1. Introduction
Perovskite solar cells (PSCs) have emerged as a promising technology owing to their potential to meet the desired characteristics of existing silicon solar cells such as eco-friendly, abundant materials, high cell efficiency, low cost, and large-scale solution manufacturing process. PSCs have also attracted considerable attention owing to their simple and cost-effective manufacturing, coupled with their high solar cell performance. While most research efforts have focused on the development of perovskite materials. Moreover, studies on glass electrodes, electron transport layers (ETLs), and hole transport layers (HTLs) are equally crucial for optimizing PSCs and enabling their application in various devices. The pioneering work on the liquid-state PSCs was initiated by the Miyasaka group [1]. However, the liquid-state PSCs using a liquid electrolyte can lead to the decomposition of the perovskite (CH3NH3PbI3, MAPbI3) structure, resulting in a drop in energy conversion efficiency (ECE). To address this issue, solid-state PSCs were developed, achieving an efficiency of 9.7%. While the MAPbI3 crystals formed from their precursors exhibited a very fast reaction, challenges arose in producing large crystals and uniform film. In order to enhance the crystal sizes of MAPbI3, subsequent research explored the use of TiO2 with a mesoporous (MP) structure [2], giving rise to the concept of a mesoscopic structure [3]. Moreover, PSCs using Al2O3 as the electron transport layer (ETL), despite its lack of conductivity compared to MP-TiO2, achieved an ECE of 12% [4]. By refining the solar cell design, researchers have been able to produce highly efficient PSCs. Various PSC architectures, including mesoscopic [5,6], planar [7,8], bilayer, and inverted PSCs [9,10], have demonstrated efficiency exceeding 20% (now the world record is 34.6% obtained with perovskite/silicon tandem solar cell).
To further enhance the efficiency of PCSs, various techniques have been developed to fabricate high-quality MAPbI3 films, including solution processes [11,12], and vacuum flash-assisted solution processes [13]. However, in mesoscopic PSCs using MP-TiO2 with pore sizes below 50 nm, the separated charges exhibit different electrode arrival carrier mobilities owing to electron transfer from MAPbI3 to MP-TiO2.This leads to charge trapping and detrapping in the mesoscopic structure, resulting in I-V hysteresis. Consequently, considerable efforts have been devoted to achieving high performance and stability, including optimizing the scan direction and speed to minimize the hysteresis of the I-V curve and overcome this challenge. Therefore, low trap density, high conductivity, and mobility are highly sought-after properties for ETLs, which can be achieved by modifying their surface morphology. To address these challenges, research efforts have been directed toward exploring various ETL architectures, including nanorods [14], nanowires [15], nanotubes [16], and inverse opals [17]. Additionally, the introduction of a passivation layer between the transport layer and perovskite layer as well as the surface modification of the perovskite thin film with hydrophobic groups have been used to enhance PSC efficiency and stability. Furthermore, researchers have explored modifying the ETL properties by incorporating other metal ions, such as NiO [18], CuO [19], and SnO2 [20], as well as doping metal oxides with hetero metal ions, including Zn-doped TiO2 [21], Nb-doped TiO2 [22], and Sm-doped TiO2 [23]. Metal ions and metal oxides are particularly promising materials owing to their large bandgap, high conductivity, and reduced sensitivity to ultraviolet (UV) light. These modifications have demonstrated the potential to reduce the hysteresis of the I-V response, leading to improved PSC efficiency.
Incorporating aluminum (Al) into TiO2 materials introduces novel energy levels that influence charge transport characteristics in semiconductors. During the Al doping process, Al3+ ions substitute into TiO2, modifying its electronic structure [24]. Based on the density functional theory (DFT) calculations, it is well known that the holes are strongly localized in the O2p nonbonding orbital of three-coordinated O2− ions. Consequently, dopants hinder carrier mobility by trapping photogenerated carriers at extrinsic defects [25]. We carried out preliminary work for the test of a passivation effect with an Al-doped TiO2 layer as ETL in the various architecture of PSCs and found that the Al-TiO2 layer in a specific cell structure can play an important role as a buffer layer, enhancing the ECE [26]. Replacing the mesoporous TiO2 (MP-TiO2) layer with an insulating material such as Al2O3, led to an increase in open-circuit voltage (Voc). This improvement can be attributed to a reduction in the quasi-Fermi level splitting. Non-stoichiometric TiO2 contains deep electron trap sites, which facilitate charge recombination. By reducing the recombination rate through passivation, the efficiency of PSCs can be enhanced. This indicates that passivation strategies can effectively minimize charge recombination of MAPbI3 and improve device performance, requesting a systematic study about the exact role of the buffer layer.
In this study, therefore, we fabricated the PSCs by incorporating a crystalline Al-doped TiO2 thin film as a buffer layer in a mesoscopic structure to achieve a passivation effect. We hypothesized that this specific PSC design would reduce carrier recombination at the buffer layer/MAPbI3 interface by increasing the HOMO-energy level of the ETL. Consequently, Voc would improve owing to the wider band gap and enhanced conduction band energy level of crystalline Al-doped TiO2 buffer layer, reflecting improvement of charge carrier transport characteristics with different Al-contents. Experimental and theoretical analyses confirmed our hypothesis, with the optimal cell performance obtained using a MAPbI3 layer deposited on a combination of MP-TiO2 and 7 mol% crystalline Al-doped TiO2 buffer layer as the new ETL.
2. Materials and Methods
2.1. Synthesis of Crystalline Al-Doped TiO2 Buffer Layer and Fabrication of PSCs
The detailed synthesis of crystalline Al-doped TiO2 buffer layers and fabrication of PSCs have been previously reported [26]. Briefly, a Ti precursor sol (0.15 M titanium diisopropoxidebis(acetylacetonate) solution) was spin-coated onto a fluorine-doped thin oxide (FTO) substrate to form a compact TiO2 layer (cp-TiO2), which was then dried at 155 °C. Subsequently, a mesoporous TiO2 (MP-TiO2) layer was deposited onto the cp-TiO2/FTO substrate using a TiO2 powder paste. To create crystalline Al-doped TiO2 buffer layers with varying Al-contents, 3, 5, 7, and 10 mol% aluminum isopropoxide (Sigma-Aldrich, Burlington, NJ, USA, 98.00%) precursors were added to the Ti precursor sol solution. After stirring for 3 h, the buffer layers were synthesized using a spin-coating method. The layers were then dried at 125 °C and annealed at 500 °C for 1 h. To obtain the perovskite (MAPbI3) active layer, 1 M PbI2 (476 mg/mL in dimethylformamide (DMF) and CH3NH3I (7mg/mL in isopropanol (IPA)) were sequentially deposited onto the MP-TiO2 layer. The perovskite layer was subsequently dried at 100 °C. The hole transport layer (HTL) was prepared by mixing 28.8 μL of Spiro-MeOTAD (Lumtec) (72.3 mg/mL in chlorobenzene) with 4-tert-butylpyridine (Sigma-Aldrich, 96%). The TSFI stock solution (Sigma-Aldrich, 99.8%) was stirred for 24 h before being deposited onto the perovskite layer. Finally, a gold (Au) electrode was thermally evaporated onto the HTL. The detailed structures of the fabricated PSCs are shown in Figure 1.
2.2. Characterization
To characterize the crystallinity, structure, and crystal phase of the thin films, X-ray diffraction (XRD) measurements were performed using Cu Kα radiation (λ = 1.5416 Å) with a 40 kV beam voltage and a 30 mA beam current. The optical absorbance of the buffer layers was measured in the wavelength range of 300–500 nm using a UV-Vis-NIR spectrometer (Shimadzu, Kyoto, Japan, UV-3600). X-ray photoelectron spectroscopy (XPS) and Raman spectroscopy were subsequently used to analyze the surface species, atomic composition, and functionality. To evaluate the photovoltaic performance of the solar cell devices, photocurrent density–voltage (I-V) curves (SUN 2000, Huawei, Shenzhen, China) were measured using a xenon lamp under an AM 1.5 filter at 100 mW/cm2 illumination in open circuit conditions. The analysis of electrochemical impedance spectroscopy (EIS) was subsequently performed in a frequency range from 1 MHz to 100 mHz using SUN 2000 Instruments under AC voltage with a perturbation amplitude of 10 mV applied during the EIS measurement.
