Next Article in Journal
Study of the Influence of Melamine and Expanded Graphite on Selected Properties of Polyurethane Foams Based on Uracil Derivatives
Previous Article in Journal
Effects of Pre- and Post-Processing on Pin-Bearing Strength of 3D-Printed Composite Specimens with Circular Notches
Previous Article in Special Issue
Multifunctional Sensor for Strain, Pressure, and UV Light Detections Using Polyaniline and ZnO Nanostructures on a Flexible Substrate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Polymer Functional Layers for Perovskite Solar Cells

1
Department of Physics, Incheon National University, Incheon 22012, Republic of Korea
2
Department of Advanced Materials Engineering, Kyonggi University, Suwon 16227, Republic of Korea
3
Division of Advanced Materials, Korea Research Institute of Chemical Technology (KRICT), Daejeon 34114, Republic of Korea
*
Authors to whom correspondence should be addressed.
Polymers 2025, 17(19), 2607; https://doi.org/10.3390/polym17192607
Submission received: 31 August 2025 / Revised: 20 September 2025 / Accepted: 24 September 2025 / Published: 26 September 2025
(This article belongs to the Special Issue Polymer Thin Films: Synthesis, Characterization and Applications)

Abstract

Perovskite solar cells (PSCs) are next-generation solar cells; they are replacing silicon-based solar cells due to their higher efficiency, greater cost-effectiveness, and enhanced potential for various applications. Exceeding the efficiency of crystalline silicon-based solar cells, the commercialization of PSCs has driven not only the development of perovskite photoactive materials but also charge transport layer advancements, interfacial engineering, and processing technologies. PSCs were developed later than dye-sensitized solar cells and organic solar cells; the adoption of techniques previously employed in these technologies is significant to enhancing their performance. Among them, polymers are widely employed in perovskite solar cells to facilitate efficient charge transport, provide interfacial passivation, enhance mechanical flexibility, enable solution-based processing, and improve environmental stability. In this review, we highlight the roles of polymer materials as charge transport layers, interfacial layers, and other functional layers for highly efficient and stable PSCs.

1. Introduction

Perovskite solar cells (PSCs) have attracted considerable attention as a renewable energy source due to their rapid increase in efficiency and potential for diverse applications [1,2,3,4]. Since their initial power conversion efficiency (PCE) of 3.8% in 2009, continuous developments in composition engineering of perovskite, charge transport materials, and interfacial engineering have enabled the PCE to reach 27% [1,2,3,4,5,6,7,8]. In the early stages, PSCs were developed on porous TiO2 scaffolds, originally used in dye-sensitized solar cells (DSSCs), with perovskite materials infiltrating the pores and serving as the light-absorbing layer [5,9,10,11]. In addition, dye-sensitized perovskite and liquid-type hole transport layers (HTLs) have been replaced with solid-type layers because solid-state solar cells have many advantages, such as elimination of the liquid electrolyte leakage issue, higher mechanical durability, and better scalability for large-area cells, making them more suitable for commercialization [9,10,11]. In this context, spiro-OMeTAD, originally used in light-emitting diodes and later adopted in solid-state DSSCs, started to be applied as the HTL in PSCs. N. G Park’s group first reported all-solid-state thin film PSCs using submicron thick mesoporous TiO2 with an infiltrated perovskite absorber and spiro-OMeTAD HTL [9].
Owing to their facile fabrication process, planar-type PSCs have been developed, and a wide range of materials originally used in dye-sensitized solar cells, light-emitting diodes, and OSCs began to be incorporated [11,12,13,14,15]. Alternatively, new materials such as metal oxides, organic molecules, and polymers are sometimes developed and applied to other solar cells first and then to PSCs, or vice versa [16,17,18,19]. From these efforts, both mesoporous n–i–p and planar p–i–n structures have achieved comparably high efficiencies, while extensive research is currently focused on enhancing their stability and enabling large-area scalability [8,13,20,21,22]. Most recently, a 20% exceedance of PCE on an area greater than 200 cm2 has been certified and listed in the National Renewable Energy Laboratory (NREL) chart [6,23]. However, high-temperature processes and spin-coating techniques, which are challenging to implement in commercial manufacturing and flexible applications, are still commonly employed [23,24,25,26,27]. Polymers have been utilized to enhance the performance of perovskite solar cells because of their compatibility with large-area fabrication processes, such as low-temperature and printing techniques, and their flexible nature [28,29,30].
In this review, we discuss in detail how polymers can be utilized as various functional layers, such as charge transport, interfacial, electrode encapsulation, and anti-reflection layers, in the construction of highly efficient and stable PSCs. The core challenges presented by polymers for functional layers in PSCs include the need for precise energy-level alignment, stable conductivity control, and reliable doping strategies, while maintaining film uniformity and large-area processability. We briefly discuss the fundamental principles and suitability of using various polymers as various functional layers. Our aim with this review is also to address the potential use of polymer functional layers in all perovskite tandem solar cells to achieve high efficiencies, as well as the necessary improvements required for their further development.

