Next Article in Journal
Scaling Amphiphilicity with Janus Nanoparticles: A New Frontier in Nanomaterials and Interface Science
Previous Article in Journal
Insights into the Adsorption of Carbon Dioxide in Zeolites ITQ-29 and 5A Based on Kinetic Measurements and Molecular Simulations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Efficiency and Mechanical Stability in Flexible Perovskite Solar Cells via Phenethylammonium Iodide Surface Passivation

by
Ibtisam S. Almalki
1,
Tamader H. Alenazi
2,
Lina A. Mansouri
2,
Zainab H. Al Mubarak
2,
Zainab T. Al Nahab
2,
Sultan M. Alenzi
1,
Yahya A. Alzahrani
1,
Ghazal S. Yafi
3,
Abdulmajeed Almutairi
4,
Abdurhman Aldukhail
1,
Bader Alharthi
1,
Abdulaziz Aljuwayr
1,
Faisal S. Alghannam
1,
Anas A. Almuqhim
1,
Huda Alkhaldi
2,
Fawziah Alhajri
2,
Nouf K. AL-Saleem
2,*,
Masfer Alkahtani
1,
Anwar Q. Alanazi
1,* and
Masaud Almalki
1,*
1
Future Energy Technologies Institute, King Abdulaziz City for Science and Technology (KACST), P.O. Box 6086, Riyadh 11442, Saudi Arabia
2
Department of Physics, College of Science and Humanities, Imam Abdulrahman Bin Faisal University, P.O. Box 1982, Jubail 35811, Saudi Arabia
3
Department of Chemistry, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
4
Quantum Technologies & Advanced Computing Institute, King Abdulaziz City for Science and Technology (KACST), P.O. Box 6086, Riyadh 11442, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(14), 1078; https://doi.org/10.3390/nano15141078
Submission received: 21 May 2025 / Revised: 2 July 2025 / Accepted: 9 July 2025 / Published: 11 July 2025
(This article belongs to the Section Solar Energy and Solar Cells)

Abstract

Flexible perovskite solar cells (FPSCs) hold great promise for lightweight and wearable photovoltaics, but improving their efficiency and durability under mechanical stress remains a key challenge. In this work, we fabricate and characterize flexible planar FPSCs on a polyethylene terephthalate (PET). A phenethylammonium iodide (PEAI) surface passivation layer is introduced on the perovskite to form a two-dimensional capping layer, and its impact on device performance and stability is systematically studied. The champion PEAI-passivated flexible device achieves a power conversion efficiency (PCE) of ~16–17%, compared to ~14% for the control device without PEAI. The improvement is primarily due to an increased open-circuit voltage and fill factor, reflecting effective surface defect passivation and improved charge carrier dynamics. Importantly, mechanical bending tests demonstrate robust flexibility: the PEAI-passivated cells retain ~85–90% of their initial efficiency after 700 bending cycles (radius ~5 mm), significantly higher than the ~70% retention of unpassivated cells. This work showcases that integrating a PEAI surface treatment with optimized electron (SnO2) and hole (spiro-OMeTAD) transport layers (ETL and HTL) can simultaneously enhance the efficiency and mechanical durability of FPSCs. These findings pave the way for more reliable and high-performance flexible solar cells for wearable and portable energy applications.

1. Introduction

Flexible perovskite solar cells (FPSCs) have emerged as a cutting-edge photovoltaic technology, offering the PCEs of perovskite materials in lightweight, bendable formats [1,2,3,4,5]. The urgency for renewable and portable energy solutions, driven by global energy demand and the limitations of conventional rigid silicon solar panels, has intensified research into flexible photovoltaics. Rigid perovskite solar cells (PSCs) based on metal halide perovskites (ABX3 structures) have achieved remarkable efficiency gains in the past decade, now rivaling established PV technologies with laboratory PCEs over 26% [6,7]. Early demonstrations of FPSCs, however, showed modest efficiencies; for example, Kumar et al. (2013) reported a PCE of only ~2.6% for a flexible device [8]. Thanks to advances in materials and interfaces, recent flexible devices have reached much higher efficiencies [9,10,11,12,13]. Notably, some state-of-the-art flexible perovskite cells can maintain most of their initial efficiency even after many bending cycles, highlighting substantial progress in mechanical resilience [14]. These trends underscore the potential of FPSCs to become competitive with traditional solar modules, enabling solar power in applications where rigid panels are impractical (wearables, roll-up devices, aerospace, etc.) [15,16,17,18,19].
Despite their promise, FPSCs face unique challenges related to both device architecture and stability. The choice of flexible substrate is critical: common polymer films like polyethylene terephthalate (PET) and polyethylene naphthalate (PEN) offer transparency, light weight, and flexibility, but must withstand the processing temperatures of device fabrication [20,21,22,23,24,25,26,27]. PET can tolerate temperatures up to ~150 °C and serves as a mechanically robust support for the thin-film stack [28,29,30,31,32]. Another challenge is engineering each layer of the solar cell for flexibility without compromising performance. Transparent electrodes, such as indium tin oxide (ITO), are standard in rigid cells but tend to crack under strain; nevertheless, with careful handling, ITO-coated PET can be employed for lab-scale flexible cells. Charge transport layers must also be optimized for the flexible context. Tin oxide (SnO2) has become a popular ETL for high-efficiency PSCs due to its suitable band alignment and low-temperature processability [33,34,35,36,37]. However, differences in surface properties between plastic and rigid substrates demand careful tuning of such ETLs; studies have shown that achieving >15% efficient FPSCs requires ETLs tailored to plastic substrates [38,39]. On the hole transport side, spiro-OMeTAD (doped with lithium bis(trifluoromethanesulfonyl)imide and 4-tert-butylpyridine) remains a benchmark material delivering reliable performance, though its hygroscopic dopants can impact long-term stability [40,41,42].
A particularly important aspect for improving both efficiency and stability is interface engineering of the perovskite active layer. Thin-film halide perovskites are prone to defects at grain boundaries and surfaces, which act as non-radiative recombination centers and initiation points for degradation (e.g., under moisture or bias stress) [43,44,45]. Significant advances in perovskite photovoltaics, including optimized interface engineering and passivation strategies, have driven efficiency improvements. Recent work on coherent interface formation in perovskite solar cells highlights effective strategies to enhance charge extraction and reduce interface defects, directly contributing to higher open-circuit voltages and overall device performance [46]. Additionally, developments in tandem structures incorporating perovskite and organic semiconductors emphasize the importance of versatile interfacial strategies for multi-functional photovoltaic devices [47]. Orientation control and the engineering of thin-film interfaces play critical roles in determining the photovoltaic properties of layered materials. Advanced deposition techniques, such as those demonstrated for oriented Pb-based thin films, have underscored the correlation between crystallographic orientation, interfacial quality, and dielectric properties [48]. Similarly, low-temperature synthesis approaches developed for dielectric oxide films highlight the value of controlled processing conditions in obtaining defect-minimized thin-film materials [49]. Surface passivation strategies using large organic cations have proven effective in addressing these issues [50,51,52]. Phenethylammonium iodide (PEAI) is one such organic salt that has been widely used to form a two-dimensional (2D) perovskite layer at the surface of three-dimensional (3D) perovskite films [53,54,55]. The bulky phenethylammonium (PEA+) cation can terminate the perovskite lattice, creating a thin PEA2PbI4-like 2D capping layer that electronically passivates dangling bonds and defect sites on the surface. This 2D/3D interface has been shown to reduce surface recombination, leading to higher open-circuit voltage and fill factor in devices [56,57,58,59,60]. Additionally, the hydrophobic organic cations can act as a barrier to moisture ingress, improving environmental stability. By healing defects and preventing charge traps, PEAI and similar large-cation post-treatments have enabled notable efficiency improvements and increased operational stability in PSCs. In flexible cells, these benefits could extend further to mechanical stability: a well-passivated, quasi-2D surface may better tolerate strain by preventing crack initiation at grain boundaries. There is growing evidence that such interface layers can enhance mechanical durability; for instance, devices with certain 2D interfaces or additives have retained >90% efficiency after thousands of bending cycles.
The bending stability of flexible perovskite solar cells (FPSCs) is a critical determinant of their practical applicability in portable and wearable electronics. Recent investigations have demonstrated that although well-optimized FPSCs can retain over 90% of their original efficiency after numerous bending cycles, the underlying mechanisms are governed by a complex interplay of mechanical strain, optical modulation, and interfacial delamination [61,62]. For instance, Du et al. showed that when FPSCs are bent to a radius of 90°, they maintain up to 95.12% of their initial power conversion efficiency (PCE), and the bending direction significantly influences device performance with upward bending producing up to 9.9% higher current density than downward bending due to light convergence and enhanced carrier generation near the bending center [63]. Finite element modeling further reveals that stress tends to localize in stiff layers like ITO and Ag, often leading to cracking and electrical failure upon cyclic loading. To mitigate these effects, flexible transparent electrodes, such as PH1000, and engineering approaches, like neutral axis repositioning using protective layers, have been explored. Devices incorporating PH1000 exhibited minimal resistance variation (<14%) and retained over 60% of PCE after 2000 cycles at a tight 2 mm bending radius, in contrast to the catastrophic failure observed in ITO-based counterparts [64]. Such findings emphasize the need for an integrative design approach encompassing material selection, structural reinforcement, and stress-distribution tuning to ensure durable mechanical performance under realistic flexing conditions. While previous studies have extensively addressed PEAI passivation in rigid perovskite systems, detailed investigations into its mechanical and photovoltaic implications on flexible substrates remain limited. Addressing this gap, the current work systematically examines PEAI passivation’s impact on mechanical robustness and photovoltaic performance of flexible perovskite solar cells, thereby contributing practical insights and extending previous interfacial engineering concepts into the flexible device regime [46,47].
In this study, we focus on a flexible perovskite solar cell with a double-cation mixed-halide perovskite absorber (composition (MACl)0.33(FAPbI3)0.97(MAPbBr3)0.1) and investigate the effects of a PEAI surface treatment on its performance and bending stability. To test this, we fabricate devices on PET/ITO substrates with and without the PEAI treatment and carry out comprehensive characterization: structural (XRD, SEM), optical (UV–Vis absorption and photoluminescence, PL), photovoltaic (JV characteristics under simulated sunlight), and mechanical (bending cycle tests). The results demonstrate that PEAI passivation increases the PCE of our flexible cells by mitigating interfacial recombination, and more strikingly, it significantly enhances mechanical durability under cyclic strain. By integrating literature insights and our experimental findings, we discuss how this surface engineering approach can address key bottlenecks in FPSC performance. Our work contributes to the ongoing efforts to develop flexible perovskite photovoltaics that are not only efficient but also robust enough for real-world applications, moving FPSCs closer to commercialization.