3. Results and Discussion
Four different types of PSCs were designed and fabricated, each incorporating a different combination ofnonporous TiO2, MP-TiO2, and a buffer layer between the active layer and ETL. The detailed fabrication processes and characteristics of these PSCs have been previously reported [26]. As illustrated in Figure 1, this study focuses on the preparation of mesoscopic PSCs with various selective contacts. These architectures demonstrate the selective collection of electrons, which exhibit different mobilities as they traverse the ETL owing to variations in trap sites introduced by the nonporous TiO2, MP-TiO2, and buffer layer. Among the four different device structures depicted in Figure 1, Cell (b) exhibited the highest efficiency owing to relatively high current density (Jsc), open circuit voltage (Voc), and fill factor (FF) (refer to Figure 2a and Table 1). Notably, the Jsc and FF were considerably enhanced compared to the other cells, leading to an improvement in overall solar cell efficiency [26]. In contrast, the performance of Cell (c) and Cell (d) was reduced. Specifically, Cell (c) exhibited a decrease in FF of approximately 20% compared to the other devices, indicating lower stability. The motivation of this study was to understand why the characteristics of solar cells differed considerably.
Upon illumination of the perovskite (MAPbI3) material with sunlight, electron–hole pairs are generated in the material and subsequently injected carriers into the charge transport layer (refer to Figure 3a). Charge separation and injection can occur via two primary pathways: (1) injection of electrons into the ETL [(e−---h+) MAPbI3 perovskite → ecb−(ETL) + h+(MAPbI3 perovskite) (step 1)] and (2) injection of holes into the HTL [(e−---h+) MAPbI3 perovskite → h+(HTL) + e−(MAPbI3 perovskite) (step 2)]. The carriers injected into the charge transport layers can undergo various processes, including exciton annihilation through photoluminescence [(e−---h+) MAPbI3 perovskite → photoluminescence (step 3)] or non-radiative recombination [(e−---h+) MAPbI3 perovskite → non-radiative recombination (step 4)]. Additionally, back charge transfer can occur at the ETL/MAPbI3 layer interface [ecb−(ETL) + h+(MAPbI3 perovskite) → recombination (step 5)] and the MAPbI3 layer/HTL interface [h+(HTL) + e−(MAPbI3 perovskite) → recombination (step 6)]. Finally, charge recombination can take place through the recombination of electrons in the ETL and holes in the HTL [ecb−(ETL) + h+(HTL) → recombination (step 7)]. Owing to the above-mentioned processes [steps (3), (4), (5), (6), and (7)], which can reduce the efficiency of PSCs, it is desirable for the carrier lifetime of MAPbI3 to be prolonged. The implementation of a buffer layer, as depicted in Figure 3b, can decrease the charge recombination rate and enhance the efficiency of PSCs. To mitigate charge recombination in MAPbI3, we used a passivation effect by incorporating an Al-doped TiO2 buffer layer between the CP-TiO2(ETL)/MP-TiO2 interface. The carrier lifetime (or simply the lifetime) generally refers to the duration during which a charge carrier remains free to move, thereby contributing to electrical conduction. Charge carrier lifetime (or effective electron lifetime, τeff) refers to the duration that charge carriers (electrons in the conduction band and holes in the valence band) remain in a free state. It is defined as the average time taken for a minority carrier to recombine and can vary considerably depending on the materials and construction of the semiconductor. The energy released during recombination can manifest as heat or photons (optical recombination, used in LEDs and semiconductor lasers). Consequently, charge carrier lifetime is a critical factor in bipolar transistors and solar cells. The recombination lifetime (τr) especially refers to the time frame for the recombination of injected or photogenerated electrons and holes. It serves as a fundamental parameter in the analysis of semiconductor optoelectronic devices, particularly solar cells. Surface recombination occurs at the top of the solar cell, emphasizing the importance of using material layers with superior surface passivation properties to minimize the impact of light exposure over periods. As illustrated in Figure 1, MAPbI3 possesses two distinct interfaces: the ETL/MAPbI3 interface and the MAPbI3/HTL interface. Minimizing the recombination rate of electrons and holes at each of these interfaces is crucial. Furthermore, it is essential to modulate the energy levels of the conduction and valence bands through techniques such as self-passivation, doping, and interlayer passivation, to enhance device performance.
In this study, our primary objective was to investigate the impact of various ETLs on the electron transfer ability of carriers generated in the active layer. Building upon our previous PL analysis [26], which revealed a reduction in the PL area for Cell (b) compared to other cells, we hypothesized that the rate of carrier recombination was influenced by the material properties, structure, and characteristics of the ETL. These factors can affect both the generation of excitons and the mobility of charge carriers. Our findings suggest that the presence of a buffer layer, such as Al-doped TiO2, at the MAPbI3/ETL interface facilitates the transfer of generated electrons to the electrode. This indicates an increase in carrier lifetime owing to enhanced recombination resistance. Consequently, the electron acceptor ability was enhanced when the PSC was fabricated using an Al-doped TiO2 buffer layer. Conversely, in the Cell (c) and Cell (d) structures, the hindered generation of electrons impedes proper crystal growth. This suggests that in poorly formed crystals, these paration of electrons and holes is compromised, hindering charge injection [steps (1) and (2)], back charge transfer [steps (5) and (6)], and charge recombination [step (7)]. Instead, exciton annihilation [steps (3) and (4)] occurs at a considerably higher rate. These effects adversely impact the recombination rate, ETL functionality, and overall solar cell performance. To check this kind of passivation effect systematically, we first prepared the crystalline Al-doped TiO2 buffer layers with different Al-doping concentrations based on Cell (b). Then, the analysis of X-ray diffraction (XRD), Raman spectroscopy, and X-ray photoelectron spectroscopy (XPS) were sequentially utilized to prove the passivation effect as well as to understand the charge transfer behavior of the PSCs. For the analysis of XRD, we made all structures of perovskite without HTL, and the thickness of whole ETLs was fixed to 100 nm for both pure TiO2 and Al-TiO2 thin films.
Figure 4a shows XRD patterns of both pure crystalline TiO2 and crystalline Al-doped TiO2 thin films with different Al-doping concentrations. All observed peaks corresponding to TiO2 can be attributed to the anatase (A) and rutile (R) phases of TiO2, which were preserved upon the incorporation of Al precursor up to 10 mol%. No additional peaks associated with TiO2 material were detected during the Al precursor doping process. The ionic radii of the Ti4+, Ti3+, and Al3+ ions are 0.61, 0.64, and 0.53 nm, respectively. As Ti3+ ions are replaced with Al3+ ions, the unit cell size of crystalline Al-doped TiO2 decreases compared to TiO2. As we can see from the right-hand side figure of Figure 4a, both the (101) anatase peak and the (110) rutile peak move slightly towards a higher degree as doping proceeds gradually, indicating that the unit cell is to expand its original size. In addition, the (110) rutile peak at 26.8 degreess harpens with increasing Al-doping concentrations up to 10 mol%, while the (101) anatase peak at 25.5 degrees broadens. The crystallite size of Al-doped TiO2, calculated using the Scherrer equation, increases with the Al-doping concentration. Consequently, the incorporation of Al ions introduces stress into the TiO2 lattice during the doping process, resulting in a slight shift in the peak positions toward higher degrees, indicating unit cell expansion. Figure 4b shows the UV-Vis absorption spectra, which exhibit a shift toward shorter wavelengths with increasing Al-doping, a phenomenon known as blue shift. This observation suggests the possibility of an increase in the conduction band energy level of the ETL owing to the blue shift effect. Consequently, the Fermi level in the device may be increased, leading to an enhancement in either Voc or FF. Based on the Tauc-plots, which can be obtained from Figure 4b, moreover, we observed a variation in reducing optical bandgap from 3.20 to 2.24 eV with increasing Al-doping concentration (mol% Al) between 0 and 7.Then, there will be further decreasing the optical bandgap up to 2.07 eV when 10 mol% Al is doped. This means that the conduction band (i.e., HOMO-energy level) of ETLs will be changeable as the Al-doping concentration is changed. In this study, we observed that the HOMO-energy levels of the ETLs gradually increase as the Al-doping content increases.