2. Polymers for PSCs

A PSC device structure consists of substrate/bottom electrode/electron transport layer (ETL) or hole transport layer (HTL)/Perovskite/HTL or ETL/top electrode [11]. A configuration of HTL/perovskite/ETL from the bottom is denoted as a p–i–n structure, while ETL/perovskite/HTL is denoted as an n–i–p structure. In the case of PSCs, the early development was based on mesoporous TiO2, leading to the initial advancement of n–i–p structured PSCs, which are referred to as the “normal” structure, while p–i–n structures were developed later and are thus called the “inverted” structure [9,10,11]. For high-efficiency n–i–p structured PSCs, metal oxide-based ETLs, such as SnO2 and TiO2, and small-molecule organic hole transport layers, like Spiro-OMeTAD, have been widely used [11]. With the development of p–i–n structures, polymeric materials such as PEDOT:PSS and PTAA began to be employed due to their high mobility and transparency in the visible light range [12]. In addition to these materials, various polymers are employed as functional layers to improve efficiency and stability and enable large-area fabrication.
In OSCs, polymers have been widely employed to form entire layers including electrodes, HTLs, photoactive layers, ETLs, and interfacial layers [17,19]. In PSCs, except for cases where polymers are added as additives to the photoactive perovskite layer, they are generally used either as individual functional layers or in combination with other materials to form complexes [30,31,32]. Depending on their bandgap, conductivity, and energy levels, polymers are mainly utilized as ETLs or HTLs, and they can also be employed as electrodes, an encapsulation layer, anti-reflection layer, and interfacial layer, such as a passivation layer, as shown in Figure 1. In PSCs, polymer ETLs facilitate charge collection not primarily through high conductivity- based transport, but rather via dipole formation inducing vacuum level shifting, as typically observed with PFN-Br or PEI derivatives. Typical examples of HTLs are P3HT, PTAA and PEDOT:PSS, the HOMO level is between 5.0~5.2 eV, which is well aligned to the perovskite photoactive layer. The conductivity can be enhanced by the doping process; notably, the conductivity of PEDOT:PSS can be enhanced to values exceeding 1000 S cm−1 through appropriate doping, which has rendered it suitable for use as a transparent electrode. [10,11,33]. Moreover, polymers with high charge mobility and minimal absorption in the visible region can also be utilized as recombination layers in perovskite tandem solar cells [34,35,36,37,38].
Therefore, when selecting or engineering polymers for functional layers in PSCs, several molecular design principles are critically important. First, the energy level alignment, particularly the HOMO/LUMO positions, should match those of adjacent perovskite or transport layers to facilitate efficient charge extraction or blocking. Second, the side-chain polarity and functional groups must be tailored to enhance miscibility, film formation, and interfacial compatibility with the perovskite surface. Backbone rigidity and planarity influence charge mobility and film morphology, especially in conducting or semiconducting polymers. Additionally, optical bandgap engineering is essential when transparency is needed (e.g., in electrodes or interlayers), while chemical stability and hydrophobicity are crucial for long-term device durability. An integrated molecular design approach that balances electronic, morphological, and interfacial properties is thus indispensable for optimizing polymer performance across different PSC applications. In the following sections, the utilization of polymers as functional layers will be discussed based on previously reported research and development cases.