2. Methodology

2.1. Device Fabrication

Flexible substrates of PET (Figure 1a) coated with indium tin oxide (ITO) were used as the transparent bottom electrode. The PET/ITO substrates (dimensions ~2.5 cm × 2.5 cm, sheet resistance ~100 Ω/□) were cleaned sequentially with detergent, deionized water, ethanol, and isopropanol, and then dried with nitrogen. A UV-ozone treatment for 15 min was applied to improve surface cleanliness and wettability prior to layer deposition. A compact SnO2 ETL was deposited on the ITO-coated PET via spin coating. We employed a commercially available SnO2 colloidal precursor (2.5% in water) diluted as necessary. The SnO2 solution was dispensed onto the substrate and spun at 4000 rpm (acceleration 2000 rpm/s) for 20 s, yielding a thin, uniform film. The films were then annealed at 150 °C for 60 min on a hotplate to densify the SnO2 layer and enhance its conductivity. This low-temperature process is compatible with PET and results in a transparent n-type layer that facilitates electron extraction from the perovskite. The thickness of the SnO2 ETL is on the order of ~30–50 nm (as observed in cross-section SEM). A double-cation lead mixed-halide perovskite with nominal formula (MACl)0.33(FAPbI3)0.97(MAPbBr3)0.1 was prepared via a one-step spin-coating process in a nitrogen glovebox. The precursor solution was formulated by dissolving the appropriate stoichiometric amounts of methylammonium chloride, methylammonium bromide (MABr), formamidinium iodide (FAI), lead iodide (PbI2), and lead bromide (PbBr2) in a mixed solvent of N,N-dimethylformamide (DMF) and dimethylsulfoxide (DMSO). Specifically, we used a recipe of: 33 mg MACl, 247 mg FAI, 5 mg MABr, 722.4 mg PbI2, and 16.5 mg PbBr2, dissolved in 1 mL of DMF:DMSO (4:1 v/v). The solution was spin-coated in two steps: 2000 rpm for 10 s, followed by 6000 rpm for 30 s. During the second step, an anti-solvent (chlorobenzene, ~200 µL) was dripped onto the spinning substrate ~12 s before the end of the spin cycle. The anti-solvent induces rapid crystallization of the perovskite, resulting in uniform film formation. Immediately after spin coating, the films were annealed at 120 °C for 25 min. For selected samples, a surface passivation layer was applied to the perovskite film after annealing. A PEAI (Figure 1a) solution (6 mg mL−1 in isopropanol) (approximately 0.02 M) was prepared and dynamically spin-coated onto the cooled perovskite film at 3000–4000 rpm for 10–20 s. This step yields two sets of samples: “w/o PEAI” (no passivation) and “with PEAI” (passivated). Next, the hole transport layer was deposited. Spiro-OMeTAD [2,2′,7,7′-tetrakis(N,N-di-p-methoxyphenylamine)-9,9′-spirobifluorene] was used as the HTL, prepared with standard dopants to ensure adequate conductivity. In a nitrogen glovebox, spiro-OMeTAD powder (≈73 mg mL−1) was dissolved in chlorobenzene with additives: lithium bis(trifluoromethanesulfonyl)imide (Li-TFSI) solution (520 mg mL−1 in acetonitrile, 21.36 µL added) and 4-tert-butylpyridine (tBP, 34.5 µL added). The resulting doped spiro-OMeTAD solution (~70 mM spiro) was spin-coated onto the perovskite (or PEAI-treated perovskite) at 4000 rpm for 20 s. This produces a ~150–200 nm thick transparent HTL. Finally, a gold back contact was deposited to complete the device. Gold (~90 nm thick) was thermally evaporated under high vacuum (<10−5 Torr) onto the spiro-OMeTAD layer through a shadow mask. A schematic representation of the full device is illustrated in Figure 1b.

2.2. Characterization Techniques

X-ray diffraction (XRD): The crystalline structure of the perovskite films (with and without PEAI) was analyzed by XRD using a Bragg–Brentano θ-2θ geometry (Cu Kα radiation, λ = 1.5406 Å). Scans were recorded from 10° to 40° 2θ at a step of 0.02° to identify the prominent crystallographic peaks of the perovskite.
X-ray photoelectron spectroscopy (XPS) was performed using monochromatic Al Kα radiation (1486.6 eV) to investigate the chemical interactions at the perovskite/PEAI interface, particularly assessing changes in binding energies and chemical environments associated with surface passivation. Spectra were calibrated using the adventitious carbon peak (C 1s) at 284.8 eV as the internal reference. High-resolution scans were collected for Pb 4f, I 3d, and N 1s core levels to elucidate the chemical state modifications resulting from the PEAI treatment.
Ultraviolet Photoelectron Spectroscopy (UPS) measurements were conducted to investigate the energy-level alignment and work function changes at the perovskite/charge transport interfaces. A He I photon source (21.22 eV) was used, and the spectra were acquired under ultra-high vacuum conditions (~10−8 Torr). The work function and valence band maximum (VBM) positions were determined from the secondary electron cutoff and valence band edge onset, respectively.
Scanning electron microscopy (SEM): The morphology of the perovskite layers was observed by SEM. Top-view SEM images were taken to evaluate grain size, surface coverage, and the presence of pinholes or defects. Cross-sectional SEM images were also obtained by cleaving devices and imaging the layer stack at a tilted angle.
Optical measurements: UV–Vis absorbance spectra were recorded for the perovskite films (on glass or PET/ITO substrates) using a UV–Vis spectrophotometer in the 400–800 nm wavelength range. The absorption onset (band edge) provides an estimate of the perovskite bandgap. Photoluminescence (PL) spectra were measured using a fluorescence spectrometer (or a custom setup with a 532 nm excitation laser and CCD detector) in ambient conditions. Steady-state PL was collected from films with and without PEAI to gauge differences in radiative recombination.
Current–voltage (JV) characteristics: The photovoltaic performance of completed devices was measured under simulated AM 1.5G sunlight. A solar simulator with a 100 mW cm−2 illumination (calibrated with a silicon reference cell) was used. The JV curves were recorded using a source meter (Keithley 2400) by sweeping the voltage from 0 V to ~1.2 V with a scan rate of ~100 mV/s−1 with a delay time of 10 ms at each point. The cells were masked to a defined active area (~0.16 cm2) to avoid edge effects and ensure accurate current density calculations. We report the champion and average PCE values. All measurements were performed in ambient air at ~25 °C and relative humidity < 30% to minimize moisture-related degradation during testing.
Mechanical bending test: To evaluate mechanical stability, a cyclic bending test was carried out on the flexible devices. Each device was repeatedly bent inward (concave facing the perovskite side) to a bending radius of approximately 5 mm (unless otherwise noted) and released, constituting one cycle, and then the device was carefully placed back in the solar simulator and its JV curve measured under 1-sun illumination.