Figure 5 shows the Raman spectra of the crystalline Al-doped TiO2 samples. Rutile and anatase phases were determined: Rutile peaks at 143 cm−1 (B1g), 447 cm−1 (Eg), 611 cm−1 (A1g); Anatase peaks at 144 cm−1 (Eg), 197 and 640 cm−1 (Eg), 398 and 515 cm−1 (B1g), respectively. Notably, the primary active Raman mode (Eg) of the anatase phase was observed at 144 cm−1, which corresponds to -Ti-Ti vibrations in the octahedral chains [27]. Meanwhile, the active Raman modes reported for aluminum oxide at 384 and 417 cm−1 were not detected, confirming the substitutional incorporation of Al into the TiO2 crystalline structure [27]. Another noteworthy observation is the decrease in the Raman intensity of TiO2 associated with the vibrational modes of both anatase and rutile as the Al-dopant concentrations increase. Upon zooming in on the A(Eg) vibrational mode, the Raman peak intensities exhibit a decreasing trend with the following sequence of Al mol% contents: 0, 3, 5, 7, and 10 mol%, respectively (refer to the enlarged figures on the right-hand side). Additionally, the peak positions shift toward higher wavelengths (similar to the UV-visible result), indicating a relative decrease in bond length in the lattice for both anatase and rutile phases [27]. Furthermore, the effects of Al-dopant incorporation into the TiO2 crystalline structure can be observed through the changing trends of the primary active Raman mode (Eg) in Figure 5. The decrease in peak intensity and shift toward higher wavenumbers (namely bathochromic shift) signify the grain boundary disorder of TiO2 induced by Al-doping [28]. Notably, despite the increasing Al-doping concentration, all Raman peaks associated with the anatase phase persist, while the rutile phase peaks gradually diminish. This suggests that the remaining peak may be attributed to the insertion of Al3+ ions into the TiO2 lattice. Liu et al. obtained similar results and identified a new Raman vibrational peak at approximately 317 cm−1 in their Al-doped TiO2 crystalline materials [28].
Al-doping was further corroborated by XPS. XPS was measured to confirm the changes in the oxidation number of O1s, Ti2p, and Al2s in TiO2 during the 7 mol% Al-doping process. Figure 6 shows high-resolution XP spectra of O1s, Ti2p, and Al2p for doped (red color) and undoped (black color) samples. The Al2p peak (refer to the right-hand side figure) increases peak intensity in the Al (7 mol%)-doped TiO2. However, no Al2p peak was detected in the pure TiO2 sample. In the pure TiO2, the Ti2p3/2 and Ti2p1/2 peaks (refer to the middle figure) are located at binding energies of 458.8 and 464.5 eV, respectively. Upon the incorporation of Al into the TiO2 crystalline lattice, the Ti2p peaks exhibit a slight shift toward lower binding energy, indicating a reduced oxidation state compared toTi4+ in pure TiO2. The primary O1s peak (refer to the left-hand side figure), with an asymmetric shape, is located at a binding energy of 529.0 eV and may correspond to bonding with both Ti and Al. Additionally, an O1s peak with an asymmetric shape at 531.0 eV was identified and attributed to oxygen species in the Al-doped TiO2 buffer layers. Only a minor shift toward a lower O1s binding energy was observed, potentially resulting from the change in electronegativity of the Ti bond from -O-Ti to Al-O-Ti in the crystal structure of the buffer layer. Consistent with previous studies [28,29], the O1s peak observed at approximately 529.0 eV exhibits an asymmetric shape for both pure and doped TiO2, suggesting the presence of defect sites and/or potential surface contamination. This indicates that all our samples may have experienced a degree of contamination during the measurement process (or surface oxidation owing to exposure to air), leading to a chemical shift of up to 0.6 eV in the O1s binding energy.
To check the detailed passivation effect of buffer layers with different Al-doping concentrations, the characteristics of solar cell devices, photocurrent density–voltage (I-V) curves shown in Figure 2b were measured and their measured parameters are summarized in Table 2. As shown in Table 2 and Figure 2b, the Jsc, Voc, and FF parameters exhibit a gradual increase up to an Al-doping concentration of 7 mol%, contributing to an improvement in cell performance from 9.24% to 11.87%. This enhancement can be attributed to the smoother transfer of electrons from the MAPbI3 layer to the electrode through the ETL. In the 7 mol% Al-doping range, the ECE shows a consistent increase, primarily owing to the increase in current density (Jsc). However, beyond 7 mol% Al-doping, the ECE increases. These observations suggest that the electron acceptor ability is enhanced by the Al-doping process below 7 mol%, indicating a reduction in the recombination rate and an increase in the Fermi-energy level with an appropriate amount of Al-doping. Consequently, the crystalline Al-doped TiO2 buffer layer contributes to an improvement in recombination resistance. In general, photovoltaic performance parameters such as short-circuit current density, open-circuit voltage, and photon current conversion efficiency can be extracted from J-V measurements. While J-V measurements provide insights into the photovoltaic performance of solar cells, they do not offer information about photoelectrochemical behavior such as series and charge transfer resistance, electron lifetime, and charge collection efficiency. These parameters can be obtained using EIS. EIS measurements generate the Nyquist plot (see Figure 7a) as well as the Bode plot [refer to Figure 7b], which depict the charge transfer resistance (Rr or Rct) and angular frequency (ωrec) for the solar cell under study.
To prove this, therefore, the EIS analysis for the three fabricated PSCs without and with buffer layers was performed. As we can see in Figure 7a, the EI spectra of PSCs contain two RC arcs. One is related to the contact resistance of the interfaces at high frequency with the same area. The other is attributed to recombination resistance and chemical capacitance (Cμ) of the device at low frequencies [30]. Thus, EIS results can give the values of both charge transfer resistance that can induce with Cμ and recombination resistance at the ETL/MAPbI3 interfaces and MAPbI3/counter electrode in the field of PSC. For example, the reduction in frequency in the EIS measurement generated a higher electron lifetime, which is explained in Equation (5) and Table 3. In this scenario, the relationship is inversely proportional to the maximum frequency. For the analysis of EIS data, both the Bode plot depicted in Figure 7b and a traditional equivalent circuit model depicted in Figure 7c were also utilized in his study. As shown in Figure 7c, a constant photo-generated current source, a series parasitic resistance (i.e., series resistance, Rs), and a parallel parasitic resistance (i.e., parallel resistance, Rp) are also generally included. When comparing the J-V results with the EIS data presented in Figure 7a, it is evident that the PSC with the 10 mol% crystalline Al-doped TiO2 layer exhibits higher Rs and lower recombination resistance compared to the PSC with the 7 mol% crystalline Al-doped TiO2 buffer layer. This indicates a decrease in carrier transport ability owing to a relatively high Fermi-energy level of the ETL, suggesting that electrons encounter difficulty in overcoming the energy barrier. Consequently, exciton annihilation [steps (3) and (4)] and back charge transfer [step (6)] become more likely to occur, while electron injection [step (1)] is hindered. This ultimately leads to electron–hole recombination. As the Al-doping concentration increases, the HOMO-energy level of the ETL gradually increases. The PSC with the 7 mol% crystalline Al-doped TiO2 buffer layer demonstrates the most efficient charge injection [step (1)] and minimal back charge transfer [step (5)] as a result of the passivation effect. This observation confirms that the solar cell parameters with the 7 mol% crystalline Al-doped TiO2 buffer layer were improved, leading to the best cell performance (11.87%) in comparison to the non-doping sample (6.37%). However, the solar cell performance declined with the 10 mol% crystalline Al-doped TiO2 sample. This suggested that the conduction band energy level became excessively high, making it more challenging for electrons to move through the ETL, thereby increasing the recombination rate. To clarify these aspects, we performed theoretical calculations to obtain various electron transport parameters as follows.
