Abdelaal S. A. Ahmed*a,
Fatma S. M. Hashemb,
Abu-Bakr A. El-Adasya and
Tarek A. Seaf El-Nasrac
aChemistry Department, Faculty of Science, Al-Azhar University, Assiut 71524, Egypt. E-mail: abdelaalsaiyd@azhar.edu.eg; abdelaalsaiyd@gmail.com
bChemistry Department, Faculty of Science, Assuit University, Assiut 71524, Egypt
cDepartment of Chemistry, College of Science and Arts, Jouf University, Sakaka 2014, Saudi Arabia
First published on 7th July 2025
Flexible perovskite solar cells (FPSCs) have received extensive interest for application in wearable electronic devices owing to their high flexibility, light weight, and compatibility with irregular-shaped electronic devices. The roll-to-roll manufacturing method, which exhibits high capacity for mass production, has enabled significant advancements in FPSCs via composition engineering, interface modification, fabrication process optimization, and new charge transport materials. Devices utilizing high-quality perovskite films have achieved a power conversion efficiency (PCE) exceeding 24%. This paper comprehensively outlines the recent advancements in the development of FPSCs, including flexible substrates, electrodes, low-temperature interlayers, and diverse methodologies for fabricating high-quality perovskite films. This review also discusses the existing challenges and outlines future opportunities for growth in this rapidly evolving field.
Till now, various kinds of FSCs have been developed, including flexible silicon-based solar cells and TFSCs (e.g., CIGS and CdTe).32 Unfortunately, their general applicability is constrained by the complex fabrication methods linked to the high cost and use of various dangerous metals.33 Thus, great attention has been devoted to developing alternative FSCs from the third generation, such as flexible-DSSCs (FDSSCs),34 flexible-QDSSCs (FQDSSCs), and flexible-PSCs (FPSCs).35 Compared with other flexible photovoltaic systems, FPSCs have emerged as viable options due to their excellent photovoltaic efficiency and lightweight, flexible design.19,36 It is anticipated that FPSCs will find specialized uses in wearable and portable electronics, self-powered electronics,37 and smart integrated buildings.38 More significantly, FPSCs can be made using the R-to-R printing technique,39 which will facilitate their mass manufacturing and advance their future commercialization.40,41 Consequently, the development of FPSCs is crucial for the real-world use of PSCs. Thus, in this work, we aimed to present the current developments in FPSCs, including flexible substrates, electrode preparation, and the production of perovskite films. Then, the real application of FPSCs with common examples is introduced. Finally, a conclusion and outlook for the further development of FPSCs are provided.
Based on the density of states (DOS) and partial charge density, the valence band maximum (VBM) in halide perovskites exhibits strong antibonding character between the Pb s and I p orbitals. In contrast, the conduction band minimum (CBM) is primarily formed by the Pb p states. This unique electronic structure, featuring both ionic and covalent characteristics, stems from the molecular structure of the perovskite as well as the properties of the individual ions.49 In conventional semiconductors, the VBM and CBM are typically dominated by p and s orbitals, respectively.49 However, in halide perovskites, this trend is reversed. This inversion is one of the factors contributing to their exceptional optoelectronic performance. Another key advantage of perovskite solar cells (PSCs) is their high optical absorption, in which a perovskite absorber layer with a thickness of less than 500 nm can achieve a PCE of 15%. In comparison, the absorber layers in first and second-generation photovoltaics usually require much thicker layers; about 300 μm and 2 μm, respectively. The optical absorption of semiconductors is typically evaluated using two parameters, as follows: (i) the transition matrix elements between the VB and CB states, which determine the probability of each photoelectric transition, and (ii) the joint density of states (JDOS), which represents the total number of possible photoelectric transitions. Therefore, the optical absorption coefficient is closely linked to the electronic structure of the materials. Fig. 1 illustrates the mechanisms of optical absorption in first, second, and third-generation photovoltaics.
![]() | ||
Fig. 1 Scheme representing the optical absorption of the absorbers in the 1st, 2nd, and 3rd generations of PVs. Reprinted with permission from ref. 49. Copyright 2015, Royal Society of Chemistry. |
In silicon absorbers, optical absorption primarily occurs through transitions involving Si p and Si s/p orbitals. However, because silicon is an indirect bandgap semiconductor, the probability of electronic transitions between their VB and CB is significantly lower; approximately two times lower compared to materials with a direct bandgap. Consequently, the absorber layer must be two orders of magnitude thicker to achieve sufficient absorption, which substantially increases the overall material cost. Other absorber materials, such as GaAs and halide perovskites (e.g., CH3NH3PbI3), possess direct bandgaps, which result in significantly stronger optical absorption compared to silicon. However, their electronic structures differ in notable ways. In CH3NH3PbI3, the lower portion of its CB is primarily composed of degenerate Pb p orbitals, while in GaAs, the CB edge arises from a more dispersive s band. Additionally, in CH3NH3PbI3, its edge transition originates from the hybridized (Pb s, I p) orbitals to Pb p orbitals. The transition probability between the VBM and CBM in CH3NH3PbI3 is comparable to that in GaAs, owing to the strong intra-atomic Pb s → Pb p transition. Taken together, these features enable halide perovskites to exhibit even greater optical absorption than GaAs.
Based on the above-mentioned characteristics, perovskite materials demonstrate strong potential for photovoltaic applications, both in conventional and flexible formats. Among them, organic–inorganic lead halide perovskites, CH3NH3PbI3, are some of the most widely used absorber materials for photovoltaics due to their optimal bandgap (1.57 eV),53 and low exciton binding energy (50 meV).54 The use of perovskite materials in opto electronic devices dates back to 1994, as reported by Tsutsui et al.55 In this study, they employed the (C6H5C2H4NH3)2PbI4 perovskite material in devices to study their fluorescence under electrical excitation. At low temperatures, an emission peak at 520 nm with a narrow full width at half-maximum (FWHM) of about 10 nm was observed. By utilising light-emitting diodes (LEDs), the obtained perovskite-based LEDs achieved brightness levels nearly five-times higher than conventional organic light-emitting diodes (OLED). Since then, the remarkable optical and electronic properties of perovskite materials have attracted significant attention for photovoltaic applications.56 The photovoltaic performance of perovskite materials was first reported in 2009 by Kojima et al.,57 in which a hybrid perovskite was used as a highly effective light-harvesting material in liquid electrolyte-based DSSCs. The assembled device achieved a PCE of 3.9%. However, a major limitation was its poor stability, given that the perovskite rapidly degraded in the presence of the liquid electrolyte. Thus, to address this issue, subsequent research introduced solid-state organic hole transport materials as alternatives to the liquid electrolyte.58–60 In less than a decade, PSCs have achieved remarkable progress, with their PCE exceeding 25%. This value exceeds the PCEs of polycrystalline Si-PVs (22.3%), thin film crystalline Si PVs (21.2%), CuInSe2 (22.6%), and CdTe PVs (22.1%). These improvements in the overall performance of PSCs can assigned to their many intrinsic properties such as high absorption coefficients, high defect tolerance, desirable/tunable bandgap, mechanical flexibility, and long charge carrier diffusion lengths.61,62 Additionally, the cost-effective materials, mechanical durability, and low-temperature fabrication procedures (typically <150 °C) of PSCs make them suitable for realizing flexible perovskite solar cells (FPSCs) using a plastic substrate. Furthermore, FPSCs using plastic substrates would produce the most competitive power per weight among the solar cells. Consequently, FPSCs can be employed in specialized fields including electronic textiles, large-scale industrial roofs, portable electric chargers, and unmanned aerial vehicle (UAV) power sources.35,63–68 FPSCs have shown a rapid performance improvement, and the first study on FPSCs was reported in 2013 by Mathews et al.,69 in which the assembled FPSCs displayed a PCE of 2.62%. Subsequent developments have led to continuous performance enhancements, with recent devices achieving PCEs as high as 24.3% (Fig. 2),35,69–80 which is considered the highest performance among TFSC devices. In comparison to alternative photovoltaic technologies such as Si, Cu(In, Ga)Se2 (CIGS), CdTe, and organic solar cells, FPSCs have demonstrated a remarkable specific power-per-weight (W g−1).36,81,82 For example, Kaltenbrunner et al.83 developed a lightweight FPSC that achieved a PCE exceeding 12% by utilizing an ultra-thin PET substrate with a thickness of 1.4 μm. These flexible devices maintained their original performance under 40% compression. Furthermore, owing to their extremely low thickness, the assembled devices showed a specific power output reaching 23 W g−1. Furthermore, Kang et al.84 reported that FPSCs assembled with silver nanowire (AgNW) transparent electrodes fabricated on 1.3 μm-thick polyethylene naphthalate foils demonstrated a specific power output of 29.4 W g−1, outperforming other thin-film photovoltaics such as amorphous silicon (a-Si; 8.31 W g−1),85 organic solar cells (OSC; 10 W g−1),86 lead sulfide quantum dot solar cells (PbS QDs; 12.3 W g−1),87 and CdS/CdTe solar cells (0.254 W g−1),88 which is considered one of the highest values in photovoltaic technology.