3. Results and Discussion

3.1. Structural Properties (XRD, XPS, and SEM)

XRD analysis confirmed that the films crystalized into the expected perovskite structure with high phase purity (Figure 1c). Both PEAI-passivated and control perovskite films showed dominant diffraction peaks at 2θ ≈ 14.0° and 28.4°, corresponding to the (110) and (220) lattice planes of the tetragonal perovskite crystal (in pseudocubic indexing, these relate to the (100) and (200) reflections) [65]. Additional peaks at ~23.5°, 31.8°, and 40.5° were observed, assignable to higher-order perovskite reflections ((211), (310), and (224) planes, respectively), indicating good crystallinity. Crucially, the XRD patterns did not exhibit a significant peak at 12.6° that would indicate the presence of hexagonal PbI2 impurity. In the control (unpassivated) film, a very weak shoulder at approximately ~12.6° could be discerned, suggesting trace amounts of residual PbI2 or slight decomposition at grain boundaries. In contrast, the PEAI-treated film’s XRD showed this feature to be suppressed below the noise level, implying that the PEAI treatment helped eliminate residual PbI2 by reacting with under-coordinated lead at the surface to form PEA2PbI4 [66]. The XRD pattern of the PEAI-treated film exhibits low-angle reflections characteristic of PEA2PbI4, confirming the formation of an n = 1 Ruddlesden–Popper (RP) phase at the perovskite surface (Figure S1). The overall intensity of the perovskite peaks was slightly higher (by ~10%) for the PEAI-passivated sample, and the FWHM of the main (110) peak was marginally narrower. Overall, the PEAI layer significantly enhanced the crystallographic orientation, as evidenced by an intensity ratio increase of approximately 3.17 for the dominant (110) peak, accompanied by a minor peak shift of about −0.02° in 2θ, suggesting slight structural adjustments likely due to surface interactions or lattice strain induced by PEAI (Table S1). These differences suggest that the surface passivation may also lead to improved crystallinity or larger effective grain size, possibly because the PEAI aids in healing surface imperfections and promotes better ordering at the film surface. Complementary XPS analysis further supported the mechanism of PEAI-induced passivation. The Pb 4f XPS spectra revealed a pronounced reduction in the intensity of metallic lead (Pb0) signals after PEAI treatment, highlighting the effective passivation of Pb-related defect states. This interaction significantly decreases the density of non-radiative recombination centers, thus contributing to the observed improvements in device performance and crystallinity enhancement.
Top-view SEM images provide complementary information on the perovskite film surface. The control perovskite films exhibited grain sizes on the order of 200–500 nm with full surface coverage (no pinholes), consistent with high-quality solution processing (Figure 1d). The grain boundaries are moderately visible as thin lines, and some small PbI2 crystals could be occasionally observed at triple junctions of grain boundaries in the control film. In the PEAI-passivated film, the surface looked smoother and grain boundaries were somewhat less pronounced (Figure 1e). This could be because the PEAI post-treatment forms a thin overlayer that covers the grain boundaries. Indeed, the literature reports have noted that organic cation treatments like PEAI can lead to slightly enlarged apparent grain size and a smoother topography due to the formation of a 2D perovskite at the surface that “fills in” surface gaps [67]. In our case, the difference in grain structure was in favor of the PEAI-treated surface, which appeared to have a larger grain size that was free of any small crystallites or residues and had a uniform contrast in SEM, implying a homogeneous capping layer. The grain size distribution analysis demonstrates a clear increase in the average grain size from approximately 0.6 µm in the reference film to approximately 1.2 µm following PEAI treatment (Figure S2). This enlargement and narrower distribution suggest that PEAI effectively promotes grain growth, leading to enhanced morphological uniformity and potentially improved photovoltaic performance. These morphological observations support the XRD findings that PEAI passivation produces a cleaner, possibly more ordered perovskite surface. Such a surface is expected to have fewer defect sites, which correlates with the improved electronic and optical properties discussed below. Cross-sectional SEM image of a flexible perovskite solar cell (with PEAI passivation), showing the layer structure on the PET/ITO substrate (Figure 1f). The bright top layer is the evaporated Au contact, beneath which the darker spiro-OMeTAD (HTL) and perovskite layers are visible. The SnO2 ETL is too thin to clearly distinguish at this scale, but it resides between the perovskite and the bottom ITO electrode (on PET). The total thickness of the device stack (excluding PET) is on the order of 800–900 nm, with the perovskite layer being ~500 nm thick, as indicated by the scale bar. The layers appear conformal and continuous, with no large voids or delamination gaps observed. Notably, the perovskite layer in the PEAI-treated device exhibits a dense and uniform morphology, which is favorable for charge transport.

3.2. Optical Properties (UV–Vis and PL)

The optical absorption spectra of the perovskite films on glass (both with and without PEAI) are nearly identical in the visible range, indicating that the bulk perovskite remains unchanged by the surface treatment (Figure 2a). The films show a strong absorption onset at approximately 820 nm, corresponding to a bandgap of approximately 1.51–1.55 eV. In the wavelength range of 400–750 nm, the absorbance exceeds 80%, demonstrating the perovskite’s excellent light-harvesting capability over the visible spectrum. Both types of films exhibit a sharp band edge, indicating low Urbach energy (disorder) and high material quality [68]. Photoluminescence measurements, however, reveal a clear difference between passivated and non-passivated films (Figure 2b). Under identical excitation conditions, the PEAI-treated perovskite film exhibits a significantly stronger PL emission at ~800 nm (near the band-edge peak) compared to the untreated film. The integrated PL intensity of the PEAI-passivated sample is roughly double that of the control. This enhancement in PL indicates a lower density of non-radiative recombination centers in the presence of the PEAI layer. Surface defects, which act as traps for charge carriers and quench PL, have been effectively passivated by the phenethylammonium cations. The optical characterization provides evidence that the PEAI treatment passivates surface defects, as seen by the boosted PL intensity. This is a strong indication that solar cells made from these films should exhibit higher open-circuit voltage and reduced recombination losses, since non-radiative recombination at surfaces is a major factor limiting VOC in perovskite solar cells [55,69,70].