Based on the diffusion-recombination transmission line model, the measured impedance in Figure 7a is attributed to the following equations [31,32,33,34].
Z(ω) = [RtRr/(1 + i ω/ωrec)]1/2coth[(Rt/Rr)1/2 (1 + i ω/ωrec)1/2](1)
where Z represents the electrochemical impedance for the photoanode thin film, i = (−1)1/2, ω denotes the angular frequency, ωrec signifies the frequency of the charge transfer process, Rt is the transport resistance of the electrons and Rr is the charge transfer resistance associated with the recombination of electrons at the perovskite/HTL interface. Rr can also be referred to as the charge transfer resistance (Rct) [34]. Additionally, Equation (1) reveals the relationship between two opposing processes that occur in solar cells: electron transfer across the semiconductor layer (ωd) and carrier loss owing to electron recombination (ωrec). These processes can be mathematically expressed as follows:ωd/ωrec = Rct/Rt = (Ln/L)2(2)
ωd = 1/RtCμ = Deff/L2(3)
and ωrec = 1/RctCμ = Deff/Ln2(4)
where Deff represents the effective electron chemical diffusion coefficient, L is the thickness of the photoanode film layer, Ln is the effective electron diffusion length of the photoanode, and Cμ denotes the chemical capacitance, respectively [31,33]. When Rt ≪ Rct or ωrec ≪ ωd, the parallel connection of Rct and Cμ results in the formation of a recombination arc, and Equation (1) can be simplified to the following form:Z(ω) = Rt/3 + [Rct/(1 + i ω/ωrec)](1A)
The small and large arcs of a conventional Warburg impedance appear at high and low frequencies, respectively, as shown in Figure 7a. A large arc corresponds to a low or slower recombination effect compared to electron diffusion through the TiO2 layer. Furthermore, Rt ≪ Rct is a desirable condition for optimal solar cell performance because a large Rct contributes to reduced electron recombination in the cell. Conversely, when Rt ≫ Rct or ωrec ≫ ωd, a considerable combination effect arises in the solar cell owing to the Gerischer impedance, leading to the impedance curve with a smaller arc at high frequency as observed in Figure 7a. Consequently, the recombination time is shorter than the electron diffusion across the TiO2 layer. Based on the above-mentioned model theory, various electron transport parameters, such as the effective electron lifetime (τeff), maximum peak frequency (fmax), and effective rate constant (keff) for recombination, can be extracted from the impedance spectra of solar cells. For instance, τeff is derived from the fmax of the impedance spectra [refer to Figure 7b], which corresponds to the photoanode in the medium frequency region [35]. This maximum frequency also represents the effective rate constant (keff) for recombination. These three parameters are interrelated as follows:
τeff = 1/ωrec = 1/2πfmax = RctCμ(5)
ωrec = keff = 1/τeff(6)
This implies that the effective electron lifetime (τeff) aligns with the angular frequency at the apex of the arc in Figure 7a, expressed as ωrec = 2πfmax = τeff−1, where ωrec denotes the angular frequency of the electron recombination reaction [36]. Equation (6) effectively elucidates the separation of a lifetime into two primary components. The separation occurs because Rrec encompasses both a density term, (the chemical capacitance) and a kinetic constant (keff). Consequently, the product in Equation (6) isolates the latter. The electrochemical capacitance (Cμ), which is linked to minority carrier accumulation, and the recombination resistances together yield a virtually constant lifetime across varying illumination intensities. In the presence of impurity levels, for instance, lifetimes in silicon exhibit high variability [37], which is evenmore pronounced in amorphous inorganic semiconductors such as Al-TiO2. It is crucial to emphasize that recombination constitutes an interfacial charge transfer event occurring at the boundary between the semiconductor and ionic/hole carrier. Given that the distance for electron tunneling is typically approximately 1 nm, which is considerably smaller than the average size of nanoparticles in a solar cell, it is beneficial to differentiate between bulk traps. In any transient measurement used to determine lifetime, a modification of the Fermi level entails a change in the occupancy of bulk traps. Consequently, we can derive the free electron lifetime (Tt) by multiplying two experimentally determined quantities: the diffusion coefficient (Dn) and the effective electron lifetime (τeff). However, in this study, we used the values of the effective electron chemical diffusion coefficient (Deff) as obtained theoretically.
Tt = Dn τeff/Do = Deff τeff/Do(7)
The denominator of Equation (7) represents a constant estimated to be D0 = 0.4 cm2s−1 [38]. It has been well established that the mobility–lifetime product corresponds to the free carrier recombination time in amorphous silicon solar cells. Consequently, the electrochemical capacitance (Cμ) was also calculated using Equation (8), utilizing the values of τeff and Rct extracted from the Nyquist plot.
τeff = 1/ωrec = 1/2πfmax = RctCμ = 1/keff(8)
Hence, the effective electron chemical diffusion coefficient (Deff) and effective electron diffusion length (Ln) of the photoanode can also be derived from Equation (9) [36].
Ln = Deff τeff(9)
This implies that the electron diffusion length (Ln) can be derived from Equations (3) and (4) and is determined by the product of the diffusion coefficient (Deff) and the electron lifetime (τeff). Based on Equation (9), we observe that Ln increases with increasing electron lifetime, assuming that the diffusion coefficient (Deff) remains constant at a fixed temperature. It is also noteworthy to examine the observed properties of Ln in our solar cell devices. We initially note that Ln is not a transient quantity, but rather a steady-state parameter; therefore, it can be expressed in terms of the IS resistances as Ln = L(Rrec/Rt)1/2, where Rt represents the transport resistance [refer to Equation (2)]. Consequently, bulk traps do not interfere with the diffusion length. Notably, the solid hole conductor OMeTAD (HTL) cell exhibits the lowest charge transfer resistance, which considerably impairs performance despite the cell’s attainment of higher photovoltage [39,40]. The model predicts that Ln is proportional to the square root of the free carrier lifetime. In solar cells, the diffusion length typically exhibits a slight increase with increasing bias voltage. For instance, for MP-TiO2 with MP-TiO2/Perovskite/HTL interface, the diffusion length is observed to increase with voltage in the absence of illumination, while a contrasting trend is observed under illumination: it decreases with the potential [38,39,40].