![]() | ||
Fig. 2 Efficiency evolution of FPSCs from 2013 to 2025. Reprinted with permission from ref. 35, 69–72 and 74–80. Copyright 2013, 2014, 2017, 2020, 2021, Royal Society of Chemistry. Copyright 2015, 2016, 2022, 2023, 2025, Springer Nature. Copyright 2018, 2019 Wiley-VCH, Copyright 2024, American Chemical Society, [2013 PCE = 2.62%, 2014 PCE = 7.0%, 2015 PCE = 15.3%, 2016 PCE = 14.0%, 2017 PCE = 16.80%, 2018 PCE = 18.40%, 2019 PCE = 19.51%, 2020 PCE = 19.90%, 2021 PCE = 21.10%, 2022 PCE = 24.40%, 2023 PCE = 23.86%, 2024 PCE = 23.81%, 2025 PCE = 24.43%] . |
Mechanical flexibility is another critical factor for the real applicability of FPSC devices. The fabrication of deformable and stretchable FPSCs is possible through the use of flexible substrates with excellent bending resilience.35,89,90 For example, Park et al. utilized a shape-memory polymer as a substrate to develop highly deformable FPSCs.90 The assembled FPSC devices displayed excellent mechanical stability, as indicated by the loss of only 40% of their initial value after 50 cycles of complete deformation and recovery. Additionally, one of the primary characteristics of FPSCs is their ability to work with flexible substrates, which permits high-throughput R-to-R production and may attract significant industry participants. In addition to enabling the manufacture of conventional silicon-based devices, R-to-R processing also makes it possible to create modules with unique shapes. Moreover, a significant advantage of FPSCs is their ability to work with flexible substrates, making them well-suited for high-speed R-to-R manufacturing, which can lead to their industrial production. The R-to-R technique not only accelerates production relative to conventional silicon devices but also facilitates the development of modules in various smart shapes.91–94 For example, Galagan et al.95 used the R-to-R slot die coating method to prepare FPSCs on flexible substrates. The fabricated FPSC devices with a mask of 0.04 cm2 displayed a PCE of 14.5%. Additionally, the overall performance of the manufactured devices remained unchanged after 1000 bending cycles with a bending radius of 10 mm. In another study, Bu et al.96 used the slot-die print method for preparing high-quality tin oxide (SnO2) films for FPSCs. The small-size FPSCs achieved a PCE of 17.18%, while the devices with larger sizes (5 × 6 cm2) displayed an efficiency of over 15%.
All the above-mentioned amazing characteristics make FPSCs ideal for wearable and portable electronic applications. However, the relatively lower PCE of FPSCs than that for rigid devices is still a challenge that seriously limits their further commercialization, and thus needs to be addressed.97 The roughness and deformability of polymer substrates are common factors responsible for damage to the quality of the charge transport and perovskite layers, lowering the efficiency of FPSCs.98 In addition to mechanical issues, the high sheet resistance and low transparency of flexible substrates negatively impact the FF and JSC of FPSCs, particularly in the context of large-area device fabrication.99 Furthermore, the low processing temperature (∼100 °C) necessary for flexible substrates limits the crystallinity of the charge transport layers and perovskite films, thereby reducing the device efficiency.74 Therefore, in the last few years, great efforts have been devoted to overcoming these limitations, and hence it is important to review and summarize the recent progress in the development of FPSCs.
![]() | ||
Fig. 3 Configuration of the PSC device; (a) n-i-p mesoscopic, (b) n-i-p planar, and (c) p-i-n planar PSC structures. Reprinted with permission from ref. 56. Copyright 2018, Royal Society of Chemistry. |
![]() | ||
Fig. 4 (a–g) SEM of perovskite films deposited onto m-TiO2 films, (h) cross sectional SEM image at 150 °C, and (i) J–V curves of the assembled FPSC devices. Reprinted with permission from ref. 107 Copyright 2014, Wiley-VCH. |
Besides controlling the morphology, the choice of solvent is also important, given that it affects the solubility of the precursors and the formation of intermediate phases.108,109 The commonly used organic solvents include dimethylformamide (DMF), dimethyl sulfoxide (DMSO), methyl-2-pyrrolidinone (NMP), γ-butyrolactone (GBL), and toluene, which are utilized individually or as mixtures.110–113 For example, Seok et al.114 utilized a mixed solvent of GBL/DMSO and toluene drop-casting method to prepare homogenous and dense perovskite layers of CH3NH3I–PbI2–DMSO intermediate phase. The assembled device achieved a PCE of 16.2%.
The two-step method is an alternative strategy that was very well-liked in the early stage of perovskite research. It involves depositing two precursors flowed by thermal annealing.59 In this approach, a PbI2 precursor is first deposited onto the conductive side of the substrate, followed by exposure to a CH3NH3I solution.115,116 Fig. 5 illustrates the differences between the low-temperature one-step and two-step methods.117 Generally, the two-step method yields an improved perovskite morphology and higher performance than that made using the one-step process.59,118 This technique was first introduced by Mitzi et al.,119 where a PbI2 film was immersed in a CH3NH3I solution.
![]() | ||
Fig. 5 Fabrication of CH3NH3PbI3 films via one-step and two-step spin-coating techniques. Reprinted with permission from ref. 117 Copyright 2014, AIP Publishing. |
Due to its simplicity and reproducibility, the two-step deposition method has been widely adopted for the fabrication of PSC devices.120 For instance, Huang et al.121 successfully prepared pin-hole-free CH3NH3PbI3 perovskite layers using a solution process by sequentially spin-coating PbI2 and CH3NH3I layers (Fig. 6a). The PbI2 and CH3NH3I precursors were dissolved in DMF and 2-propanol, respectively. Then, the resulting perovskite films were heated at 100 °C for varying durations. As shown in the SEM images in Fig. 6b–d, the perovskite film exhibited enhanced continuity compared to that prepared from premixed precursors. The assembled FPSC device showed a PCE of 15.4%, outperforming the devices fabricated by the interdiffusion method (14.5%). This demonstrates that the two-step deposition method has potential for the fabrication of efficient and low-cost perovskite layers at low temperatures. However, the incomplete conversion of PbI2 and the uncontrolled configuration of the perovskite crystals are the main challenges.122 Thus, to address these issues, great efforts have focused on improving the morphology and enhancing the conversion of PbI2. The use of additives such as polymers, inorganic acids, and solvents has been proven to be an effective strategy for improving the quality of perovskite films.123–125 Various organic polymers have been successfully utilized as additives.126 For instance, Su et al.127 reported that introducing 1wt% poly(ethylene glycol) (PEG) into the perovskite precursors enhanced the PCE of the assembled device by 25% (Fig. 6i). This improvement was assigned to the better control of the crystal size and aggregation (Fig. 6e and f). Moreover, PEG exhibits strong interactions with perovskite molecules, which enhances the moisture resistance, and consequently improves the long-term stability of the devices.