3.3. Photovoltaic Performance (J–V Characteristics)

The photovoltaic performance of flexible perovskite devices was measured under standard test conditions (1 sun, AM 1.5G). Without the PEAI passivation, the control FPSCs yielded an average PCE of approximately 13–14%, with the best cell reaching 14.5% (Figure 3a). These control devices typically showed a VOC of ~1.00 V, a Jsc of ~24.8 mA cm−2, and an FF of ~0.60–0.65. The somewhat modest VOC (∼1.0 V) and FF (60–65%) in the control devices indicate there is room for reducing recombination and resistive losses, precisely the issues that surface passivation aims to address. Introducing the PEAI surface treatment led to a notable improvement in device metrics. The PEAI-passivated FPSCs achieved PCEs in the range of 15–17%, with champion devices at ~16.5% PCE under forward-scan illumination. Figure 3b shows representative JV curves comparing a control device and a PEAI-passivated device. The most pronounced change is an increase in VOC for the PEAI device: VOC improved to ~1.08 V (an increase of ~80 mV relative to control). This directly reflects the reduced recombination at the perovskite/HTL interface due to defect passivation, allowing the device to sustain a higher photovoltage. The fill factor also improved, reaching ~70% in passivated cells. Meanwhile, the Jsc of the devices remained high and comparable between passivated and control cells. This indicates that the PEAI layer is not impeding charge collection or optical absorption; carriers can still efficiently reach the contact electrodes. External quantum efficiency (EQE) spectra and integrated current density of reference and PEAI-passivated devices. The PEAI-treated device shows a higher EQE across most of the visible range, with an integrated Jsc of ~25 mA cm−2, consistent with the JV measurements (Figure S3). Overall, the average PCE gain from PEAI treatment was on the order of 2–3 percentage points absolute. This relative improvement (~20%) aligns well with previous reports of organic halide salt passivation in PSCs [44,71]. The improved performance of the PEAI-passivated flexible perovskite solar cells (FPSCs) can be attributed in part to enhanced radiative recombination characteristics, as evidenced by the time-resolved photoluminescence (TRPL) and photoluminescence quantum yield (PLQY) results. As shown in Figure S4, the PEAI-treated films exhibit a notably shorter carrier lifetime (τ = 12.5 ns) compared to the reference (τ = 28 ns), indicating accelerated radiative recombination due to defect passivation. Additionally, the PLQY increases from 4% to 15%, confirming that a greater fraction of photoexcited carriers recombine radiatively rather than through non-radiative pathways. These findings suggest that PEAI effectively suppresses surface traps and enhances interfacial quality, which translates to higher open-circuit voltage and improved device efficiency in the flexible architecture. While our TRPL and PLQY measurements clearly demonstrate reduced non-radiative recombination through PEAI passivation, further detailed studies assessing recombination dynamics under mechanical stress (e.g., post-bending conditions) are warranted. These investigations would enhance understanding of the mechanistic relationship between mechanical stability and electronic properties in flexible perovskite solar cells. Furthermore, the PEAI treatment significantly affects the energy level alignment at the perovskite interface, as evidenced by Ultraviolet Photoelectron Spectroscopy (UPS) analysis (Figure S5). Specifically, the introduction of PEAI shifted the perovskite film’s Fermi level (EF) deeper from 4.51 eV (reference film) to 4.70 eV (PEAI-passivated film). Correspondingly, the valence band maximum (VBM) moved closer to EF, decreasing the offset from 1.01 eV to 0.72 eV. These changes result in improved alignment with the HOMO level of Spiro-OMeTAD (HTL), facilitating more efficient hole extraction and reducing energy loss at the perovskite/HTL interface. This improved energy alignment helps explain the observed enhancements in photovoltaic device performance, particularly the increased Voc and FF.

3.4. Mechanical Flexibility and Stability

A primary advantage of FPSCs is their ability to endure bending, enabling deployment in wearable and portable formats. We evaluated the mechanical robustness of our devices by subjecting them to repeated bending and measuring the retention of photovoltaic performance. The initial (unbent) efficiencies have been described above (≈16.5% with PEAI, ≈14% without). After bending the devices (radius ~5 mm) for 700 cycles, the cells were tested again under the same illumination conditions. Strikingly, the PEAI-passivated devices retained approximately 88% of their original PCE on average after 700 cycles. In contrast, the control devices (no PEAI) retained only about 70% of their initial PCE after the same bending treatment. This difference highlights a substantial improvement in mechanical durability imparted by the PEAI surface layer. Specifically, a typical PEAI-treated device degraded from 16.0% initial PCE down to ~14.0% after 700 bends, whereas a control device might go from 14.0% to ~9.8% in the same test. The dominant loss in both cases was a reduction in the fill factor and a slight drop in Voc, while Jsc remained fairly stable (within 5% of the initial). While 700 cycles are a considerable stress test (far exceeding the bending a typical wearable device might undergo daily), we note that literature reports have demonstrated FPSCs enduring >1000 or even >10,000 bending cycles with specialized architectures. For instance, Li et al. reported flexible cells retaining over 90% of initial efficiency after 1000 bends, and other groups using reinforcing structures have achieved stable performance over 10,000 cycles [72]. Our results, reaching ~88% after 700 cycles without any encapsulation or special structural enhancements, are in line with these findings and confirm that surface passivation is a viable strategy to improve mechanical resilience. Figure 3d shows a photograph of a flexible perovskite solar cell during a bending test (device curvature radius ~5 mm). The cell is bent along the PET substrate, illustrating its mechanical flexibility. Even under significant bending, the layered structure (PET/ITO/SnO2/perovskite/spiro/Au) remains intact without visible cracking. The incorporation of the PEAI passivation layer contributes to this robustness by forming a flexible 2D interface that can accommodate strain. The PEAI at the surface likely acts to “bridge” grain boundaries and adhere the perovskite crystals together, reducing the likelihood of mechanical fracture under flexing [73]. Additionally, PEA+ cations might impart some ductility at the interface, as 2D perovskites are known to be more pliable compared to 3D perovskites [74]. The outcome is an FPSC that not only performs well initially but also shows enhanced longevity under repetitive bending, a critical attribute for any practical deployment of solar cells on flexible platforms. Figure 3 underlines the potential of such devices for wearable technology, for example, solar textiles or device-integrated solar chargers, where constant movement and bending are expected. The visual integrity of the cell in the bent state correlates with the electrical data, indicating sustained performance. The thermal stability of devices was rigorously examined to assess the efficacy of the PEAI passivation layer under prolonged heating conditions (60 °C) at a relative humidity of ~ 25%. Over a period of 500 h, PEAI-treated devices demonstrated significantly enhanced stability, maintaining roughly 85% of their initial photovoltaic performance, as illustrated in Figure S6. In contrast, untreated control devices exhibited pronounced efficiency loss, retaining only approximately 60% of their original efficiency under identical conditions. This marked difference underscores the effectiveness of PEAI in stabilizing the perovskite interfaces against thermal degradation. Variations across five individual devices per condition were consistently minor, as indicated by the small standard deviations, highlighting the reproducibility and robustness of the PEAI passivation strategy. Overall, the developed flexible perovskite solar cells demonstrated promising mechanical stability, enduring extensive bending tests exceeding 700 cycles at a bending radius of 5 mm without significant performance degradation. This performance surpasses or is comparable to many recent state-of-the-art flexible perovskite devices, highlighting the effectiveness of the implemented PEAI passivation strategy in enhancing both structural robustness and reliability for flexible device applications (Table S2). Compared to recently reported passivation methods such as PMMA, PEABr, BAI, and OAI, our PEAI-treated flexible PSCs demonstrate superior photovoltaic efficiency (17%) and notable mechanical resilience, retaining approximately 88% of their initial efficiency after 700 bending cycles (Table S3). The impressive mechanical robustness demonstrated by PEAI-treated perovskite solar cells under bending conditions is promising for flexible photovoltaic applications. The enhancement in mechanical stability likely originates from the microstructural improvements facilitated by PEAI treatment, as evidenced by larger grain sizes and reduced grain boundary defects, contributing to minimized stress concentration points and crack initiation sites [75,76]. Furthermore, the Ruddlesden–Popper phase generated via PEAI passivation provides additional resistance to moisture ingress and environmental degradation, thus potentially extending device lifetime under operational stresses [77]. Integration of these mechanically robust and microstructurally optimized films into practical device architectures requires attention to interface engineering and encapsulation strategies, where optimal alignment of energy levels and minimized interfacial recombination are critical to achieving high device performance and durability under realistic environmental conditions [78,79].