In addition, we can determine the recombination lifetimes (τr) from a photovoltage (Voc) decay curve [41] shown in Figure 8. Figure 8 is the photovoltage decay curve plotting 1/Voc and Jsc as a function of Al-doping amounts in mol%, and realizes a tendency to oppose between τr and τeff (see Table 3). Table 3 summarizes the experimentally and theoretically obtained PSC parameters, including charge transfer resistance (Rct), maximum peak frequency (fmax), effective electron lifetime (τeff), effective rate constant (keff) for recombination, effective electron chemical diffusion coefficient (Deff), electrochemical capacitance (Cμ), free electron lifetime (Tt), effective electron diffusion length (Ln), and recombination lifetime (τr). From Figure 7a,b, we can determine the Rct and ωrec values, and by using Equation (6), ωrec = 2πfmax = τeff−1, we could obtain fmax values as well as τeff values which are listed in the 2nd and 3rd columns, respectively. As we can see in the 3rd column, the obtained τeff values are 3.016 × 10−3 s for 0 mol% Al-doping, 7.271 × 10−3 s for 7 mol% Al-doping, and 4.398 × 10−3 s for 10 mol% Al-doping, respectively. In contrast, the τr values are 1.29 (0 mol%), 0.97 (7 mol%), and 1.12 s (10 mol%), respectively. The increase in τeff facilitates electron diffusion and transfer owing to the increase in Ln, because τr is closely related to both free electron lifetime (Tt) and electron diffusion length (Ln), as well as electrochemical capacitance (Cμ). The response time for recombination, commonly referred to as the “lifetime”, invariably comprises a recombination resistance and a capacitance, and it is directly revealed by IS measurement. However, the time constant of the recombination arc of IS must generally be expressed as τr = RrecCμ [31]. The nature of the electrochemical capacitance (Cμ) is crucial for correctly interpreting τr as a lifetime. Under specific conditions, the transient measurement induces charging or discharging of the depletion layer in a Si-based solar cell, resulting in the measured “lifetime” having no correlation with recombination time. In such cases, the τr values correspond to depletion capacitance and cannot be associated with the minority carrier lifetime. It is important to emphasize that when determining electron lifetime in solar cell devices, it is essential to verify that the recombination resistance is indeed associated with interfacial charge transfer at the nanostructured metal oxide [42]. Carrier recombination lifetimes are often regarded as indicators of perovskite film quality, with longer decay lifetimes being associated with higher-performing materials. Carrier recombination kinetics has been characterized as a combination of trap-assisted, monomolecular (first-order), and bimolecular (second-order) recombination processes. While most studies agree that radiative bimolecular recombination is dominant at high initial carrier densities (n0 > 1017 cm−3), reports on kinetics at lower excitation densities (relevant to solar cell operation) vary from single-exponential to bi-exponential and to stretched-exponential functions with varying degrees of accuracy [43].
As evident from Table 2 and Table 3, our highest performing PSC with a 7 mol% crystalline Al-doped TiO2 buffer layer exhibited the Rct of 258.73Ω along with a low effective rate constant (keff) of 137.533 s−1 for recombination. It also demonstrated a relatively long effective electron lifetime (τeff) of 7.271 ms (accompanied by a maximum electrochemical capacitance value of 28.10 μF) among the various solar cell devices. This is potentially attributable to the suppression of recombination, as indicated by the minimal recombination lifetime (0.97 s) and electron chemical diffusion coefficient (11.77 cm2s−1) at the CP-TiO2(ETL)/Buffer layer(Al-doped TiO2)/MP-TiO2/Perovskite/HTL interfaces. Consequently, a relatively long effective electron diffusion length (Ln) of 85.61 μm is achieved. Mahmud et al. observed that halide perovskite-based solar cells using a nano-mesoporous TiO2 layer exhibited a slower decay rate, which translates to a longer effective electron lifetime, reduced charge recombination with enhanced charge collection efficiency [44]. Similarly, Xiao et al. reported relatively long charge carrier lifetimes under 0.3 sun illumination [45]. Carrier lifetimes in the microsecond range suggest a lack of non-radiative pathways, and such extended carrier lifetimes can fundamentally lead to high Voc. We speculate that the primary factor contributing to the extended carrier lifetimes is a combination of enhanced perovskite crystal grain size and effective passivation achieved through the crystalline Al-doped TiO2 layer [46]. Two plausible explanations can be offered for the increased charge carrier lifetime. First, owing to the crystalline Al-doping of the TiO2 buffer layer located between the pure TiO2 and perovskite layer, the concentration of deep trap states, which serve as recombination centers, diminishes with increasing Al content. Consequently, the effective charge carrier lifetime is prolonged owing to a reduction in Shockley–Read–Hall (SRH) recombination. Second, the incorporation of Al-TiO2 into the perovskite MAPbI3 lattice may introduce additional defects, which are shallow in nature and act as dopants for the MAPbI3 lattice. Unlike deep trap states located in the middle of the band gap, shallow traps positioned near the conduction or valence band of the semiconductor partially trap charge carriers, which must then be released back into the band prior to recombination. In this scenario, up to a certain doping level, the charge carrier lifetime would be extended because thetrapping-and-release events would decelerate the actual recombination process. Yang et al. reported that the CH3NH3Br-treated PSCs exhibited improved charge collection and surface passivation properties, possibly owing to reduced defect states [47].
Based on our comprehensive experimental observations, we propose a plausible mechanism for the Al-doping processes in TiO2 buffer layers as follows. As depicted in Figure 9, when oxygen vacancies or Ti interstitials arise in the TiO2 lattice, Ti ions are preferentially substituted by Al ions in substitutional sites rather than interstitial sites. This substitution triggers the transformation of Ti(IV) into both Ti(III) and Ti(IV)+ + e− forms owing to the presence of defect [refer to Figure 9a]. Subsequently, Ti(III) is replaced by Al(III), which exhibits enhanced stability, leading to changes in optical and electrical properties [refer to Figure 9b]. In a solar cell, an electrical device, semiconductors such as Si, TiO2, and perovskite material are subjected to light, which is converted into electricity via the photovoltaic effect. Electrons are either excited through the absorption of light, or if the material’s bandgap energy can be bridged, electron–hole pairs are generated, simultaneously creating a voltage potential. Consequently, charge carrier and electron diffusion, as well as electron recombination events, frequently occur in a solar cell. The charge carriers in the solar cell move through the semiconductor to counteract the voltage potential, which serves as the drifting force for electron movement. Additionally, electrons can be forced to move by diffusion from areas of higher electron concentration to areas of lower electron concentration. Therefore, to optimize the efficiency of a solar cell, it is highly desirable to collect as many charge carriers as possible at the cell’s electrodes. This implies that electron recombination, which can compromise solar cell efficiency, must be minimized. Consequently, we should prioritize increasing carrier lifetime while maintaining minimal electron recombination, suggesting that steps 1 and 5 in Figure 3 are of greater significance than the other steps. By successfully growing higher-quality crystals with minimal defects, we can achieve considerably higher ECE than current levels. This occurs because efficient electron–hole separation is hindered in a defective crystal lattice. Additionally, steps 3 and 4 in Figure 3 will proceed at a faster rate, which can adversely impact the recombination rate, ETL, and overall solar cell performance.
4. Conclusions
In this study, we introduce crystalline Al-doped TiO2 buffer layers for the fabrication of mesoscopic PSCs with the aim of enhancing photovoltaic performance. Our approach focuses on reducing interfacial resistance and increasing recombination resistance. As a result, the crystalline Al-doped TiO2 buffer layer increases the energy level of the conduction band, facilitating the efficient transport of electrons to the electrode owing to a passivation effect. We achieved a remarkable solar cell efficiency of 11.87% with a PSC structure of Cell (b) using a 7 mol% crystalline Al-doped TiO2 buffer layer. The validity of our approach is corroborated by J-V and EIS analysis, as well as theoretical calculations of various electron transport parameters. This is a remarkable achievement because we have obtained a substantial enhancement in power energy conversion efficiency of nearly 50% from 6.37% [obtained from Cell (c)] to 11.87%, by incorporating the newly developed ETL. With the development of new energy sources that can be synthesized from recycled photocatalytic materials (such as TiO2 powder) at a relatively low cost, the commercialization of PSCs has become a viable prospect.
Conceptualization, D.K. and J.-H.B.; methodology, D.K., J.L. and R.J.; software, J.L. and K.-H.H.; validation, D.K., R.J. and K.-H.H.; formal analysis, D.K. and J.L.; investigation, D.K., J.L. and R.J.; resources, K.-H.H.; data curation, D.K. and J.L.; writing—original draft preparation, D.K.; writing—review and editing, J.-H.B.; visualization, D.K. and R.J.; supervision, J.-H.B.; project administration, J.-H.B.; funding acquisition, J.-H.B. All authors have read and agreed to the published version of the manuscript.