![]() | ||
Fig. 6 (a) Scheme of the preparation of the perovskite layer via the interdiffusion method. SEM images of (b) the spun coated stacked PbI2 layer, (c) annealed CH3NH3PbI3 perovskite layers prepared by the interdiffusion method, and (d) annealed CH3NH3PbI3 perovskite layers prepared using spun premixed precursors. Reprinted with permission from ref. 121 Copyright 2014, Royal Society of Chemistry. SEM images of perovskite films with (e) 0 wt% PEG, (f) 1 wt% PEG, (g) 3 wt% PEG, and (h) 5 wt% PEG. (i) J–V of the devices with pristine and x wt% of PEG additive. Reprinted with permission from ref. 127 Copyright 2015, American Chemical Society. |
Inorganic acids such as hydroiodic acid (HI), hydrochloric acid (HCl), and hydrobromic (HBr) have also been employed as additives to significantly enhance the overall performance of PSCs.123 The beneficial roles of these acids can be summarized as follows:123 (i) they increase the solubility of the precursors, (ii) they prevent decomposition of the perovskite by converting elemental iodine into iodide ions, and (iii) they facilitate the formation of a pre-crystallized intermediate through interaction with the PbI2 precursor, which is crucial for improving the grain size and crystallinity of the resulting perovskite film. Several studies have demonstrated the effectiveness of these additives. For example, Wen et al. used HI acid as an additive in an isopropanol (IPA) solution of CH3NH3I during the two-step spin-coated method to fabricate high-quality CH3NH3PbI3 perovskite films. Increasing the HI concentration led to larger perovskite grain sizes. As presented in the SEM image (Fig. 7a–d), the film prepared with 0.004 vol% HI/IPA exhibited the most uniform and largest grain size. The assembled PSC device achieved a PCE of 18.21% (Fig. 7e). Similarly, Lu et al. reported that the addition of HI successfully enhanced the sensitivity of the perovskite particle sizes to the precursor (CH3NH3I/PbI2) molar ratio.128 Using a one-step spin-coating process, they fabricated a CH3NH3PbI3 film with full surface coverage and high crystallinity. The assembled planar device displayed a PCE of 19.29%, highlighting the effectiveness of HI in optimizing the film quality and device performance. Recently, Wu et al.129 demonstrated that water can serve as a promising additive in the perovskite precursors. In their study, a small amount of water was introduced in the PbI2/DMF solution to investigate its effects on the properties of PbI2 in the perovskite films fabricated using the two-step method. The resulting perovskite films were exceptionally pure, smooth, and dense, with no visible pinholes. This improvement facilitated the formation of a high-quality, pinhole-free perovskite layer with larger grain sizes and fewer defects, leading to an enhanced PCE of 18% in the inverted PSC device. These findings underscore the significant role of water in improving the perovskite film quality, and consequently the overall performance of inverted PSCs. Similarly, Meng et al. employed the two-step spin-coating method to fabricate FAxMA1−xPbI2.55Br0.45 perovskite layers by adding DMF to the FAI/MAI/IPA solution.130 According to the SEM images (Fig. 7f–i), the addition of 2% DMF enhanced the conversion of the precursors to perovskite, leading to an improved film morphology, reduced crystal defects, and better charge-transfer efficiency. The assembled PSC device achieved a PCE of 20.1% (Fig. 7j), demonstrating both high efficiency and low processing temperature, which are ideal for the fabrication of FPSCs. Furthermore, Yang et al.76 revealed that the use of dimethyl sulfide (DS) significantly slowed the perovskite crystallization process, allowing more controlled film formation. The assembled FPSC device is presented in Fig. 7k, and its cross section image is in Fig.7l. The resulting perovskite films exhibited larger crystal grains and enhanced stability (Fig. 7m and n), contributing to a maximum PCE of 18.40% in the corresponding FPSC devices (Fig. 7o). These results highlight the effectiveness of solvent engineering in optimizing the perovskite film quality and device performance.
![]() | ||
Fig. 7 (a) SEM images of MAPbI3 film prepared by two-step spin-coating deposition with (a) 0, (b) 0.002, (c) 0.004, and (d) 0.006 vol% HI/IPA. (e) J–V curves of the FPSCs. Reprinted with permission from ref. 131. Copyright 2018, AIP Advances. (c) SEM image of perovskite films obtained by different ratios of DMF in IPA, (f) w/o DMF, (g) 1% DMF, (h) 2% DMF, (i) 4%DMF. (j) J–V curves of the FPSCs. Reprinted with permission from ref. 130. Copyright 2017, American Chemical Society. (k) Scheme of the FPSC device, (i) cross-sectional SEM image of the device. SEM image of MAPbI3 films (m) without and (n) with DS additive, and (o) J–V curves of the FPSCs. Reprinted with permission from ref. 76. Copyright 2018, Wiley-VCH. |
![]() | ||
Fig. 8 (a) Scheme for the formation of the perovskite film through the vapour-assisted solution method, (b) SEM image of the perovskite film at 150 °C for 4 h, and (c) J–V curve of PSCs based on the as-prepared perovskite films (inset is the cross-sectional SEM of the FPSC). Reprinted with permission from ref. 132. Copyright 2014, American Chemical Society. |
In a separate study, the same research group employed magnetron sputtering to deposit amorphous TiO2 (am-TiO2) as the ETL for FPSCs.133 Subsequently, PbI2 was deposited onto the substrate using a vacuum evaporation process, while an aluminium plate was coated with a layer of CH3NH3I powder. The two layers were placed face-to-face and annealed at 150 °C. This method yielded a perovskite film with full surface coverage. The resulting FPSC, with an active area of 410 mm2, achieved a PCE of 15.07%, demonstrating the effectiveness of vacuum-based processing techniques for large-area flexible devices.
![]() | ||
Fig. 9 (a) Lift-off and transfer process. Reprinted with permission from ref. 141. Copyright 2005, Nature. (b) Cross-sectional SEM image of FPSC and scheme of the flexible device, (c) photograph of FPSC, and (d) J–V curve of the best device. Reprinted with permission from ref. 143. Copyright 2015, Royal Society of Chemistry. |
In parallel, significant efforts have been directed toward developing simpler and more scalable low-temperature methods.144 For example, Yang et al.144 employed 1-benzyl-3-methylimidazolium chloride as a solid-state ionic liquid (ss-IL) to form ETLs at low temperatures. This approach offers a facile and potentially low-cost alternative for the fabrication of the ETL in FPSCs, as shown in Fig. 10a and b. The energy-level diagram of the device is presented in Fig. 10c. According to the scanning Kelvin probe microscopy (SKPM) measurements, the work function (WF) of ITO decreased from 4.67 eV to 4.32 eV after coating with the ss-IL layer. Similar reductions were observed from 4.41 to 4.02 eV for FTO, 5.10 to 3.93 eV for Au, and 4.60 to 3.84 eV for Ag. These results confirm that the ss-IL coating effectively modified the WF of both the metal oxides (ITO and FTO) and metals (Au, Ag). According to the energy-level alignment, the cathode/perovskite interface lowers the WF of the ss-IL layer, lowering the energy barrier for electron extraction, and thus improving the charge collection. Furthermore, the broad bandgap of ss-IL (∼4.59 eV) served as an effective hole-blocking layer, reducing the hole diffusion into the ITO and enhancing the fill factor (FF) and JSC. The broad band gap, high electron mobility, and well-aligned the WF of ss-IL-based ETLs contributed to an enhanced PCE to 16.09% (Fig. 10d).
![]() | ||
Fig. 10 (a) Device structure, (b) cross-sectional SEM of FPSC with ss-IL as ETL, (c) energy-level, and (d) J–V curve of the best device. Reprinted with permission from ref. 144. Copyright 2016, Wiley-VCH. |
Jeong et al.145 used a UV-assisted solution process to prepare Nb-doped TiO2 (UV-Nb:TiO2) ETLs at low temperature (<50 °C). The prepared TiO2 nanocrystals (NCs) stabilized with oleic acid spin-coated, and then treated with UV (Fig. 11a). After UV treatment, the prepared UV-TiO2 displayed higher crystallinity, as indicated by the HR-TEM images (Fig. 11c and d). Due to its photocatalytic activity, the UV-treated TiO2 thin-film (UV-TiO2) degraded oleic acid, and thus the prepared TiO2 NC film displayed highly uniform, high compact high transmittance (Fig. 11b) and better-blocking effect compared to that for TiO2 film prepared at high temperature. Moreover, Nb doping enhanced the conductivity and improved the charge extraction by shifting the conduction band downward. As a result, the FPSCs using PEN/ITO and UV-Nb:TiO2 achieved a PCE of 16.01%. Moreover, the assembled devices displayed higher stability even after 1000 bending cycles with a radius of 15 mm. Further enhancement was achieved by Yang et al.,146 employing the sol–gel method for the preparation of a compact TiO2 (c-TiO2) ETL below 150 °C. The results confirmed that the c-TiO2 ETL exhibited the best photoelectric performance with the concentration of TiO2 precursor solution of 2 M and the annealing temperature of 150 °C for 30 min. The best PCE of the assembled FPSC on a PEN/ITO substrate was 6.11%, with good mechanical stability. Recently, Gu et al.147 prepared TiO2-doped carbon nanofibers (TiO2/C NFs) via electrospinning as ETLs for FPSCs. The related constructed devices exhibited a PCE of 17.61%, with high mechanical stability, highlighting the promising application of TiO2/C NFs in developing high-efficiency FPSCs.