4. Conclusions

We have developed and investigated a flexible perovskite solar cell based on a PET/ITO substrate incorporating a PEAI surface passivation layer. The inclusion of the PEAI passivation layer significantly improves the photovoltaic performance of the flexible cells. PEAI-treated devices achieved PCEs at approximately 16–17%, compared to ~14% for identical devices without PEAI. The improvement is attributed to the effective passivation of surface defects by PEAI, which suppresses nonradiative recombination and yields a more ideal diode behavior. Our results demonstrate that even on flexible substrates, where interfaces may be more defect-prone, applying a 2D cation passivation strategy can push device efficiencies closer to those of rigid, state-of-the-art PSCs. A key outcome of this work is the demonstrated improvement in mechanical stability for PEAI-passivated FPSCs. In bending tests, devices with PEAI retained nearly 90% of their initial efficiency after 700 bending cycles, significantly outperforming unpassivated control devices, which fell to ~70% of their initial efficiency. Improved efficiency and bending durability via a simple surface treatment, which fits well into the larger efforts of making perovskite solar cells more viable in unconventional scenarios. Our findings concur with the growing consensus in the perovskite research community that interface engineering (through chemical means like self-assembled monolayers or 2D cations) is crucial for both achieving high performance and mitigating degradation mechanisms. By demonstrating this on flexible cells, we bridge the gap between materials chemistry and mechanical engineering. The results encourage further interdisciplinary work to optimize the composition of both the perovskite absorber and the flexible substrates/encapsulants to ultimately produce a flexible solar cell that rivals rigid ones in efficiency and exceeds them in functionality.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano15141078/s1.

Author Contributions

Conceptualization, A.Q.A. and M.A. (Masud Almalki); Data curation, T.H.A., L.A.M., Z.H.A.M., Z.T.A.N., Y.A.A. and A.A. (Abdulmajeed Almutairi); Formal analysis, S.M.A., G.S.Y., A.A. (Abdurhman Aldukhail), B.A., A.A. (Abdulaziz Aljuwayr) and F.S.A.; Funding acquisition, N.K.A.-S.; Investigation, I.S.A., T.H.A., L.A.M., Z.H.A.M., Z.T.A.N., A.A. (Abdulmajeed Almutairi), A.A. (Abdurhman Aldukhail), A.A.A. and F.A.; Methodology, I.S.A., S.M.A. and F.S.A.; Project administration, A.Q.A. and M.A. (Masud Almalki); Resources, A.A.A.; Supervision, N.K.A.-S. and M.A. (Masud Almalki); Validation, G.S.Y.; Visualization, Y.A.A., B.A. and H.A.; Writing—original draft, H.A.; Writing—review and editing, M.A. (Masfer Alkahtani), A.Q.A. and M.A. (Masud Almalki). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deanship of Scientific Research at Imam Abdulrahman Bin Faisal University (IAU), Saudi Arabia.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Material. Further inquiries can be directed to the corresponding authors.