Data are contained within the article.
The authors declare no conflicts of interest.
Footnotes
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.
Figure 1. Device architectures featuring variation in the buffer layer or non-porous TiO2 layer in mesoscopic structure: Cell (a) FTO/CP-TiO2(ETL)/MP-TiO2/Perovskite/HTL/Au, Cell (b) FTO/CP-TiO2(ETL)/Buffer layer (Al-TiO2)/MP-TiO2/Perovskite/HTL/Au, Cell (c) FTO/CP-TiO2(ETL)/NP-TiO2/Perovskite/HTL/Au, and Cell (d) FTO/CP-TiO2(ETL)/Buffer layer(Al-TiO2)/Perovskite/HTL/Au.
Figure 2. J-V curves obtained from (a) 4 different solar cell devices shown in Figure 1, and (b) PSCs fabricated based on Cell (b) structure without and with crystalline Al-doped TiO2 buffer layers.
Figure 3. Schematic diagrams of carrier transfer processes in PSCs (a) without and (b) with buffer layer.
Figure 4. (a) XRD patterns of both pure TiO2 and crystalline Al-doped TiO2 buffer layers. Peaks labeled with “R” correspond to the rutile phase, while peaks labeled with “A” correspond to the anatase phase. The right-hand side figure of (a) presents enlarged images. (b) UV-visible spectra of the crystalline Al-doped TiO2 buffer layers. (The Al mol% content (x) increases with the arrow (0, 3, 5, 7, and 10 mol%), indicating the blue shift.)
Figure 5. Raman spectra of pure TiO2 and Al-doped TiO2 buffer layers. The Al mol% contents are increased with the sequences of 0, 3, 5, 7, and 10 mol%, respectively [see the enlarged figures on the right-hand side for the A(Eg) vibrational mode].
Figure 6. High-resolution XP spectra (O1s, Ti2p, Al2p) of pure TiO2 and Al (7 mol%)-doped TiO2 buffer layer.
Figure 7. (a) Nyquist plot and (b) Bode plot (with log scaled frequency) of mesoscopic PSCs with 0, 7, and 10 mol% crystalline Al-doped TiO2 buffer layers. Figure (c) shows a traditional equivalent circuit model used for the analysis of EIS data.
Figure 8. Variations of 1/Voc and Jsc with different Al-doping concentrations in the TiO2 buffer layer.
Figure 9. Possible mechanism for the Al-doping process in TiO2 buffer layers: (a) Un-doped and (b) Al-doped.
Summary of photovoltaic performance parameters obtained from J–V curves of
Sample Name | Jsc (mA/cm2) | Voc (V) | FF (%) | PCE (%) |
---|---|---|---|---|
Cell (a) | 14.66 | 0.898 | 65.66 | 8.64 |
Cell (b) | 15.12 | 0.916 | 66.65 | 9.24 |
Cell (c) | 12.84 | 0.908 | 54.62 | 6.37 |
Cell (d) | 12.64 | 0.913 | 64.72 | 8.06 |
Summary of photovoltaic performance parameters obtained from J–V curves of
Sample Name | Jsc (mA/cm2) | Voc (V) | FF (%) | ECE (%) |
---|---|---|---|---|
Pure TiO2 | 15.12 | 0.916 | 66.65 | 9.24 |
3 mol% Al | 15.63 | 0.919 | 67.68 | 9.722 |
5 mol% Al | 16.93 | 0.927 | 67.84 | 10.66 |
7 mol% Al | 18.79 | 0.941 | 67.14 | 11.87 |
10 mol% Al | 17.50 | 0.934 | 67.30 | 11.00 |
PSC parameters including charge transfer resistance (Rct), maximum peak frequency (fmax), effective electron lifetime (τeff), effective rate constant (keff) for recombination, effective electron chemical diffusion coefficient (Deff), electrochemical capacitance (Cμ), free electron lifetime (Tt), effective electron diffusion length (Ln),and recombination lifetime (τr) extracted from EIS measurements (
Sample Name | Rct (Ω) | fmax (Hz) | τeff (ms) | keff (s−1) | Deff (cm2s−1) | Cμ (μF) | Tt (s) | Ln (µm) | τr (s) |
---|---|---|---|---|---|---|---|---|---|
Pure TiO2 | 124.30 | 52.79 | 3.016 | 331.565 | 13.98 | 24.26 | 0.1054 | 42.17 | 1.29 |
7mol%Al | 258.73 | 21.90 | 7.271 | 137.533 | 11.77 | 28.10 | 0.2139 | 85.61 | 0.97 |
10mol%Al | 195.20 | 36.21 | 4.398 | 227.376 | 16.95 | 22.53 | 0.1864 | 74.55 | 1.12 |
References
1. Kojima, A.; Teshima, K.; Shirai, Y.; Mitasaka, T. Organometal Halide Perovskites as Visible-Light Sensitizers for Photovoltaic Cells. J. Am. Chem. Soc.; 2009; 131, pp. 6050-6051. [DOI: https://dx.doi.org/10.1021/ja809598r] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/19366264]
2. Buchka, J.; Pellet, N.; Moon, S.J.; Baker, R.H.; Gao, P.; Nazeeruddin, M.K.; Grätzel, M. Sequential deposition as a route to high-performance perovskite-sensitized solar cells. Nature; 2013; 499, pp. 316-319.