3. Polymers as ETLs for PSCs

Polymeric materials have emerged as promising alternatives to conventional fullerene-based ETLs in PSCs, offering distinct advantages in terms of energy level tunability, interfacial defect passivation, processability, and operational stability. In particular, the development of n-type conjugated polymers with tailored backbones and side chains has enabled their successful application as pure ETL materials, without blending with inorganic components. In the course of the initial demonstrations, a thiazole-imide-based conjugated polymer (PDTzTI) was demonstrated to function effectively as an ETL in inverted (p–i–n) architectures [39]. The material exhibited favorable alignment with the conduction band of the perovskite absorber and contributed to strong interfacial interactions through its sulfur and nitrogen atoms, which coordinated with undercoordinated Pb2+ ions. These features have been shown to reduce trap states and improve film morphology, resulting in a PCE of 20.86% and improved long-term stability, highlighting the potential of rational polymer design for ETL functionality (Figure 2a).
Further advancement was marked by the conception of PFNDI, a naphthalene diimide (NDI)-based polymer bearing phosphite ester side chains [40]. The molecular structure promoted solubility and enabled strong interfacial coordination, which facilitated the formation of uniform films and reduced trap-assisted recombination. Devices fabricated with PFNDI achieved efficiencies up to 18.25% and retained over 80% of their performance after prolonged exposure to ambient conditions. Spectroscopic analysis confirmed the presence of P–O–Pb interactions, thereby affirming the chemical basis of the observed passivation effect (as shown in Figure 2b). Another example involves PFN-2TNDI, an amino-functionalized copolymer containing fluorene, thiophene, and NDI units, which was implemented as an ETL in conventional (n–i–p) devices [41]. The incorporation of amino groups improved energy level alignment and enabled strong chemical passivation of undercoordinated lead ions. As a result, the PSCs displayed a PCE of 15.96% and excellent photostability under continuous 365 nm illumination, retaining 75% of their initial performance after 3000 h—significantly surpassing the UV durability of TiO2-based ETLs.
Figure 2. (a) The schematic diagram of the inverted planar perovskite solar cells with polymer electron transport layers and molecular structures of the n-type polymers PDTzTI and PBTI. Adapted with permission from Ref. [39] Copyright 2020, Elsevier. (b) XPS spectra of the perovskite films with and without PFNDI layer for (left) P 2p and (right) I 3d spectra. Adapted with permission from Ref. [40] Copyright 2020, Elsevier. (c) Schematic diagram of the SnO2 colloidal dispersion regulated by PM, and scanning electron microscopy (SEM) images of the SnO2 and SnO2-PM films. Adapted with permission from Ref. [42] Copyright 2022, Royal Society of Chemistry. (d) The chemical structure of PFN-2TNDI and (e) photovoltaic performance of PSCs with different ETLs, and (f) effects of the thickness of PFN-2TNDI ETL on the PCE of PSCs. Adapted with permission from Ref. [43] Copyright 2016, John Wiley and Sons.
Figure 2. (a) The schematic diagram of the inverted planar perovskite solar cells with polymer electron transport layers and molecular structures of the n-type polymers PDTzTI and PBTI. Adapted with permission from Ref. [39] Copyright 2020, Elsevier. (b) XPS spectra of the perovskite films with and without PFNDI layer for (left) P 2p and (right) I 3d spectra. Adapted with permission from Ref. [40] Copyright 2020, Elsevier. (c) Schematic diagram of the SnO2 colloidal dispersion regulated by PM, and scanning electron microscopy (SEM) images of the SnO2 and SnO2-PM films. Adapted with permission from Ref. [42] Copyright 2022, Royal Society of Chemistry. (d) The chemical structure of PFN-2TNDI and (e) photovoltaic performance of PSCs with different ETLs, and (f) effects of the thickness of PFN-2TNDI ETL on the PCE of PSCs. Adapted with permission from Ref. [43] Copyright 2016, John Wiley and Sons.
Polymers 17 02607 g002
Polymer–inorganic hybrid strategies have also proven effective. In one such approach, dendritic poly(amidoamine) (PM) was incorporated into SnO2 colloids to create a polymer-complexed ETL [42]. The presence of PM prevented nanoparticle agglomeration, enhanced surface smoothness, and improved wettability. These morphological and interfacial enhancements facilitated the growth of high-quality perovskite films with reduced trap densities (Figure 2c). Devices utilizing such a hybrid ETL architecture reached a PCE of 22.93% with an open-circuit voltage (Voc) of 1.17 V. These devices exhibited remarkable stability under ambient conditions. Additionally, conjugated polymers based on polyfluorene and NDI frameworks, functionalized with amino groups, have been shown to suppress hysteresis and enhance open-circuit voltage by improving contact quality and energy alignment at the electron-collecting interface [43]. In comparison with metal oxide ETLs, these polymeric systems demonstrated superior light stability, retaining device performance under extended UV irradiation (Figure 2d–f).
Overall, these studies illustrate the versatility of polymer-based ETLs. Whether employed as pure semiconducting transport layers or integrated into hybrid systems with inorganic oxides, conjugated polymers offer a tunable platform for optimizing charge extraction, interface quality, and long-term stability. The compatibility of these materials with solution processing and low-temperature fabrication further supports their integration into flexible and scalable perovskite photovoltaic technologies.