Acknowledgments

The authors gratefully acknowledge the support provided by the Deanship of Scientific Research at Imam Abdulrahman Bin Faisal University (IAU), Saudi Arabia, which enabled the successful execution of this study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Han, G.S.; Jung, H.S.; Park, N.G. Recent cutting-edge strategies for flexible perovskite solar cells toward commercialization. Chem. Commun. 2021, 57, 11604–11612. [Google Scholar] [CrossRef] [PubMed]
  2. Subudhi, P.; Punetha, D. Progress, challenges, and perspectives on polymer substrates for emerging flexible solar cells: A holistic panoramic review. Prog. Photovolt. Res. Appl. 2023, 31, 753–789. [Google Scholar] [CrossRef]
  3. Fan, Z.; Zanelli, A.; Monticelli, C.; Li, Q. Flexible Photovoltaic Solar Design. In Lightweight Energy: Membrane Architecture Exploiting Natural Renewable Resources; Springer International Publishing: Cham, Switzerland, 2022; pp. 93–149. [Google Scholar]
  4. Demchyshyn, D.I.S. Flexible Perovskite Photovoltaics for Biomedical and Aerospace Applications; Johannes Kepler University Linz: Linz, Austria, 2023. [Google Scholar]
  5. Zanelli, A.; Monticelli, C. Life cycle assessment of organic solar cells and perovskite solar cells with graphene transparent electrodes. J. Renew. Energy 2022, 195, 906–917. [Google Scholar]
  6. Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal halide perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc. 2009, 131, 6050–6051. [Google Scholar] [CrossRef]
  7. Best Research-Cell Efficiency Chart|Photovoltaic Research|NREL. Available online: https://www.nrel.gov/pv/cell-efficiency.html (accessed on 12 June 2025).
  8. Caprihan, R.; Kumar, A.; Stecke, K.E. Evaluation of the impact of information delays on flexible manufacturing systems performance in dynamic scheduling environments. Int. J. Adv. Manuf. Technol. 2013, 67, 311–338. [Google Scholar] [CrossRef]
  9. Goje, A.A.; Ludin, N.A.; Fahsyar, P.N.A.; Syafiq, U.; Chelvanathan, P.; Syakirin, A.D.A.G.; Teridi, M.A.; Ibrahim, M.A.; Su’ait, M.S.; Sepeai, S.; et al. Review of flexible perovskite solar cells for indoor and outdoor applications. Mater. Renew. Sustain. Energy 2024, 13, 155–179. [Google Scholar] [CrossRef]
  10. Xu, R.; Pan, F.; Chen, J.; Li, J.; Yang, Y.; Sun, Y.; Zhu, X.; Li, P.; Cao, X.; Xi, J.; et al. Optimizing the Buried Interface in Flexible Perovskite Solar Cells to Achieve Over 24% Efficiency and Long-Term Stability. Adv. Mater. 2024, 36, 2308039. [Google Scholar] [CrossRef]
  11. Wu, Y.; Xu, G.; Shen, Y.; Wu, X.; Tang, X.; Han, C.; Chen, Y.; Yang, F.; Chen, H.; Li, Y.; et al. Stereoscopic polymer network for developing mechanically robust flexible perovskite solar cells with an efficiency approaching 25%. Adv. Mater. 2024, 36, 2403531. [Google Scholar] [CrossRef]
  12. Liu, P.; Wang, H.; Niu, T.; Yin, L.; Du, Y.; Lang, L.; Zhang, Z.; Tu, Y.; Liu, X.; Chen, X.; et al. Ambient scalable fabrication of high-performance flexible perovskite solar cells. Energy Environ. Sci 2024, 17, 7069–7080. [Google Scholar] [CrossRef]
  13. Li, Z.; Jia, C.; Wan, Z.; Cao, J.; Shi, J.; Xue, J.; Liu, X.; Wu, H.; Xiao, C.; Li, C.; et al. Boosting mechanical durability under high humidity by bioinspired multisite polymer for high-efficiency flexible perovskite solar cells. Nat. Commun. 2025, 16, 1771. [Google Scholar] [CrossRef]
  14. Song, F.; Zheng, D.; Feng, J.; Liu, J.; Ye, T.; Li, Z.; Wang, K.; Liu, S.F.; Yang, D. Mechanical durability and flexibility in perovskite photovoltaics: Advancements and applications. Adv. Mater. 2024, 36, e2312041. [Google Scholar] [CrossRef] [PubMed]
  15. Yang, D.; Yang, R.; Priya, S.; Liu, S. Recent advances in flexible perovskite solar cells: Fabrication and applications. Angew. Chem. Int. Ed. 2019, 58, 4466–4483. [Google Scholar] [CrossRef] [PubMed]
  16. Wang, Z.; Zhu, X.; Feng, J.; Yang, D.; Liu, S. Semitransparent flexible perovskite solar cells for potential greenhouse applications. Sol. RRL 2021, 5, 2100264. [Google Scholar] [CrossRef]
  17. Wojciechowski, K.; Forgács, D. Commercial applications of indoor photovoltaics based on flexible perovskite solar cells. ACS Energy Lett. 2022, 7, 3729–3733. [Google Scholar] [CrossRef]
  18. Gao, Y.; Huang, K.; Long, C.; Ding, Y.; Chang, J.; Zhang, D.; Etgar, L.; Liu, M.; Zhang, J.; Yang, J. Flexible perovskite solar cells: From materials and device architectures to applications. ACS Energy Lett. 2022, 7, 1412–1445. [Google Scholar] [CrossRef]
  19. Qamar, M.Z.; Khalid, Z.; Shahid, R.; Tsoi, W.C.; Mishra, Y.K.; Kyaw, A.K.K.; Saeed, M.A. Advancement in indoor energy harvesting through flexible perovskite photovoltaics for self-powered IoT applications. Nano Energy 2024, 129, 109994. [Google Scholar] [CrossRef]
  20. Li, Y.; Meng, L.; Yang, Y.; Xu, G.; Hong, Z.; Chen, Q.; You, J.; Li, G.; Yang, Y.; Li, Y. High-efficiency robust perovskite solar cells on ultrathin flexible substrates. Nat. Commun. 2016, 7, 10214. [Google Scholar] [CrossRef]
  21. Jung, H.S.; Han, G.S.; Park, N.G.; Ko, M.J. Flexible perovskite solar cells. Joule 2019, 3, 1850–1880. [Google Scholar] [CrossRef]
  22. Xie, M.; Wang, J.; Kang, J.; Zhang, L.; Sun, X.; Han, K.; Luo, Q.; Lin, J.; Shi, L.; Ma, C.Q. Super-flexible perovskite solar cells with high power-per-weight on 17 μm thick PET substrate utilizing printed Ag nanowires bottom and top electrodes. Flex. Print. Electron. 2019, 4, 034002. [Google Scholar] [CrossRef]
  23. Kim, S.; Oh, H.; Jeong, I.; Kang, G.; Park, M. Influence of a solvent trap in ITO/PEN substrates on the performance of flexible perovskite solar cells and light-emitting diodes. ACS Appl. Electron. Mater. 2021, 3, 3207–3217. [Google Scholar] [CrossRef]
  24. Castriotta, L.A.; Pineda, R.F.; Babu, V.; Spinelli, P.; Taheri, B.; Matteocci, F.; Brunetti, F.; Wojciechowski, K.; Di Carlo, A. Light-Stable Methylammonium-Free Inverted Flexible Perovskite Solar Modules on PET Exceeding 10.5% on a 15.7 cm2 Active Area. ACS Appl. Mater. Interfaces 2021, 13, 29576–29584. [Google Scholar] [CrossRef] [PubMed]
  25. Dkhili, M.; Lucarelli, G.; De Rossi, F.; Taheri, B.; Hammedi, K.; Ezzaouia, H.; Brunetti, F.; Brown, T.M. Attributes of high-performance electron transport layers for perovskite solar cells on flexible PET versus on glass. ACS Appl. Energy Mater. 2022, 5, 4096–4107. [Google Scholar] [CrossRef] [PubMed]
  26. Holzhey, P.; Prettl, M.; Collavini, S.; Mortan, C.; Saliba, M. Understanding the impact of surface roughness: Changing from FTO to ITO to PEN/ITO for flexible perovskite solar cells. Sci. Rep. 2023, 13, 6375. [Google Scholar] [CrossRef] [PubMed]
  27. Lin, J.; He, Y.; Mo, H.; Khaleed, A.; Ren, Z.; Li, Y.; Cao, Y.; Tang, J.; Wang, W.T.; Feng, S.P.; et al. Substrate Modifications for Stability Improvements of Flexible Perovskite Solar Cells. Energy Technol. 2024, 12, 2300958. [Google Scholar] [CrossRef]
  28. Boehme, M.; Charton, C. Properties of ITO on PET film in dependence on the coating conditions and thermal processing. Surf. Coat. Technol. 2005, 200, 932–935. [Google Scholar] [CrossRef]
  29. Carneiro-da-Cunha, M.G.; Cerqueira, M.A.; Souza, B.W.S.; Carvalho, S.; Quintas, M.A.C.; Teixeira, J.A.; Vicente, A.A. Physical and thermal properties of a chitosan/alginate nanolayered PET film. Carbohydr. Polym. 2010, 82, 153–159. [Google Scholar] [CrossRef]
  30. Zardetto, V.; Brown, T.M.; Reale, A.; Di Carlo, A. Substrates for flexible electronics: A practical investigation on the electrical, film flexibility, optical, temperature, and solvent resistance properties. J. Polym. Sci. B Polym. Phys. 2011, 49, 638–648. [Google Scholar] [CrossRef]
  31. Lee, H.S.; Bang, J.O.; Lee, H.J.; Lee, G.J.; Chai, K.H.; Jung, S.B. Presented at the Electronic Components and Technology Conference, Lake Buena Vista, FL, USA, 31 May–3 June 2011; pp. 1796–1799.
  32. Faraj, M.; Ibrahim, K.; Ali, M.J. PET as a plastic substrate for the flexible optoelectronic applications. Optoelectron. Adv. Mater.-Rapid Commun. 2011, 5, 879–882. [Google Scholar]
  33. Wang, C.; Guan, L.; Zhao, D.; Yu, Y.; Grice, C.R.; Song, Z.; Awni, R.A.; Chen, J.; Wang, J.; Zhao, X.; et al. Water Vapor Treatment of Low-Temperature Deposited SnO2 Electron Selective Layers for Efficient Flexible Perovskite Solar Cells. ACS Energy Lett. 2017, 2, 2118–2124. [Google Scholar] [CrossRef]
  34. Qiu, X.; Yang, B.; Chen, H.; Liu, G.; Liu, Y.; Yuan, Y.; Huang, H.; Xie, H.; Niu, D.; Gao, Y.; et al. Efficient, stable and flexible perovskite solar cells using two-step solution-processed SnO2 layers as electron-transport-material. Org. Electron. 2018, 58, 126–132. [Google Scholar] [CrossRef]
  35. Xie, H.; Yin, X.; Chen, P.; Liu, J.; Yang, C.; Que, W.; Wang, G. Solvothermal synthesis of highly crystalline SnO2 nanoparticles for flexible perovskite solar cells application. Mater. Lett. 2019, 234, 311–314. [Google Scholar] [CrossRef]
  36. Liu, C.; Zhang, L.; Zhou, X.; Gao, J.; Chen, W.; Wang, X.; Xu, B. Hydrothermally Treated SnO2 as the Electron Transport Layer in High-Efficiency Flexible Perovskite Solar Cells with a Certificated Efficiency of 17.3%. Adv. Funct. Mater. 2019, 49, 1807604. [Google Scholar] [CrossRef]
  37. Chen, L.; Liu, Z.; Qiu, L.; Xiong, J.; Song, L.; Du, P. Multifunctional Regulation of SnO2 Nanocrystals by Snail Mucus for Preparation of Rigid or Flexible Perovskite Solar Cells in Air. ACS Nano 2023, 17, 23794–23804. [Google Scholar] [CrossRef] [PubMed]
  38. Mishra, S.; Ghosh, S.; Singh, T. Progress in materials development for flexible perovskite solar cells and future prospects. ChemSusChem 2021, 14, 512–538. [Google Scholar] [CrossRef]
  39. Liu, J.; Ye, T.; Yu, D.; Liu, S.; Yang, D. Piezocatalytic Techniques in Environmental Remediation. Angew. Chem. Int. Ed. 2023, 135, 202213927. [Google Scholar] [CrossRef]
  40. Wen, J.; Zhang, K.; Zhang, X.; Wang, Y.; Song, Y. Optimizing Hole Transport Materials for Perovskite Solar Cells: Spiro-OMeTAD Additives, Derivatives and Substitutes. Chem. Asian J. 2025, 25, e202500056. [Google Scholar] [CrossRef]
  41. Wan, Z.; Wei, R.; Jiang, S.; Wang, Y.; Yin, H.; Zeng, H.; Azam, M.; Luo, J.; Jia, C. Increasing the Li-TFSI doping concentration in Spiro-OMeTAD enables efficient and stable perovskite solar cells. J. Mater. Chem. C Mater. 2025, 13, 10690–10699. [Google Scholar] [CrossRef]
  42. Xie, G.; Chen, L.; Liu, J.; Yu, J.; Yin, H.; Li, H.; Yang, Y.; Liang, A.; Chen, Y. Recent Progress on Spiro-type Hole Transport Materials for Efficient and Stable Perovskite Solar Cells. J. Mater. Chem. C Mater. 2025, 13, 2040–2057. [Google Scholar] [CrossRef]
  43. Kanda, H.; Shibayama, N.; Huckaba, A.J.; Lee, Y.; Paek, S.; Klipfel, N.; Roldán-Carmona, C.; Queloz, V.I.E.; Grancini, G.; Zhang, Y.; et al. Band-bending induced passivation: High performance and stable perovskite solar cells using a perhydropoly (silazane) precursor. Energy Environ. Sci. 2020, 13, 1222–1230. [Google Scholar] [CrossRef]