3. Im, J.H.; Jang, I.H.; Pellet, N.; Grätzel, M.; Park, N.G. Growth of CH3NH3PbI3 cuboids with controlled size for high-efficiency perovskite solar cells. Nat. Nanotechnol.; 2014; 9, pp. 927-932. [DOI: https://dx.doi.org/10.1038/nnano.2014.181] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/25173829]
4. Lee, M.M.; Teuscher, J.; Miyasaka, T.; Murakami, T.N.; Snaith, H.J. Efficient Hybrid Solar Cells Based on Meso-Super structured Organometal Halide Perovskites. Science; 2012; 338, pp. 643-647. [DOI: https://dx.doi.org/10.1126/science.1228604]
5. Jeon, N.J.; Noh, J.H.; Yang, W.S.; Kim, Y.C.; Ryu, S.C.; Seo, J.W.; Seok, S.I. Compositional engineering of perovskite materials for high-performance solar cells. Nature; 2015; 517, pp. 476-481. [DOI: https://dx.doi.org/10.1038/nature14133] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/25561177]
6. Yang, W.S.; Noh, J.H.; Jeon, N.J.; Kim, Y.C.; Ryu, S.C.; Seo, J.W.; Seok, S.I. High-performance photovoltaic perovskite layers fabricated through intramolecular exchange. Science; 2015; 348, pp. 1234-1237. [DOI: https://dx.doi.org/10.1126/science.aaa9272] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/25999372]
7. Liu, M.; Johnston, M.B.; Snaith, H.J. Efficient planar heterojunction perovskite solar cells by vapour deposition. Nature; 2013; 501, pp. 395-398. [DOI: https://dx.doi.org/10.1038/nature12509] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/24025775]
8. Liu, D.; Kelly, T.L. Perovskite solar cells with a planar heterojunction structure prepared using room-temperature solution processing techniques. Nat. Photonics; 2014; 8, pp. 133-138. [DOI: https://dx.doi.org/10.1038/nphoton.2013.342]
9. 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, pp. 542-546. [DOI: https://dx.doi.org/10.1126/science.1254050]
10. Heo, J.H.; Han, H.J.; Kim, D.; Ahn, T.K.; Im, S.H. Hysteresis-less inverted CH3NH3PbI3 planar perovskite hybrid solar cells with 18.1% power conversion efficiency. Energy Environ. Sci.; 2015; 8, pp. 1602-1608. [DOI: https://dx.doi.org/10.1039/C5EE00120J]
11. Lee, K.; Kim, H.J. Precursor Controlfor Air-Processable Anti-Solvent-Free Perovskite Photovoltaic Cellsin High-Humidity Condition. Appl. Sci. Converg. Technol.; 2024; 33, pp. 126-129. [DOI: https://dx.doi.org/10.5757/ASCT.2024.33.5.126]
12. Hwang, K.-H.; Nam, S.-H.; Kim, D.I.; Seo, H.J.; Boo, J.-H. The influence of DMSO and ether via fast-dipping treatment for a perovskite solar cell. Sol. Energy Mater. Sol. Cells; 2018; 180, pp. 386-395. [DOI: https://dx.doi.org/10.1016/j.solmat.2017.10.024]
13. Li, X.; Bi, D.; Yi, C.Y.; Décoppet, J.D.; Luo, J.; Zakeeruddin, S.M.; Hagfeldt, A.; Grätzel, M. A vacuum flash–assisted solution process for high-efficiency large-area perovskite solar cells. Science; 2016; 353, pp. 58-62. [DOI: https://dx.doi.org/10.1126/science.aaf8060] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/27284168]
14. Kim, H.S.; Lee, J.W.; Yantara, N.; Boix, P.P.; Kulkarni, S.A.; Mhaisalkar, S.; Grätzel, M.; Park, N.G. High Efficiency Solid-State Sensitized Solar Cell-Based on Submicrometer Rutile TiO2 Nanorod and CH3NH3PbI3 Perovskite Sensitizer. Nano Lett.; 2013; 13, pp. 2412-2417. [DOI: https://dx.doi.org/10.1021/nl400286w] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/23672481]
15. Im, J.H.; Luo, J.; Franckevicius, M.; Pellet, N.; Gao, P.; Moehl, T.; Zakeeruddin, S.M.; Nazeeruddin, M.K.; Grätzel, M.; Park, N.G. Nanowire Perovskite Solar Cell. Nano Lett.; 2015; 15, pp. 2120-2126. [DOI: https://dx.doi.org/10.1021/acs.nanolett.5b00046] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/25710268]
16. Wanga, X.; Lia, Z.; Xuc, W.; Kulkarnia, S.A.; Batabyala, S.K.; Zhangd, S.; Caoc, A.; Wong, L.H. TiO2 nanotube arrays based flexible perovskite solar cells with transparent carbon nanotube electrode. Nano Energy; 2015; 11, pp. 728-735. [DOI: https://dx.doi.org/10.1016/j.nanoen.2014.11.042]
17. King, J.S.; Graugnard, E.; Summers, C.J. TiO2 Inverse Opals Fabricated Using Low Temperature Atomic Layer Deposition. Adv. Mater.; 2005; 7, pp. 1010-1013. [DOI: https://dx.doi.org/10.1002/adma.200400648]
18. Hu, L.; Peng, J.; Wang, W.; Xia, Z.; Yuan, J.; Lu, J.; Huang, X.; Ma, W.; Song, H.; Chen, W. et al. Sequential Deposition of CH3NH3PbI3 on Planar NiO Film for Efficient Planar Perovskite Solar Cells. ACS Photonics; 2014; 1, pp. 547-553. [DOI: https://dx.doi.org/10.1021/ph5000067]
19. Zuo, C.; Ding, L. Solution-Processed Cu2O and CuO as Hole Transport Materials for Efficient Perovskite Solar Cells. Small; 2015; 11, pp. 5528-5532. [DOI: https://dx.doi.org/10.1002/smll.201501330]
20. Jiang, Q.; Zhang, L.; Wang, H.; Yang, X.; Meng, J.; Liu, H.; Yin, Z.; Wu, J.; Zhang, X.; You, J. Enhanced electron extraction using SnO2 for high-efficiency planar structure HC(NH2)2PbI3-based perovskite solar cells. Nat. Energy; 2016; 2, 16177. [DOI: https://dx.doi.org/10.1038/nenergy.2016.177]
21. Wu, M.C.; Chan, S.-H.; Jao, M.-H.; Su, W.F. Enhanced short-circuit current density of perovskite solar cells using Zn-doped TiO2 as electron transport layer. Sol. Energy Mater. Sol. Cells; 2016; 57, pp. 447-453. [DOI: https://dx.doi.org/10.1016/j.solmat.2016.07.003]
22. Numata, Y.; Ishikawa, R.; Sanehira, Y.; Kogo, A.; Shirai, H.; Miyasaka, T. Nb-doped amorphous titanium oxide compact layer for formamidinium-based high efficiency perovskite solar cells by low-temperature fabrication. J. Mater. Chem. A; 2018; 6, pp. 9583-9591. [DOI: https://dx.doi.org/10.1039/C8TA02540A]
23. Xiang, Y.; Ma, Z.; Zhuang, J.; Lu, H.; Jia, C.; Luo, J.; Li, H.; Cheng, X. Enhanced Performance for Planar Perovskite Solar Cells with Samarium-Doped TiO2 Compact Electron Transport Layers. J. Phys. Chem. C; 2017; 121, pp. 20150-20157. [DOI: https://dx.doi.org/10.1021/acs.jpcc.7b05880]
24. Wang, H.; He, J.; Boschloo, G.; Lindstrom, H.; Hagfeldt, A.; Lindquist, S.E. Electrochemical Investigation of Trapsin Nanostructured TiO2 Film. J. Phys. Chem. B; 2001; 105, pp. 2529-2533. [DOI: https://dx.doi.org/10.1021/jp0036083]
25. Murashkina, A.; Rudakova, A.V.; Ryabchuk, V.K.; Nikitin, K.V.; Mikhailov, R.V.; Emeline, A.V.; Bahnemann, D.W. Influence of the Dopant Concentration on the Photoelectrochemical Behavior of Al-DopedTiO2. J. Phys. Chem. C; 2018; 122, pp. 7975-7981. [DOI: https://dx.doi.org/10.1021/acs.jpcc.7b12840]
26. Kim, D.I.; Lee, J.W.; Jeong, R.H.; Park, S.; Hwang, K.-H.; Nam, S.-H.; Boo, J.-H. Performance enhancement of perovskite solar cell using Al-TiO2 thin film as electron transporting buffer layer. Energy Proc. CUE2021; 2021; 15, 19.