![]() | ||
Fig. 11 (a) Scheme of the UV process for the formation of a compact TiO2 thin film and (b) FTIR spectra of spin-coated TiO2 thin films. HR-TEM images of oleic acid-capped TiO2 NCs (c) before and (d) after UV treatment. (e) J–V curves of the FPSCs with various ETLs on FTO glass. (f) Normalized photovoltaic parameters of the FPSC device after bending with a radius of 15 mm for 1000 cycles. (g) Photograph of the FPSC based on with UV-Nb:TiO2 ETL coated on a ITO/PEN substrate. Reprinted with permission from ref. 145. Copyright 2016, Elsevier. |
Despite the great progress that has been achieved utilizing TiO2 as ETLs in FPSCs, its lower electron mobility compared to conventional ETLs still a challenge.144 As a result, intensive research has been devoted to developing alternative metal oxides with improved charge transport properties. Among them, zinc oxide (ZnO) has been highlighted as a potential alternative ETL in FPSCs to replace the conventional TiO2 due to its excellent properties such as higher electron mobility and easy processability from solution at low temperature without a sintering process.148–150 The first work applying ZnO as the ETL in FPSCs was reported by Kumar et al.,69 who utilized the chemical bath deposition (CBD) method to prepare a ZnO compact layer. The FPSCs prepared using the ZnO ETL showed a PCE of 2.62% on a flexible PET/ITO substrate. Higher improvements were achieved by Liu and Kelly,148 where they deposited ZnO films onto ITO substrates using the spin-coating technique, and the whole the structure of device is presented in Fig. 12a, and its photograph is shown in Fig. 12b. According to the energy-level alignment shown in Fig. 12c, the photogenerated charge carriers within the perovskite layer can be efficiently separated, in which electrons can be transferred to the ZnO layer, while holes can move to the spiro-OMeTAD HTL. The device constructed on a flexible PET/ITO substrate displayed a PCE of 10.2% (Fig. 12d). Further enhancement was achieved by Chu et al.,151 where they modified the surface quality and wettability of the ZnO ETL by an ionic liquid at room temperature (Fig. 12g and h). The modification greatly improved the charge mobility of the ZnO ETL and improved the crystallinity of the perovskite film. The perovskite film on the modified ZnO ETL (Fig. 12f) presented larger and more homogeneous grains with better crystallinity on the pristine substrate (Fig. 12e). Thus, the assembled device on a flexible substrate displayed a PCE of 12.1%.
![]() | ||
Fig. 12 (a) FPSC device structure, (b) photograph of ITO/ZnO/CH3NH3PbI3/spiro-OMeTAD/Ag device on a PET substrate, (c) energy levels of the different parts of the device, and (d) J–V curves of the FPSC under illumination/in the dark. Reprinted with permission from ref. 148. Copyright 2013, Springer Nature. SEM image of the MAPbI3 perovskite film on (e) pristine ZnO ETL, (f) modified ZnO ETL, (g) cross-sectional SEM image and (h) photograph of the assembled FPSC. Reprinted with permission from ref. 151. Copyright 2018, Elsevier. |
SnO2 is another promising alternative ETL to TiO2. Typically, SnO2 exhibits a wide optical band gap (3.6–4.0 eV), high transparency, high mobility, excellent chemical stability, and easy low-temperature preparation.152 To date, much research has been done on utilizing SnO2 as the ETL in FPSCs. For example, Park et al.153 developed an Li-doped SnO2 (Li:SnO2) ETL at low temperature. The doped Li remarkably enhanced the conductivity and lowered the CBM of SnO2, which enhanced the electron injection and transport from the perovskite layer. The assembled FPSC achieved a PCE of 14.78%, with higher mechanical durability after 500 bending cycles at a radius of 10 mm, which showed a retention of 91.9% of the initial values. Furthermore, Shi et al.152 developed a simple method to synthesize SnO2 nanoparticles (NPs) at room temperature, followed by spin coating on a flexible substrate, and finally the SnO2 thin-film was treated with UV-ozone. The assembled FPSC device displayed a PCE of 15.27%. Similarly, SnO2 films were fabricated using plasma-enhanced atomic layer deposition (PEALD) for use as ETLs in FPSCs.154 Notably, water vapor treatment significantly enhanced the charge transport properties of the SnO2 layer, leading to a PCE of 18.36% for the assembled devices. A higher performance was achieved by Zhong et al.,155 where they prepared metal ion-modified SnO2 (M-SnO2) via a hydrolysis process and coated it on a flexible substrate at room temperature to prepare an RT-SnO2 ETL-based FPSC (Fig. 13a). Doping SnO2 with metal ions remarkably enhanced the charge extraction and suppressed the interfacial charge recombination. As a result, the constructed FPSCs achieved a PCE of 19.3% (Fig. 13b and c), together with outstanding strong mechanical stability (Fig. 13d). Further enhancement was achieved by Paik et al.,156 where they prepared an SnO2–TiO2 hybrid as the ETL. The related fabricated FPSC device with an SnO2–TiO2 ETL displayed strong mechanical reliability due to its strong adhesion to the substrate. The FPSC device with the SnO2–TiO2 ETL achieved a PCE of 21.02%. Recently Long et al.157 achieved a higher performance by utilizing SnO2 embedded in gold nanoparticles (Au NPs) ETL. The presence of Au NPs significantly improved the conductivity and electron mobilities of the ETL, which promoted electron extraction and transport, resulting in a decrease in charge recombination at the ETL/perovskite interface. Therefore, the assembled FPSC device with the SnO2 ETL embedded with Au NPs showed a PCE of 23.08%, which is higher than the FPSCs with the pristine SnO2 ETL (21.65%).
![]() | ||
Fig. 13 (a) Scheme of the FPSC device, (b) J–V curve, (c) stabilized current density and PCE, and (d) normalized PCE as a function of bending cycles at a radius of 5 mm. Reprinted with permission from ref. 155. Copyright 2022, Wiley-VCH. |
In addition to metal oxides, a variety of organic materials has been utilized as ETLs, owing to their ease of preparation at low temperatures; an essential requirement for the fabrication of flexible devices. Fullerene-based materials have emerged as promising ETLs in FPSCs due to their excellent electron mobility, solution-processability, and potential for interface engineering.158 Recent advancements have focused on enhancing the performance and stability of these materials through structural modifications and hybrid compositions.159 For example, Yoon et al.160 fabricated a hysteresis-free planar CH3NH3PbI3 PSC with vacuum-processed C60 ETL deposited on a PEN/ITO substrate at room-temperature without the hole blocking layer (Fig. 14a). They reported that the C60 layer deposited on perovskites reduced the photocurrent hysteresis, and thus enhanced the overall efficiency of the hysteresis-free FPSC, achieving a PCE of 16.0% (Fig. 14b). The device retained 95% of its original power conversion efficiency after 100 bending cycles at a bending radius of 5 mm. However, after 1000 cycles, the efficiency dropped by 20% compared to its initial value (Fig. 14c). A further enhancement was achieved by Wang et al.,161 where they highlighted the use of a carboxyl-functionalized fullerene, C60 pyrrolidine tris-acid (CPTA), as the ETLs in n-i-p planar FPSCs. CPTA forms a uniform film that covalently binds to the ITO substrate, reducing the hysteresis and improving the mechanical flexibility. Because CPTA is solution-processable, it supports the development of lightweight, FPSCs. The devices using ITO/CPTA/CH3NH3PbI3/spiro-OMeTAD/Au structures on flexible substrates achieved a PCE of 17%. Recently, Hou et al.162 reported that introducing a cross-linkable fullerene (FTAI) enhances the conductivity and elasticity of grain boundaries in tin-based perovskite films. The resulting devices exhibited a PCE of 14.91% and maintained 90% of their initial efficiency after 10000 bending cycles, demonstrating excellent mechanical stability. Thus, these materials offer favorable energy level alignment and efficient electron extraction, making them attractive candidates for use in FPSCs.