  44. Zhang, H.; Pfeifer, L.; Zakeeruddin, S.M.; Chu, J.; Grätzel, M. Tailoring passivators for highly efficient and stable perovskite solar cells. Nat. Rev. Chem. 2023, 7, 632–652. [Google Scholar] [CrossRef]
  45. Wang, J.; Bi, L.; Fu, Q.; Jen, A.K.Y. Methods for passivating defects of perovskite for inverted perovskite solar cells and modules. Adv. Energy Mater. 2024, 14, 2401414. [Google Scholar] [CrossRef]
  46. Gao, C.; Jia, S.; Yin, X.; Li, Z.; Yang, G.; Chen, J.; Li, Z.; An, X. Enhancing open-circuit voltage in FAPbI 3 perovskite solar cells via self-formation of coherent buried interface FAPbIx Cl3−x. Chem. Commun. 2025, 61, 2758–2761. [Google Scholar] [CrossRef] [PubMed]
  47. Li, N.; Hu, X.; Tang, Y.; Lei, Y.; Lau, Y.S.; Chen, Q.; Sui, X.; Zhu, F. Perovskite/organic tandem device to realize light detection and emission dual function. Chem. Eng. J. 2024, 490, 151573. [Google Scholar] [CrossRef]
  48. Zhu, B.P.; Guo, W.K.; Zhu, B.P.; Zhou, Q.; Shung, K.K.; Zhu, B.P.; Shen, G.Z. Structure and electrical properties of (111)-oriented Pb(Mg1/3Nb2/3)O3-PbZrO3-PbTiO3 thin film for ultra-high-frequency transducer applications. IEEE Trans. Ultrason. Ferroelectr. Freq. Control 2011, 58, 1962–1967. [Google Scholar] [CrossRef]
  49. Zhu, B.P.; Wang, Z.Y.; Zhang, Y.; Yu, Z.S.; Shi, J.; Xiong, R. Low temperature fabrication of the giant dielectric material CaCu3Ti4O12 by oxalate coprecipitation method. Mater. Chem. Phys. 2009, 113, 746–748. [Google Scholar] [CrossRef]
  50. Jiang, Q.; Zhao, Y.; Zhang, X.; Yang, X.; Chen, Y.; Chu, Z.; Ye, Q.; Li, X.; Yin, Z.; You, J. Surface passivation of perovskite film for efficient solar cells. Nat. Photonics 2019, 13, 460–466. [Google Scholar] [CrossRef]
  51. Zhang, H.; Wang, S.; Hou, Y.; Zhang, F.; Hao, Y.; Song, J.; Qu, J. Comparison of surface-passivation ability of the BAI salt and its induced 2D perovskite for high-performance inverted perovskite solar cells. RSC Adv. 2021, 11, 23249–23258. [Google Scholar] [CrossRef]
  52. Xia, J.; Liang, C.; Gu, H.; Mei, S.; Li, S.; Zhang, N.; Chen, S.; Cai, Y.; Xing, G. Surface passivation toward efficient and stable perovskite solar cells. Energy Environ. Mater. 2023, 6, e12296. [Google Scholar] [CrossRef]
  53. Wang, F.; Geng, W.; Zhou, Y.; Fang, H.-H.; Tong, C.-J.; Loi, M.A.; Liu, L.-M.; Zhao, N. Phenylalkylamine passivation of organolead halide perovskites enabling high-efficiency and air-stable photovoltaic cells. Adv. Mater. 2016, 28, 9986–9992. [Google Scholar] [CrossRef]
  54. Yoo, H.-S.; Park, N.-G. Post-treatment of perovskite film with phenylalkylammonium iodide for hysteresis-less perovskite solar cells. Sol. Energy Mater. Sol. Cells 2018, 179, 57–65. [Google Scholar] [CrossRef]
  55. Kim, G.M.; Sato, H.; Ohkura, Y.; Ishii, A.; Miyasaka, T.; Kim, G.M.; Sato, H.; Ohkura, Y.; Ishii, A.; Miyasaka, T. Phenethylamine-Based Interfacial Dipole Engineering for High Voc Triple-Cation Perovskite Solar Cells. Adv. Energy Mater. 2022, 12, 2102856. [Google Scholar] [CrossRef]
  56. Huang, X.; Bi, W.; Jia, P.; Tang, Y.; Lou, Z.; Hu, Y.; Cui, Q.; Hou, Y.; Teng, F. Enhanced efficiency and light stability of planar perovskite solar cells by diethylammonium bromide induced large-grain 2D/3D hybrid film. Org. Electron. 2019, 67, 101–108. [Google Scholar] [CrossRef]
  57. Liu, S.; Sun, Z.; Lei, X.; Miao, T.; Zhou, Q.; Chen, R.; Wang, J.; Ren, F.; Pan, Y.; Cai, Y.; et al. Stable Surface Contact with Tailored Alkylamine Pyridine Derivatives for High-Performance Inverted Perovskite Solar Cells. Adv. Mater. 2024, 37, 2415100. [Google Scholar] [CrossRef] [PubMed]
  58. AlSabeh, G.; Almalki, M.; Kasemthaveechok, S.; Ruiz-Preciado, M.A.; Zhang, H.; Vanthuyne, N.; Zimmermann, P.; Dekker, D.M.; Eickemeyer, F.T.; Hinderhofer, A.; et al. Helical interfacial modulation for perovskite photovoltaics. Nanoscale Adv. 2024, 6, 3029–3033. [Google Scholar] [CrossRef]
  59. Li, L.; Lai, Y.; Rao, H.; Cai, W.; Zhong, X.; Guo, H.; Pan, Z. 2D/3D Heterojunction Engineering for Hole Transport Layer-Free Carbon-Based Perovskite Solar Cells. ChemSusChem 2025, 18, e202402549. [Google Scholar] [CrossRef]
  60. Xu, Z.; Guo, Z.; Li, H.; Zhou, Y.; Liu, Z.; Wang, K.; Li, Z.; Wang, H.; Qaid, S.M.H.; Mohammed, O.F.; et al. Efficient and stable inverted MA/Br-free 2D/3D perovskite solar cells enabled by α-to-δ phase transition inhibition and crystallization modulation. Energy Environ. Sci. 2025, 18, 1354–1365. [Google Scholar] [CrossRef]
  61. Shi, Y.R.; Chen, C.H.; Lou, Y.H.; Wang, Z.K. Strategies of perovskite mechanical stability for flexible photovoltaics. Mater. Chem. Front. 2021, 5, 7467–7478. [Google Scholar] [CrossRef]
  62. Ma, Y.; Lu, Z.; Su, X.; Zou, G.; Zhao, Q. Recent progress toward commercialization of flexible perovskite solar cells: From materials and structures to mechanical stabilities. Adv. Energy Sustain. Res. 2023, 4, 2200133. [Google Scholar] [CrossRef]
  63. Du, D.; Qiao, F.; Guo, Y.; Wang, F.; Wang, L.; Gao, C.; Zhang, D.; Liang, J.; Xu, Z.; Shen, W.; et al. Photovoltaic performance of flexible perovskite solar cells under bending state. Sol. Energy 2022, 245, 146–152. [Google Scholar] [CrossRef]
  64. Xie, H.; Liang, T.; Yin, X.; Liu, J.; Liu, D.; Wang, G.; Gao, B.; Que, W. Mechanical stability study on PEDOT: PSS-based ITO-free flexible perovskite solar cells. ACS Appl. Energy Mater. 2022, 5, 3081–3091. [Google Scholar] [CrossRef]
  65. Zhao, C.; Zhou, Z.; Almalki, M.; Hope, M.A.; Zhao, J.; Gallet, T.; Krishna, A.; Mishra, A.; Eickemeyer, F.T.; Xu, J.; et al. Stabilization of highly efficient perovskite solar cells with a tailored supramolecular interface. Nat. Commun. 2024, 15, 7139. [Google Scholar] [CrossRef] [PubMed]
  66. Kim, M.; Kim, G.H.; Lee, T.K.; Choi, I.W.; Choi, H.W.; Jo, Y.; Yoon, Y.J.; Kim, J.W.; Lee, J.; Huh, D.; et al. Methylammonium chloride induces intermediate phase stabilization for efficient perovskite solar cells. Joule 2019, 3, 2179–2192. [Google Scholar] [CrossRef]
  67. Ligavo, M.M.; Waita, S.; Nyongesa, F.W.; Diale, M. Surface passivation of perovskite thin films through sequential physical vapour deposition for improved stability, structural, optical and morphological properties. Opt. Mater. 2025, 159, 116613. [Google Scholar] [CrossRef]
  68. Li, M.; Huang, P.; Zhong, H.J. Current understanding of band-edge properties of halide perovskites: Urbach tail, rashba splitting, and exciton binding energy. Phys. Chem. Lett. 2023, 14, 1592–1603. [Google Scholar] [CrossRef]
  69. Almalki, M.; Alotaibi, M.H.; Alanazi, A.Q.; Eickemeyer, F.T.; Alenzi, S.M.; Alzahrani, Y.A.; Piveteau, L.; Alymani, A.Y.; Albadri, A.; Albrithen, H.; et al. Interfacial Modulation through Mixed-Dimensional Heterostructures for Efficient and Hole Conductor-Free Perovskite Solar Cells. Adv. Funct. Mater. 2024, 34, 2309789. [Google Scholar] [CrossRef]
  70. Wang, H.; Zhang, Z.; Wang, X.; Duan, L.; Luo, J. Phenyltrimethylammonium chloride additive for highly efficient and stable FAPbI3 perovskite solar cells. Nano Energy 2024, 123, 109423. [Google Scholar] [CrossRef]
  71. Pradhan, N. Alkylammonium halides for facet reconstruction and shape modulation in lead halide perovskite nanocrystals. Acc. Chem. Res. 2021, 54, 1200–1208. [Google Scholar] [CrossRef]
  72. Li, H.; Wang, W.; Yang, Y.; Wang, Y.; Li, P.; Huang, J.; Li, J.; Lu, Y.; Li, Z.; Wang, Z.; et al. Kirigami-based highly stretchable thin film solar cells that are mechanically stable for more than 1000 cycles. ACS Nano 2020, 14, 1560–1568. [Google Scholar] [CrossRef]
  73. Wang, S.; Yang, B.; Han, J.; He, Z.; Li, T.; Cao, Q.; Yang, J.; Suo, J.; Li, X.; Liu, Z.; et al. Polymeric room-temperature molten salt as a multifunctional additive toward highly efficient and stable inverted planar perovskite solar cells. Energy Environ. Sci. 2020, 13, 5068–5079. [Google Scholar] [CrossRef]
  74. Liu, C.; Sun, J.; Tan, W.L.; Lu, J.; Gengenbach, T.R.; McNeill, C.R.; Ge, Z.; Cheng, Y.B.; Bach, U. Alkali cation doping for improving the structural stability of 2D perovskite in 3D/2D PSCs. Nano Lett. 2020, 20, 1240–1251. [Google Scholar] [CrossRef]
  75. Yu, S.; Wang, Y.; Chen, T.; Li, M.; Zhang, X.; Huang, B.; Xu, J.; Wang, G. An inclined groove and its optimization design method for improving the energy performance at the saddle zone of axial flow pumps. Energy 2025, 328, 136527. [Google Scholar] [CrossRef]
  76. Li, J.; Zhu, X.; Feng, C.; Wen, M.; Zhang, Y. A simple and efficient three-dimensional spring element model for pore seepage problems. Eng. Anal. Bound Elem. 2025, 176, 106225. [Google Scholar] [CrossRef]
  77. Wang, Y.; Xia, S.; Chen, K.; Zhang, J.; Yu, C.; Wu, J.; Wang, P.; Zhang, W.; Wu, Y. Balancing Intermediates Formation on Atomically Pd-Bridged Cu/Cu2O Interfaces for Kinetics-Matching Electrocatalytic C–N Coupling Reaction. Angew. Chem. Int. Ed. 2025, 64, e202503011. [Google Scholar] [CrossRef] [PubMed]
  78. Lv, H.; Zeng, J.; Zhu, Z.; Dong, S.; Li, W. Study on prestress distribution and structural performance of heptagonal six-five-strut alternated cable dome with inner hole. Structures 2024, 65, 106724. [Google Scholar] [CrossRef]
  79. Ma, C.; Mu, R.; Li, M.; He, J.; Hua, C.; Wang, L.; Liu, J.; Totis, G.; Yang, J.; Liu, K.; et al. A multi-scale spatial–temporal interaction fusion network for digital twin-based thermal error compensation in precision machine tools. Expert Syst. Appl. 2025, 286, 127812. [Google Scholar] [CrossRef]
Figure 1. (a) Molecular structure of polyethylene terephthalate (PET) and phenethylammonium iodide (PEAI). (b) A schematic representation of the FPSC. (c) XRD diffraction pattern of reference and PEAI-treated perovskite film. Top-view SEM images of (d) reference and (e) PEAI-treated perovskite films. (f) Cross-sectional SEM image of PEAI-treaed perovskite film.
Figure 1. (a) Molecular structure of polyethylene terephthalate (PET) and phenethylammonium iodide (PEAI). (b) A schematic representation of the FPSC. (c) XRD diffraction pattern of reference and PEAI-treated perovskite film. Top-view SEM images of (d) reference and (e) PEAI-treated perovskite films. (f) Cross-sectional SEM image of PEAI-treaed perovskite film.
Nanomaterials 15 01078 g001
Figure 2. (a) UV-Vis and (b) PL spectra for the reference and PEAI-passivated perovskite films.
Figure 2. (a) UV-Vis and (b) PL spectra for the reference and PEAI-passivated perovskite films.
Nanomaterials 15 01078 g002
Figure 3. (a) Statistical representation of photovoltaic parameters, (b) J-V curve of the champion devices. (c) Stress bending test of the reference and PEAI-passivated FPSCs. (d) A photograph of a bent PEA-passivated FPSC.
Figure 3. (a) Statistical representation of photovoltaic parameters, (b) J-V curve of the champion devices. (c) Stress bending test of the reference and PEAI-passivated FPSCs. (d) A photograph of a bent PEA-passivated FPSC.
Nanomaterials 15 01078 g003
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