27. Areerachakul, N.; Sakulkhaemaruethai, S.; Johir, M.A.H.; Vigneswaran, S. Photocatalytic degradation of organicpollutants from wastewater using aluminium doped titanium dioxide. J. Water Process Eng.; 2019; 27, pp. 177-184. [DOI: https://dx.doi.org/10.1016/j.jwpe.2018.12.006]
28. Liu, S.; Liu, G.; Feng, Q. Al-doped TiO2 mesoporous materials: Synthesis and photodegradation properties. J. Porous Mater.; 2009; 17, pp. 197-206. [DOI: https://dx.doi.org/10.1007/s10934-009-9281-8]
29. Murashkina, A.A.; Murzin, P.D.; Rudakova, A.V.; Ryabchuk, V.K.; Emeline, A.V.; Bahnemann, D.W. Influence of the Dopant Concentration on the Photocatalytic Activity: Al-DopedTiO2. J. Phys. Chem. C; 2015; 119, pp. 24695-24703. [DOI: https://dx.doi.org/10.1021/acs.jpcc.5b06252]
30. Dualeh, A.; Moehl, T.; Tetreault, N.; Teuscher, J.; Gao, P.; Nazeeruddin, M.K.; Grätzel, M. Impedance Spectroscopic Analysis of Lead Iodide Perovskite Sensitized Solid-State Solar Cells. ACS Nano; 2014; 8, pp. 362-373. [DOI: https://dx.doi.org/10.1021/nn404323g]
31. Bisquert, J. Theory of the impedance of electron diffusion and recombination in a thin layer. J. Phys. Chem. B; 2002; 106, pp. 325-333. [DOI: https://dx.doi.org/10.1021/jp011941g]
32. Bisquert, J.; Fabregat-Santiago, F.; Mora-Seró, I.; Garcia-Belmonte, G.; Giménez, S. Electron life time in dye-sensitized solarcells: Theory and interpretation of measurements. J. Phys. Chem. C; 2009; 113, pp. 17278-17290. [DOI: https://dx.doi.org/10.1021/jp9037649]
33. Guerrero, A.; Bisquert, J.; Garcia-Belmonte, G. Impedance spectroscopy of metal halide perovskite solar cells from the perspective of equivalent circuits. Chem. Rev.; 2021; 121, pp. 14430-14484. [DOI: https://dx.doi.org/10.1021/acs.chemrev.1c00214] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/34845904]
34. Wang, Q.; Moser, J.-E.; Grätzel, M. Electrochemical Impedance Spectroscopic Analysis of Dye-Sensitized Solar Cells. J. Phys. Chem. B; 2005; 109, pp. 14945-14953. [DOI: https://dx.doi.org/10.1021/jp052768h] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/16852893]
35. Hoshikawa, T.; Ikebe, T.; Kikuchi, R.; Eguchi, K. Effects of electrolyte in dye-sensitized solar cells and evaluation by impedance spectroscopy. Electrochim. Acta; 2006; 51, pp. 5286-5294. [DOI: https://dx.doi.org/10.1016/j.electacta.2006.01.053]
36. Omar, A.; Alia, M.S.; AbdRahim, N. Electron transport properties analysis of titanium dioxide dye-sensitized solarcells (TiO2-DSSCs) based natural dyes using electrochemical impedance spectroscopy concept: A review. Sol. Energy; 2020; 207, pp. 1088-1121. [DOI: https://dx.doi.org/10.1016/j.solener.2020.07.028]
37. McIntosh, K.R.; Paudyal, B.B.; Macdonald, D.H. Generalized procedure to determine the dependence of steady-state photoconductance lifetime on the occupation of multiple defects. J. Appl. Phys.; 2008; 104, 084503. [DOI: https://dx.doi.org/10.1063/1.2999640]
38. Jennings, J.R.; Peter, L.M. A Reappraisal of the Electron Diffusion Length in Solid-State Dye-Sensitized Solar Cells. J. Phys. Chem. C; 2007; 111, pp. 16100-16104. [DOI: https://dx.doi.org/10.1021/jp076457d]
39. Fabregat-Santiago, F.; Bisquert, J.; Cevey, L.; Chen, P.; Wang, M.; Zakeeruddin, S.M.; Grätzel, M. Electron Transport and Recombination in Solid-State Dye Solar Cell with Spiro-OMeTAD as Hole Conductor. J. Am. Chem. Soc.; 2009; 131, pp. 558-562. [DOI: https://dx.doi.org/10.1021/ja805850q]
40. Wang, M.; Chen, P.; Humphry-Baker, R.; Zakeeruddin, S.M.; Grätzel, M. The Influence of Charge Transport and Recombination on the Performance of Dye-Sensitized Solar Cells. Chem. Phys. Chem.; 2009; 10, pp. 290-299. [DOI: https://dx.doi.org/10.1002/cphc.200800708] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/19115326]
41. Pockett, A.; Eperon, G.E.; Peltola, T.; Snaith, H.J.; Walker, A.B.; Peter, L.M.; Cameron, P.J. Characterization of planar lead halide perovskite solar cells by impedance spectroscopy, open circuit photovoltage decay and intensity modulated photovoltage/photocurrent spectroscopy. J. Phys. Chem. C; 2015; 119, pp. 3456-3465. [DOI: https://dx.doi.org/10.1021/jp510837q]
42. Cameron, P.J.; Peter, L.M. How does back-reaction at the conducting glass substrate influence the dynamic photo-voltage response of nanocrystalline dye-sensitized solar cells?. J. Phys. Chem. B; 2005; 109, pp. 7392-7398. [DOI: https://dx.doi.org/10.1021/jp0407270]
43. de Quilettes, D.W.; Vorpahl, S.; Stranks, S.; Nagaoka, H.; Eperon, G.; Ziffer, M.; Snaith, H.J.; Ginger, D. Impact of microstructure on local carrier lifetime in perovskite solar cells. Science; 2015; 348, pp. 683-686. [DOI: https://dx.doi.org/10.1126/science.aaa5333] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/25931446]
44. Mahmud, M.A.; Duong, T.; Peng, J.; Wu, Y.; Shen, H.; Walter, D.; Nguyen, H.T.; Mozaffari, N.; Tabi, G.D.; Catchpole, K.R. et al. Origin of Efficiency and Stability Enhancement in High-Performing Mixed Dimensional 2D-3D Perovskite Solar Cells: A Review. Adv. Funct. Mater.; 2022; 32, 2009164. [DOI: https://dx.doi.org/10.1002/adfm.202009164]
45. Xiao, Z.; Dong, Q.; Bi, C.; Shao, Y.; Yuan, Y.; Huang, J. Solvent annealing of perovskite-induced crystal growth for photovoltaic-device efficiency enhancement. Adv. Mater.; 2014; 26, pp. 6503-6509. [DOI: https://dx.doi.org/10.1002/adma.201401685] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/25158905]
46. Yoo, J.J.; Seo, G.; Chua, M.R.; Park, T.G.; Lu, Y.; Rotermund, F.; Kim, Y.-K.; Moon, C.S.; Jeon, N.J.; Correa-Baena, J.-P. et al. Efficient perovskite solar cells via improved carrier management. Nature; 2021; 590, pp. 587-593. [DOI: https://dx.doi.org/10.1038/s41586-021-03285-w]
47. Yang, M.; Zhang, T.; Schulz, P.; Li, Z.; Li, G.; Kim, D.H.; Guo, N.; Berry, J.J.; Zhu, K.; Zhao, Y. Facile fabrication of large-grain CH3NH3PbI3−xBrx films for high-efficiency solar cells via CH3NH3Br selective Ostwald ripening. Nat. Commun.; 2016; 7, 12305. [DOI: https://dx.doi.org/10.1038/ncomms12305] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/27477212]
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2025 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
Perovskite solar cells (PSCs) characterized by high energy conversion efficiency (ECE) and low manufacturing costs, exhibit promising potential for commercialization in the near term. For commercialization, it is very important to prevent the decomposition of perovskite by ultraviolet (UV) radiation in the air environment. Also, the mesoscopic architecture of PSCs presents considerable opportunities for the solar cell industry, offering potential for recycling of spent photocatalytic materials such as TiO2, and exploration of new energy resources. To solve these problems, therefore, this study introduces a strategy to mitigate these challenges using a crystalline Al-doped TiO2 buffer layer as the electron transport layer (ETL) in conjunction with a mesoporous TiO2 layer in the fabrication of PSCs. Among various Al concentrations in the crystalline Al-doped TiO2 buffer layer fabricated via spin-coating, an optimum concentration of 7 mol% Al yielded the highest cell performance in the specific perovskite solar cell structure. These solar cells exhibited an impressive ECE of 11.87%, representing a substantial enhancement of nearly double the ECE (6.37%) achieved with the conventional ETL. This remarkable improvement can be attributed to the passivation effect of the newly developed ETL, which combines a crystalline Al-doped TiO2 buffer layer with a mesoporousTiO2 layer. Electrochemical impedance spectroscopy (EIS) analysis was performed in conjunction with theoretical calculations of charge transport parameters to substantiate this claim.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details

1 Department of Chemistry, Sungkyunkwan University, Suwon 16419, Republic of Korea;
2 Department of Chemistry, Sungkyunkwan University, Suwon 16419, Republic of Korea;