![]() | ||
Fig. 14 (a) Scheme of the FPSC with a cross-sectional SEM image, (b) J–V curve, and (c) efficiency as a function of bending cycles with radii of 10 mm and 5 mm. Reprinted with permission from ref. 160. Copyright 2016, Royal Society of Chemistry. |
Polymers offer numerous benefits as charge-transporting materials for flexible PSCs, such as great mechanical flexibility, customizable optoelectrical characteristics by adjusting their chemical structures, and low-temperature solution processability.166 To date, many materials have been successfully utilized as HTL materials. Polymers display potential ability as HTLs due to their low-temperature solution processability, tunable optoelectrical properties, and high mechanical flexibility. Poly(3,4-ethylenedioxythiophene)–poly(styrenesulfonate) (PEDOT:PSS) is a promising candidate due to its high work function (WF), which is around 5.0 to 5.2 eV.78,167 A device with the PET/ITO/PEDOT:PSS/perovskite/PCBM/Al structure demonstrated a PCE of 9.20%.168 In further work by Pak et al.,90 they deposited PEDOT:PSS on a flexible substrate polymer (Noland Optical Adhesive 63) to be utilized as the HTL for FPSCs with an inverted architecture (Fig. 15a and b). The capacity of the device to restore its shape after random crumpling is shown in Fig. 15c. It was discovered that following annealing on a hot plate for 10 s at 80 °C, the crumpled device was fully recovered. Fig. 15d and e demonstrate that the performance of the device deteriorated as a result of cracks created by particularly severe bending. There are two types of crumpled regions that can be distinguished by the degree of bending strains induced by random crumpling, low-strain (r ≥ 1 mm) and high-strain (r < 1 mm) applied regions (Fig. 15d). Multiple cracks were discovered in the high-strain applied region, but no break was seen in the low-strain applied zone (Fig. 15e). The PCE of the assembled device before and after bending at a radius of 1 mm was 10.75% and 10.4%, respectively. After crumpling, the PCE decreased from its initial value of 10.2% to 6.1% upon crumpling recovery (Fig. 15c).
![]() | ||
Fig. 15 (a) Scheme of the stacked layers on an NOA63 film and a cross-sectional SEM image of the device, (b) photographs of the device during the crumpling test, (c) J–V curves before crumpling and after shape recovery from crumpling test, (d) SEM and (e) magnified SEM images of the top surface after shape recovery from crumpling test showing regions of low strain and high strain. Reprinted with permission from ref. 90. Copyright 2015, Wiley-VCH. |
However, despite these tremendous advancements, the performance of PEDOT:PSS-based flexible PSCs is still limited due to some negative characteristics. (i) The PEDOT:PSS film has poor charge-transporting ability in the out-of-plane direction because it forms lamellar structures, (ii) its insufficient work function compared with the VBM of perovskite (−5.4 eV) results in energy loss at the PEDOT:PSS/perovskite interface, and (iii) PEDOT:PSS displays a strong acidic nature (pH ≈ 1), which degrade the device layers, and thus destroys the total performance.169–171 Thus, its highly important to develop PEDOT:PSS-based HTLs or develop new polymeric HTLs for the further improvement of FPSCs. Doping PEDOT:PSS with other materials has shown promising potential to improve the performance of the assembled FPSCs. For instance, Hu et al.19 prepared nano-cellular-doped PEDOT:PSS (NC-PEDOT:PSS) HTL in FPSCs. They reported that the modification significantly enhanced the light absorption and charge transport. The related FPSC devices displayed a PCE of 12.32% on a large area of 1.01 cm2 and outstanding flexural endurance.
One major cause of energy loss at the PEDOT:PSS/perovskite interface, which reduces the VOC, is the relatively low WF of PEDOT:PSS (4.9–5.2 eV) compared to the perovskite VBM (∼5.4 eV).172 Thus, to overcome this, Lee et al. developed PSS-g-PANI, a water-soluble material that serves as an efficient HTL in FPSCs (Fig. 16a). The devices using PSS-g-PANI exhibited higher PCEs than that with PEDOT:PSS HTL (Fig. 16b).172
![]() | ||
Fig. 16 (a) Chemical structure of PSS-g-PANI and (b) J–V curves of the devices with PEDOT:PSS, PSS-g-PANI, and PSS-g-PANI: PFI HTLs. Reprinted with permission from ref. 172. Copyright 2016, Wiley-VCH. (c) Flexible structure, (d) energy level, and (e) J–V curves of the FPSC with PEDOT:PSS. Reprinted with permission from ref. 173. Copyright 2017, Elsevier. |
Recently, intense research has focused on replacing PEDOT:PSS with small molecules, organic polymers, and inorganic materials.173,174 Based on their tunable photophysical properties and easy preparation at low temperatures, small organic molecules have received great interest as HTLs in FPSCs.173 For example, N-(4-(9H-carbazol-9-yl)phenyl)-7-(4-(bis(4-methoxy-phenyl)amino)phenyl)-N-(7-(4-(bis(4-methoxyphenyl)amino)-phenyl)-9,9-dioctyl-9H-fluoren-2-yl)-9,9-dioctyl-9H-fluoren-2-amine (CzPAF-TPA) as the HTL demonstrated high hole mobility and an appropriate highest occupied molecular orbital (HOMO) energy level, as can be observed in Fig. 16d.173 The structure of the flexible PEDOT:PSS-based FPSC device is presented in Fig. 16c. The J–V characteristics of the FPSCs with CzPAF-TPA and PEDOT:PSS as dopant-free HTMs are displayed in Fig. 16e. The assembled FPSC device displayed a PCE of 12.46%. Alternatively, poly(triaryl amine) (PTAA) as the HTL significantly enhanced the VOC of the FPSC device with a structure of poly-ethylene naphthalate (PEN)/ITO/ZnO/MAPbI3 perovskite/PTAA/Au. The PCE of the FPSC was 15.6%.175
Conjugated polyelectrolytes are a novel class of interface materials consisting of conjugated side chains with ionic functional groups and backbones. The strong solubilities of these materials in polar solvents allow them to be employed in low-temperature solution procedures to produce thin films on surfaces. Conjugated polymers have been successfully utilized as HTLs for FPSCs. For instance, Jo et al.176 used 1,4-bis(4-sulfonatobutoxy)benzene and thiophene moieties (PhNa-1T) (Fig. 17a) as HTLs in inverted FPSCs (Fig. 17b). The incorporation of PhNa-1T into the HTL of the FPSC device significantly enhanced the charge extraction from the perovskite absorber to the HTL, suppressing charge recombination in the bulk perovskite and HTL/perovskite interface. Accordingly, the assembled FPSC device with PhNa-1T displayed a PCE of 14.7% (Fig. 17f), as well as high stability under ambient conditions (Fig. 17g).
![]() | ||
Fig. 17 (a) FPSC architecture, (b) chemical structure of PhNa-1T, (c) energy diagrams, (d) AFM images of PEDOT:PSS and PhNa-1T films deposited on ITO/PEN, (e) transmission spectra of PEDOT:PSS and PhNa-1T films deposited on ITO/PEN, (f) J–V curves of the FPSCs with PEDOT:PSS and PhNa-1T HTLs, and (g) PCE changes in the FPSCs with PEDOT:PSS and PhNa-1T HTLs as a function of exposure time to ambient atmosphere. The inset shows photographs of the corresponding degraded FPSCs. Reprinted with permission from ref. 176. Copyright 2019, Wiley-VCH. |
Inorganic materials also offer a promising alternative to PEDOT:PSS, owing to their excellent chemical stability, low production costs, and high mobility.177 Among these metal oxides, NiOx displayed promising ability as an HTL material due to its broad bandgap, deep valence band edge (∼5.4 eV), and high stability. However, the use of NiOx as the HTL in flexible devices is limited given that the typical NiOx layers require high temperatures (>300 °C) to prepare high-quality thin films.178,179 Thus, great efforts have devoted to overcoming this limitation.180 NiOx nanoparticles are usually dispersed in water and coated on a flexible substrate through solution methods such as spin-coating and spray-coating techniques. For example, Yin et al.177 reported that an NiOx nanoparticle solution spin-coated onto ITO-PEN substrates at 130 °C for 20 min in air could be sufficiently utilized as the HTL for FPSCs (Fig. 18b). The prepared FPSC device displayed a PCE of 13.43% (Fig. 18a), which is close to that achieved with the rigid device using an NiOx film prepared at higher temperature (16.47%). This indicates the potential ability of the NiOx films as HTLs for FPSCs. Furthermore, Ye et al.181 deposited NiOx nanoparticles onto a PET substrate via a low-temperature soft-cover deposition (LT-SCD) method to be utilized as HTL for FPSCs. The assembled FPSC device achieved a PCE of 15.3%. A higher performance was achieved by Najafi et al.182 In this study, NiOx and ZnO nanoparticle films were employed at room temperature as the HTL and ETL, respectively. The assembled FPSC by the flexible PEN/thin film barrier/ITO substrates displayed PCEs of 16.6% (Fig. 18d and e). More interestingly, about 85% of its output efficiency is retained after 1000 h.
