibers (a) TiO2-MgO (0.1%) before calcination (b) TiO2-MgO (0.1%) after 500 °C calcination, (c) TiO2-MgO (0.5%) before calcination, (d) TiO2-MgO (0.5%) after 500 °C calcination.
Figure 4. SEM micrographs of the electrospun nanofibers (a) TiO2-MgO (0.1%) before calcination (b) TiO2-MgO (0.1%) after 500 °C calcination, (c) TiO2-MgO (0.5%) before calcination, (d) TiO2-MgO (0.5%) after 500 °C calcination.
Nanomaterials 13 02255 g004
Figure 5. Images of SEM/EDX elemental mapping of Mg-doped fibers (a) Before calcination and (b) After calcination; the insets are the corresponding field emission SEM images. (c) Cross section of the nanofibers after calcination.
Figure 5. Images of SEM/EDX elemental mapping of Mg-doped fibers (a) Before calcination and (b) After calcination; the insets are the corresponding field emission SEM images. (c) Cross section of the nanofibers after calcination.
Nanomaterials 13 02255 g005
Figure 6. XRD patterns show the evolution of the anatase (101 A, 002 A, 200 A, 204 A, 116 A, 101 A) and rutile (110 R, 101 R, 111 R, 211 R, 220 R, 301 R) for (ad) and (012) Mg for (c,d).
Figure 6. XRD patterns show the evolution of the anatase (101 A, 002 A, 200 A, 204 A, 116 A, 101 A) and rutile (110 R, 101 R, 111 R, 211 R, 220 R, 301 R) for (ad) and (012) Mg for (c,d).
Nanomaterials 13 02255 g006
Figure 7. XPS analyses of the Ti 2p for (a) non-doped TiO2, 0.1% Mg-doped TiO2, and 0.5% Mg-doped TiO2 nanofibers. (b) Non-doped TiO2 fitting for (c) 0.1% Mg-doped TiO2 and (d) 0.5% Mg-doped TiO2. (e) Mg 1s binding energy for 0.1% Mg doped TiO2 and (f) Mg 2p MgO binding energy for 0.1% Mg doped TiO2.
Figure 7. XPS analyses of the Ti 2p for (a) non-doped TiO2, 0.1% Mg-doped TiO2, and 0.5% Mg-doped TiO2 nanofibers. (b) Non-doped TiO2 fitting for (c) 0.1% Mg-doped TiO2 and (d) 0.5% Mg-doped TiO2. (e) Mg 1s binding energy for 0.1% Mg doped TiO2 and (f) Mg 2p MgO binding energy for 0.1% Mg doped TiO2.
Nanomaterials 13 02255 g007aNanomaterials 13 02255 g007b
Figure 8. (a) UV-Vis absorbance spectra of TiO2 compact layer (CL), TiO2 nanofibers, and the TiO2 nanofibers with 0.1% and 0.5% Mg doping. (b) Band gap analyzes of the non-doped TiO2, 0.1% Mg-doped TiO2, and 0.5% Mg-doped TiO2 nanofibers.
Figure 8. (a) UV-Vis absorbance spectra of TiO2 compact layer (CL), TiO2 nanofibers, and the TiO2 nanofibers with 0.1% and 0.5% Mg doping. (b) Band gap analyzes of the non-doped TiO2, 0.1% Mg-doped TiO2, and 0.5% Mg-doped TiO2 nanofibers.
Nanomaterials 13 02255 g008
Figure 9. (a) UPS spectra of non-doped TiO2, 0.1% Mg-doped TiO2, and 0.5% Mg-doped TiO2 nanofibers. The left-hand side is the extended valence spectra, and the right-hand side is secondary-electron cut-off. (b) Energy level diagrams of non-doped TiO2 and Mg-doped TiO2 fiber ETM layers.
Figure 9. (a) UPS spectra of non-doped TiO2, 0.1% Mg-doped TiO2, and 0.5% Mg-doped TiO2 nanofibers. The left-hand side is the extended valence spectra, and the right-hand side is secondary-electron cut-off. (b) Energy level diagrams of non-doped TiO2 and Mg-doped TiO2 fiber ETM layers.
Nanomaterials 13 02255 g009
Figure 10. The PL emission intensity of the non-doped TiO2, 0.1% Mg-doped TiO2, and 0.5% Mg-doped TiO2 nanofibers.
Figure 10. The PL emission intensity of the non-doped TiO2, 0.1% Mg-doped TiO2, and 0.5% Mg-doped TiO2 nanofibers.
Nanomaterials 13 02255 g010
Table 1. Mg-doped nanofibers’ diameters before and after two-hour calcination.
Table 1. Mg-doped nanofibers’ diameters before and after two-hour calcination.
Sample NameMgCl2 (%)T (°C)D (nm)Shrinkage (%)
A0.1-249-
B0.150080%68
C0.5-262-
D0.5500133%50
Table 2. (a) The crystallite size of non-doped TiO2 and MgCl2-doped TiO2 films. (b) The crystal size of non-doped TiO2 and MgCl2-doped TiO2 films with details. All films were annealed at 500 °C.
Table 2. (a) The crystallite size of non-doped TiO2 and MgCl2-doped TiO2 films. (b) The crystal size of non-doped TiO2 and MgCl2-doped TiO2 films with details. All films were annealed at 500 °C.
(a)
Annealing Time (h)SamplesFWHMCrystal Size (nm)
3TiO225.250.889.22
2TiO225.300.7610.69
2TiO2 + 0.1% MgCl225.251.236.61
2TiO2 + 0.1% MgCl225.290.7410.86
(b)
Annealing Time & PhaseCrystal SizeAverage Crystal Size
3 h & Rutile27.3710.6714.11
36.0713.54
41.2423.10
54.319.34
56.5922.09
69.185.91
2 h & Rutile27.4215.5517.80
36.0717.99
41.2522.17
54.3415.80
56.6526.28
69.199.03
3 h & Anatase25.259.2212.23
37.7719.51
48.0317.39
55.052.83
2 h & Anatase25.3010.6914.48
36.0716.60
48.0815.81
62.7714.84
2 h & Rutile_0.1%Mg-Doped27.246.656.82
54.433.93
56.9112.45
69.824.26
2 h & Anatase_0.1%Mg-Doped25.256.614.06
37.823.51
47.994.37
62.461.76
2 h & Rutile_0.5%Mg-Doped26.093.386.25
35.9410.10
2 h & Anatase_0.5%Mg-Doped25.2910.864.00
37.540.68
49.660.45
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

Erdogar, K.; Yucel, O.; Oruc, M.E. Investigation of Structural, Morphological, and Optical Properties of Novel Electrospun Mg-Doped TiO2 Nanofibers as an Electron Transport Material for Perovskite Solar Cells. Nanomaterials 2023, 13, 2255. https://doi.org/10.3390/nano13152255

AMA Style

Erdogar K, Yucel O, Oruc ME. Investigation of Structural, Morphological, and Optical Properties of Novel Electrospun Mg-Doped TiO2 Nanofibers as an Electron Transport Material for Perovskite Solar Cells. Nanomaterials. 2023; 13(15):2255. https://doi.org/10.3390/nano13152255

Chicago/Turabian Style

Erdogar, Kubra, Ozgun Yucel, and Muhammed Enes Oruc. 2023. "Investigation of Structural, Morphological, and Optical Properties of Novel Electrospun Mg-Doped TiO2 Nanofibers as an Electron Transport Material for Perovskite Solar Cells" Nanomaterials 13, no. 15: 2255. https://doi.org/10.3390/nano13152255

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