Almalki, I.S.; Alenazi, T.H.; Mansouri, L.A.; Al Mubarak, Z.H.; Al Nahab, Z.T.; Alenzi, S.M.; Alzahrani, Y.A.; Yafi, G.S.; Almutairi, A.; Aldukhail, A.; et al. Enhanced Efficiency and Mechanical Stability in Flexible Perovskite Solar Cells via Phenethylammonium Iodide Surface Passivation. Nanomaterials 2025, 15, 1078. https://doi.org/10.3390/nano15141078

AMA Style

Almalki IS, Alenazi TH, Mansouri LA, Al Mubarak ZH, Al Nahab ZT, Alenzi SM, Alzahrani YA, Yafi GS, Almutairi A, Aldukhail A, et al. Enhanced Efficiency and Mechanical Stability in Flexible Perovskite Solar Cells via Phenethylammonium Iodide Surface Passivation. Nanomaterials. 2025; 15(14):1078. https://doi.org/10.3390/nano15141078

Chicago/Turabian Style

Almalki, Ibtisam S., Tamader H. Alenazi, Lina A. Mansouri, Zainab H. Al Mubarak, Zainab T. Al Nahab, Sultan M. Alenzi, Yahya A. Alzahrani, Ghazal S. Yafi, Abdulmajeed Almutairi, Abdurhman Aldukhail, and et al. 2025. "Enhanced Efficiency and Mechanical Stability in Flexible Perovskite Solar Cells via Phenethylammonium Iodide Surface Passivation" Nanomaterials 15, no. 14: 1078. https://doi.org/10.3390/nano15141078

APA Style

Almalki, I. S., Alenazi, T. H., Mansouri, L. A., Al Mubarak, Z. H., Al Nahab, Z. T., Alenzi, S. M., Alzahrani, Y. A., Yafi, G. S., Almutairi, A., Aldukhail, A., Alharthi, B., Aljuwayr, A., Alghannam, F. S., Almuqhim, A. A., Alkhaldi, H., Alhajri, F., AL-Saleem, N. K., Alkahtani, M., Alanazi, A. Q., & Almalki, M. (2025). Enhanced Efficiency and Mechanical Stability in Flexible Perovskite Solar Cells via Phenethylammonium Iodide Surface Passivation. Nanomaterials, 15(14), 1078. https://doi.org/10.3390/nano15141078

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