![]() | ||
Fig. 18 (a) J–V curves of the devices based on NiOx layers, (b) cross-sectional SEM image of the NiOx-based FPSC, and (c) AFM images of ITO and the NiOx film on the ITO substrate. Reprinted with permission from ref. 177. Copyright 2016, American Chemical Society. (d) Architecture of the inverted planar PSCs and (e) energy levels of the layers. Reprinted with permission from ref. 182 Copyright 2018, Wiley-VCH. |
In conclusion, because of their exceptional solution processability at room temperature, NiOx shows promise as HTLs for FPSC devices. Nevertheless, low-temperature processing frequently results in the development of defects, which significantly reduce the overall performance of the device and the quality of the film. Thus, great attempts have been made to alter NiOx nanoparticles to enhance the overall performance of FPSCs. For instance, Wang et al.183 passivated the surface defects of NiOx NPs using benzoic acid self-assembled monolayers (SAMs). This study demonstrated that 4-bromobenzoic acid could effectively perform the function of surface passivation. This SAM layer decreased the trap-assisted recombination, minimized the energy offset between the NiOx NPs and perovskite, and modified the HTL surface wettability. Thus, the perovskite crystallization was improved and more stable PSCs were produced with an improved overall performance. The fabricated FPSCs on a PET substrate displayed a PCE of 16.2%. Later, He et al.184 successfully prepared high-quality, phase-pure Cu-doped NiOx nanoparticles using a chemical co-precipitation method. The insertion of Cu improved the electrical conductivity, the work function and densely packed and pinhole-free morphology of the NiOx film. By employing the Cu-doped NiOx HTL, the assembled FPSC device displayed a PCE of 16.96%.
Flexible glass coated with ITO has been employed for FPSC applications. This is due to its excellent thermal stability (>600 °C), chemical stability, electrical conductivity, and superior gas barrier properties. When manufactured at thicknesses below a few hundred micrometres, flexible glass retains these characteristics while gaining mechanical flexibility. The first report utilising ultrathin glass in FPSCs was published in 2015 by Tavakoli et al.29 They used a 50 μm-thick glass substrate to fabricate an FPSC device, which maintained 96% of its initial PCE (12.06%) after 200 bending cycles. The PCE was further enhanced to 13.14% by introducing an antireflection film with outstanding superhydrophobic properties. More recently, Sergio et al.186 reported a PEC of 22.6% under 400 lx light-emitting diode (LED) illumination for FPSCs based on ultrathin ITO-coated flexible glass. The specific W g−1 of these devices was found to be an order of magnitude higher than that of the devices based on rigid glass and approximately 40–55% greater than that of the FPSCs using PET substrates. This highlights the significant potential of flexible glass for powering next-generation indoor electronics. However, its brittleness, relatively heavy weight, and cost remain big limitations. Metal foils are another candidate as FPSC substrates, offering excellent thermal stability, conductivity, and corrosion resistance. However, their low optical transmittance limits light absorption, thereby reducing the device efficiency.187 In contrast, plastic substrates (e.g., PET and PEN) exhibit high optical transparency, excellent flexibility, and good chemical resistance. However, their main drawbacks are low thermal tolerance and poor gas barrier properties.188 Despite these limitations, plastic-based FPSCs are currently regarded as highly promising. The choice of electrode material that interfaces with the flexible substrate also plays a critical role in the device performance. Various materials have been investigated, including silver nanowires (Ag-NWs), Al-doped ZnO (AZO), ITO, carbon nanotubes (CNTs), graphene, and organic materials. ITO remains one of the most widely used electrode materials in FPSCs. However, its high cost and poor mechanical flexibility limit its suitability for commercial flexible devices, despite its ability to yield high PCEs on PET substrates.133 Thus, research has been focused on developing alternative materials.
In search of alternatives, Ag-NWs have shown great promise in FPSCs due to their outstanding optical and electrical properties, together with compatibility with solution-based fabrication techniques.189 For example, Lee et al.190 demonstrated that the FPSC device assembled with Ag-NWs spray-coated on top spiro-MeOTAD as the HTL achieved a PCE of 7.45%. The SEM image of the Ag-NWs is presented in Fig. 19c. By coating Ag-NWs as the top electrode on titanium foil, as shown in Fig. 19a, and the cross-sectional SEM of the whole structure FPSC device (Fig. 19b), the PCE reached 7.58%.190 The relatively low performance was largely attributed to the inferior conductivity of the ITO used for comparison. To improve the stability of Ag-NWs, Lee et al. developed a transparent electrode by sandwiching the Ag-NW layer between two amorphous aluminium-doped zinc oxide (a-AZO) layers (Fig. 19d). The fabricated device based on AZO/AgNW/AZO composite electrodes formed a pinhole-free structure (Fig. 19e and f). The SEM images of the c-AZO/AgNW/AZO and a-AZO/AgNW/AZO composite electrodes (Fig. 19g–i) showed that the c-AZO and a-AZO top layers uniformly covered the AgNWs. The AgNW network structure was well preserved, with only a slight bulge at the junctions due to the annealing process at 190 °C. However, as seen in the individual c-AZO and a-AZO layers, the c-AZO/AgNW/AZO composite exhibited a rough surface with significant porosity, while the a-AZO/AgNW/AZO composite had a smoother surface texture. The FPSC assembled with the a-AZO/Ag-NW/a-AZO/PES configuration achieved a PCE of 11.23%, which is comparable to that of ITO/PEN-based devices (Fig. 19j). Moreover, this device retained 94% of its initial PCE after 400 bending cycles with a 12.5 mm bending radius (Fig. 19k).191
![]() | ||
Fig. 19 (a) Scheme, (b) cross-sectional SEM of the FPSC with Ti metal coated with a Ag-NW top electrode, and (c) SEM images of the spray-coated Ag-NWs. Reprinted with permission from ref. 190. Copyright 2016, Wiley-VCH. (d) Scheme of the preparation of a-AZO/Ag-NW/AZO electrodes. (e) Scheme of the FPSC structure based on the AZO/AgNW/AZO electrode. (f) Cross-sectional SEM image of the FPSC device. SEM images of (g) AgNW/AZO, (h) c-AZO/AgNW/AZO, and (i) a-AZO/AgNW/AZO electrodes, (j) J–V hysteresis characteristics curves, and (k) the mechanical bending test for FPSCs on ITO/PEN and a-AZO/Ag-NW/AZO/PES electrodes. Reprinted with permission from ref. 191. Copyright 2017, Wiley-VCH. |
PEDOT:PSS is widely used in FPSCs due to its high conductivity, excellent optical transmittance, uniform film coverage, and ease of fabrication through solution-processing techniques. FPSC devices incorporating PEDOT:PSS have demonstrated promising mechanical durability.192 For instance, Poorkazem et al. developed a transparent electrode composed of PEDOT:PSS and In2O3 on a PET substrate, as illustrated in Fig. 20a.192 The SEM images of the fabricated electrodes after undergoing 2000 bending cycles around a cylinder with a 4 mm radius are presented in Fig. 20b–e. It can be seen that the In2O3-based electrodes exhibited visible cracking (Fig. 20b and c), whereas the PEDOT:PSS layer retained its performance (Fig. 20d and e). The cracking in the metal-based electrodes led to increased sheet resistance, and consequently a reduction in the overall device performance. However, despite its advantages, PEDOT:PSS is known to corrode the bottom substrates over time, which adversely affects the long-term stability of FPSCs. As an alternative, graphene has gained attention for its exceptional chemical stability, superior optical transparency, high charge carrier mobility, and low electrical resistance; qualities making it a compelling candidate for use in FPSC electrodes.193
![]() | ||
Fig. 20 (a) Scheme of the M-In2O3/ZnO/CH3NH3PbI3 (top) and HC-PEDOT/SC-PEDOT/CH3NH3PbI3 (bottom) devices. SEM images of (b) PET/M-In2O3, (c) PET/M-In2O3/ZnO/CH3NH3PbI3, (d) PET/HC-PEDOT, and (e) PET/HC-PEDOT/SC-PEDOT/CH3NH3PbI3 films after 2000 bending cycles. Reprinted with permission from ref. 192. Copyright 2015, Royal Society of Chemistry. (f) Scheme diagram and (g) J–V curves of the FPSC with/without the ZEOCOAT™ layer. PCE and JSC of the FPSC with different bending cycles and (h) variation in PCE and JSC of the FPSC by bending radius. Reprinted with permission from ref. 194. Copyright 2016, Elsevier. FPSC-based on Gr-Mo/PEN: (i) device structure, (j) cross-sectional SEM image of complete device, (k) J–V curves, and (l) resistance change of the Gr-Mo/PEN and ITO/PEN films by bending at R = 4 mm, and (m) photograph of Gr-Mo/PEN substrate. Reprinted with permission from ref. 77. Copyright 2017, Royal Society of Chemistry. |
Zhike et al.194 first report the use of graphene as a transparent electrode in FPSCs. They fabricated a device with PET/graphene/poly(3-hexylthiophene) (P3HT)/MAPbI3/PC71BM/Ag structure (Fig. 20f). The graphene layer was made using chemical vapour deposition (CVD) and transferred to flexible PET substrates, with the bottom electrodes modified by P3HT films to act as HTLs. As presented in Fig. 20g, the assembled device with the ZEOCOAT™ layer on the PET substrate demonstrates a PCE of 11.5%, which is better than that achieved by the device without ZEOCOAT™ (10.4%). This enhancement is mainly assigned to the rough surface of the PET substrate. After 500 bending cycles with a bending radius of 0.175 cm, the PCE decreased by 14% due to the reduction in the JSC, as shown in Fig. 20h. The morphology of the perovskite layer, graphene electrode, and Ag top electrode in the control devices showed minimal alteration after the bending tests, which is a key factor contributing to the stability of the fabricated devices. Notably, the decline in PCE is closely linked to the reduction in JSC. In contrast, both the fill factor (FF) and VOC experienced only minor changes during the bending tests, suggesting that the resistance of the graphene electrode remained unaffected. In a further study by Yoon et al.,77 they demonstrated that the chemical bonding between the graphene layer and PET substrate is significantly responsible for the enhanced performance of FPSCs. In their study, an MoO3-modified graphene layer prepared by the CVD method was employed as a transparent anode. The structure of the assembled FPSC device is illustrated in Fig. 20i, where CH3NH3PbI3 is the perovskite layer, PEDOT:PSS is the HTL, fullerene (C60) is the ETL, and bathocuproine (BCP) is the hole blocking layer (HBL). The entire fabrication process was carried out at a relatively low temperature (<110 °C). According to the cross-sectional SEM image of the Gr-Mo/PEN device (Fig. 20j), the PEDOT:PSS layer was well formed on the hydrophobic graphene surface, given that the 2 nm-thick MoO3 interfacial layer rendered the hydrophobic graphene surface sufficiently hydrophilic. As a result, the subsequent perovskite layer was also uniformly fabricated on the Gr-Mo/PEN substrate. In addition, the surface of the perovskite layers was observed to be quite smooth. The PCE of the device increased to 16.8% (Fig. 20k). As shown in Fig. 20m, the relative resistance for the ITO/PEN film initially increased slightly, but then increased sharply after 50 bending cycles, reaching five times the initial value after 1000 cycles. In contrast, the relative resistance of the Gr-Mo/PEN film increased only 0.4 times its original value after 1000 bending cycles.
Due to their low electrical resistance and excellent optical transparency, carbon nanotubes (CNTs) have received significant attention as promising electrode materials for FPSCs. However, the overall performance remained lower than that of graphene-based flexible devices, primarily due to limitations in CNT film morphology and light transmittance. In parallel, environmental concerns regarding petroleum-based, non-biodegradable polymer substrates have received interest as biodegradable and eco-friendly alternatives for FPSCs. Among them, cellulose-based materials have emerged as promising green substrates. Nanocellulose paper (NCP), which can be extracted from abundant cellulose-rich biomass, exhibits high transparency together with respectable mechanical, thermal, and chemical stability. Moreover, it is biodegradable, cost-effective, and can be derived from biowaste. Nevertheless, the water solubility of NCP in aqueous solutions compromises its structural integrity, limiting its direct application in electronic devices.28 Thus, to address this issue, Huang et al.195 successfully extracted NCP from a viscous solution of nanocellulose of cotton and coated it with acrylic resin to form a waterproof layer, as described in Fig. 21a (steps 1–3), and then FPSCs were prepared (steps 4–6). The assembled device achieved a PCE of 4.25% (Fig. 21b), with a specific power of 0.56 W g−1. Furthermore, the device retained over 80% of its initial efficiency after 50 bending cycles (Fig. 21c), indicating its satisfactory mechanical stability.
![]() | ||
Fig. 21 (a) Scheme of the process of NCP-based substrate and NCP-FPSC, (b) J–V curves of NCP-FPSC, and (c) normalized PCE-bending cycles state on a bottle with a diameter of 15 mm. Reprinted with permission from ref. 195. Copyright 2019, Springer Nature. |
Further progress was reported by Zhu et al.,196 who extracted cellulose nanofibrils (CNFs) from bamboo to create b-CNF substrates (Fig. 22a). Subsequently, a transparent conductive indium zinc oxide (IZO) layer was sputtered onto the b-CNF substrate. The resulting b-CNF/IZO electrode exhibited high visible-range transmittance (∼85%), low surface roughness, excellent conductivity (sheet resistance of 41 Ω sq−1), and robust mechanical properties. The structure, morphology, and photovoltaic characteristics have been reported. The structure of the FPSC device based on the b-CNF substrate is CNF/IZO/PEDOT:PSS/perovskite/PCBM/Ag is presented in Fig. 22b. The coated perovskite layer showed a highly compact, large grain size, and monolithically grained films (Fig. 22c and d), which is conducive for the transportation of carriers and suppression of their recombination. The assembled FPSC achieved a PCE of 11.68% (Fig. 22e), the highest reported value among biomass-based PSCs. Additionally, the device retained approximately 70% of its original efficiency after 1000 bending cycles with a 4 mm curvature radius (Fig. 22f).
![]() | ||
Fig. 22 (a) Scheme of the preparation of b-CNF electrodes, (b) scheme of the FPSCs, (c) cross-section SEM image of the perovskite films based on the b-CNF electrode, (d) top-view SEM image of the perovskite film, (e) J–V curve of the b-CNF-based FPSCs, and (f) main parameter variation of the FPSCs upon bending tests at a 4 mm radius. Reprinted with permission from ref. 196. Copyright 2019, Wiley-VCH. |
Mechanical stability under bending stress is a critical concern for FPSCs. Seok et al.200 reported that after 300 bending cycles, the PCE of FPSCs fabricated on PEN/ITO substrates decreased by 5%. Similarly, Jung et al.143 observed that after 1000 cycles at a 10 mm bending radius, the PCE decreased from 12.20% to 6.10%, indicating significant mechanical degradation.
According to Carlo et al.,201 an ideal radius for bending ITO is 14 mm, while a smaller radius causes the ITO layer to crack. Furthermore, in the study by Liu et al.,202 they employed PET/ITO-based FPSCs to examine the inherent mechanical stability of the FPSCs with various bending radii. They reported that at a radius of 14 mm, there was no noticeable reduction in PCE after 500 cycles, while smaller radii resulted in substantial performance losses. Thus, to address these challenges, Yang et al.76 introduced a novel electrode structure by incorporating a thin silver interlayer between two ITO coatings on PET. This approach enabled a reduction in the overall ITO thickness, while enhancing both the electrical performance and mechanical flexibility. The FPSC achieved a PCE of 18.4% and maintained 83% of their original efficiency after bending with a 4 mm radius.
![]() | ||
Fig. 23 (a) Cross-sectional SEM image and (b) photographs of FPSC with Li:SnO2 ETL. Reprinted with permission from ref. 153. Copyright 2016, Elsevier. (c) Scheme of the fabrication of the textile-based FPSC with SnO2 and PCBM ETL and the corresponding cross-sectional SEM image, (d) photographs of the textile-based FPSC and (e) commercial LED lit by the fabricated textile-based FPSC under 0.8 sun illumination. Reprinted with permission from ref. 212. Copyright 2017, Royal Society of Chemistry. |
FPSCs devices are also used in self-powered wearable devices, which consist of energy harvesting and storage integrated systems. For example, Chao et al.213 designed an FPSC-photo-rechargeable lithium-ion capacitor (LIC) for self-powering wearable strain sensors (Fig. 24a). A schematic illustration of the fabricated FPSC device is displayed in Fig. 24b and c. According to the photograph in Fig. 24d, it can be seen that the assembled device was directly worn on a live beetle, demonstrating favorable wearability. The PCE of the single FPSC device with the configuration of PET/ITO/NiOx/MA1−γFAγPbI3−xClx/PCBM/BCP/Ag was 14.01%, while the value was 8.41% when FPSC and LIC were integrated into a single system (Fig. 24e). Additionally, the system delivered a high output voltage of 3 V at a discharge current density of 0.1 A g−1. A module exhibited excellent mechanical durability, maintaining its performance after 500 bending cycles (Fig. 24f). They constructed a flexible LIC based on an LTO/rGO anode and an AC cathode (Fig. 24g). A self-powered wearable sensor equipped with solar energy capabilities is shown after device integration (Fig. 24h), emphasizing its combined functions of energy harvesting, storage, and utilization, together with its ability to conform to the skin. This type of device can reliably and continuously monitor both subtle and more pronounced physiological signals without relying on an external power supply. Fig. 24i demonstrates that the PSC-LIC system achieved a consistently high and stable overall efficiency throughout photo-charging and discharge cycles, reaching a peak of 8.41% and averaging 8.19% at a current density of 0.1 A g−1. As the applied current density increased, the overall efficiency gradually declined from 8.19% at 0.05 A g−1 to 6.70% at 1 A g−1. Moreover, the LIC component within the integrated system showed an average energy storage efficiency exceeding 80%.
![]() | ||
Fig. 24 (a) Scheme of the integrated FPSC-LIC system, (b) scheme of the FPSC structure, (c) cross-sectional SEM of the FPSC device, (d) digital photograph for the as-prepared PSC unit on a live beetle (inset is the IPCE spectrum of the PSC device), (e) J–V curves of individual FPSC and four FPSCs in series, (f) stability of the photovoltaic parameters with varying bending cycles, (g) scheme of the FLIC device, (h) scheme of the PSC-LIC-sensor integrated system, and (i) Overall efficiency of the PSC-LIC device and the energy storage efficiency of the LIC as a function of the cycle number. Reprinted with permission from ref. 213. Copyright 2019, Elsevier. |
These results highlight the potential of FPSCs in enabling the autonomous operation of wearable electronics. However, the brittleness of perovskite crystals can compromise their mechanical stability and limit their large-scale application. Thus, to address this, Hu et al.19 introduced polystyrene-doped nanocellular PEDOT:PSS (NC-PEDOT:PSS) in the low-temperature process to be utilised as the HTL (Fig. 25a). Owing to the mechanical stress-relieving capability of the nanocellular scaffold, these PSCs demonstrate outstanding repeatability and stability, together with exceptional resistance to bending (Fig. 25b). As shown in Fig. 25c and d, FPSCs incorporating the NC-PEDOT:PSS layer were effectively integrated into a wearable solar power source. A total of 24 solar cell units was utilized to operate a portable fan and a multifunctional electronic watch during various body movements. The prepared NCPEDOT:PSS film displayed three different diameters, as demonstrated by the SEM image (Fig. 25e). A cross-sectional SEM image of the complete device (substrate/transparent electrode/NC-PEDOT:PSS/perovskite/PCBM/Ag) is shown in Fig. 25f. The resulting FPSC device achieved a PCE of 12.32% with excellent mechanical robustness (Fig. 25f–h). According to the SEM images in Fig. 24i and j, the PEDOT:PSS-based films developed visible cracks after 1000 bending cycles, whereas the NC-PEDOT:PSS-based films maintained a uniform structure. This indicates that the NC-PEDOT:PSS scaffold effectively relieved the stress during repeated narrow bending.
![]() | ||
Fig. 25 (a) Fabrication procedure for the NC-PEDOT:PSS and FPSC device, (b) stress release diagram for NC-PEDOT:PSS, (c and d) photographs of the FPSC as a wearable power source, (e) SEM of NC-PEDOT:PSS, (f) SEM cross-sectional image of the FPSC with the NC-PEDOT:PSS layer, (g) J–V curves of the FPSC based on PEDOT:PSS and NC-PEDOT:PSS HTL, (h) normalized average PCE of FPSC devices as a function of bending cycles with a radius of 2 mm, and (i and j) SEM images and finite-element simulation of films based on PEDOT:PSS and NC-PEDOT:PSS under bending. Reprinted with permission from ref. 19. Copyright 2017, Wiley-VCH. |
To further enhance the mechanical resilience, they described a biomimetic crystallization approach for forming high-quality perovskite films with a flexible ‘brick-and-mortar’ design (Fig. 26a). The opposing solubility characteristics of the composite matrix promoted vertically oriented micro-parallel crystal growth and provided the stretchability needed to overcome the ‘cask effect’ in PSCs.214 As illustrated in Fig. 26b, the insoluble poly(styrene-co-butadiene) (SBS) scaffold decreased the number of nucleation sites and facilitated heterogeneous nucleation by lowering the nucleation energy barrier. Simultaneously, the interaction between the soluble polyurethane (PU) and PbX2 (X = I or Br) helped to slow down the crystallization process, leading to the formation of high-quality perovskite in PSCs. According to the SEM images in Fig. 26c and d, the reference perovskite film exhibited a rough surface with an average grain size of approximately 360 nm. In contrast, the film incorporating the SBS–PU biomimetic structure displayed significantly larger crystals, exceeding 700 nm in size. It is well-established that perovskite films are more prone to fracture at the grain boundaries than within the crystal grains themselves. Thus, the devices with the SBS–PU biomimetic displayed a higher PEC than the other devices (Fig. 26e). Furthermore, an FPSC with 56.02 cm2 achieved a PCE of 7.91%, making it suitable as a skin-conforming power source (Fig. 26f). The FPSC module was capable of conveniently charging commercial wearable devices, such as a smartwatch, even during various body movements (Fig. 26g).
![]() | ||
Fig. 26 (a) Brick-mortar structure, (b) structure of wearable FPSCs with chemical structures of SBS and PU. Top-view and cross-section SEM images of (c) reference and (d) SBS–PU-based films. (e) J–V curves of the related devices, (f) I–V curves of a wearable solar-power source with SBS–PU (inset is the photograph of assembled device), and (g) photographs of wearable FPSCs as a power source to power a smartwatch. Reprinted with permission from ref. 214. Copyright 2019, Royal Society of Chemistry. |
In the context of wearable electronics, both stretchability and twistability are crucial design features, and these studies illustrate the promising potential of FPSCs to meet these demands.
However, despite the remarkable achievements, substantial challenges remain. The mechanical durability of FPSCs under repeated deformation and their long-term environmental stability in real-world conditions, particularly under conditions of moisture, temperature fluctuations, and UV exposure, are still inadequate for commercial viability. Moreover, the toxicity concerns related to lead-based perovskites are critical issues that must be addressed to ensure the commercialization and widespread adoption of this technology. Furthermore, the scalability of fabrication methods must be enhanced to meet the demands of large-scale production, while ensuring cost-effectiveness without compromising the device performance. Future research must prioritize the development of strategies to mitigate these challenges. Key areas of focus include the development of lead-free perovskites or low-toxicity alternatives, the optimization of encapsulation techniques for enhanced stability, and the improvement of mechanical resilience through advanced flexible substrates and protective coatings. Additionally, scaling up the fabrication methods such as R-to-R processing and the integration of FPSCs into multifunctional systems will be pivotal for realizing the full potential of this technology.
In conclusion, although FPSCs are poised to revolutionize the solar energy landscape, realizing their full potential will require sustained interdisciplinary efforts across materials science, engineering, and environmental sustainability. By overcoming the remaining barriers, FPSCs can play a pivotal role in the transition to a more sustainable and energy-efficient future, offering scalable, versatile solutions for the renewable energy sector.
PVs | Photovoltaics |
PSCs | Perovskite solar cells |
FPSCs | Flexible perovskite solar cells |
OPV | Organic photovoltaics |
QDSCs | Quantum dot solar cells |
DSSCs | Dye-sensitized solar cells |
PCE | Power conversion efficiency |
TFSCs | Thin-film solar cells |
ETL | Electron transport layer |
HTL | Hole transport layer |
TCO | Transparent conducting oxide |
PET | Poly(ethylene terephthalate) |
PEN | Poly(ethylene naphthalate) |
MA+ | Methylammonium (CH3NH3+) |
FA+ | Formamidine (CH3CH2NH3+) |
CBD | Chemical bath deposition |
PEALD | Plasma-enhanced atomic layer deposition |
PEDOT:PSS | Poly(3,4-ethylenedioxythiophene)–poly(styrenesulfonate) |
CB/M | Conduction band/maximum |
VB/M | Valence band/maximum |
R-to-R | Roll-to-roll |
This journal is © The Royal Society of Chemistry 2025 |