4. Polymers as HTLs for PSCs

In the device structure of PSCs, polymeric HTLs are critical components that facilitate high efficiency and ensure long-term stability. Such excellent film-forming properties, suitable energy levels, and chemical resilience render them particularly attractive for both laboratory-scale and scalable device fabrication. Most notably, the inherent hydrophobicity of polymers assists in protecting the perovskite absorber from moisture, while their tunable electronic structure enables precise energy level alignment with the perovskite valence band, thereby minimizing interfacial recombination losses (Figure 3a). Among various candidates, poly[bis(4-phenyl)(2,4,6-trimethylphenyl)amine] (PTAA) has been identified as a benchmark polymer HTL, especially in inverted (p–i–n) device configurations. PTAA offers a deep highest occupied molecular orbital (HOMO) level that facilitates efficient hole collection and forms uniform, pinhole-free films that ensure intimate contact with the perovskite layer. In contrast to small-molecule HTLs such as Spiro-OMeTAD, PTAA exhibits higher conductivity without dopants, which enables its application not only in n-i-p structured PSCs but also in p-i-n structured PSCs [10,44,45,46]. As a result, PTAA-based PSCs exhibit superior operational stability, even under environmental stress, by preventing ion migration and moisture adsorption.
Although PTAA has been widely adopted, its relatively poor wettability and limited interfacial adhesion present persistent challenges, particularly on hydrophilic substrates or in large-area printing processes. Recent studies were aimed at addressing these limitations by chemically modifying PTAA or introducing functional additives. One such strategy involves the incorporation of fluorinated side chains into the PTAA backbone. This modification not only enhances hydrophobicity but also introduces interfacial dipole interactions that lower energetic disorder, improve hole mobility, and suppress trap-assisted recombination. As illustrated in Figure 3b,c, devices with fluorinated PTAA derivatives have consistently outperformed their unmodified counterparts in Voc and PCE metrics [47]. An alternative approach involves blending PTAA with insulating polymers such as poly(methyl methacrylate) (PMMA), which improves surface coverage and wetting behavior without compromising charge transport. The resulting composite HTLs exhibit improved film homogeneity and interface contact, which contribute to the stabilization of the perovskite morphology and the suppression of interfacial defects [48].
Figure 3. (a) Energy level diagram for perovskite (MAPbI3) and representative polymeric HTLs. (b) Energy diagrams of perovskite devices with PTAA derivatives as HTLs and (c) statistical data of PCE and Voc were obtained from the devices for each different HTL. Reprinted with permission from Ref. [47] Copyright 2018, John Wiley and Sons. (d) Molecular structure and density functional theory (DFT) calculation results of the stacking model of S-Poly-TPD, and (e) J–V characteristics and (f) maximum power point (MPP) tracking properties of PSCs. Adapted with permission from Ref. [49] Copyright 2022, John Wiley and Sons.
Figure 3. (a) Energy level diagram for perovskite (MAPbI3) and representative polymeric HTLs. (b) Energy diagrams of perovskite devices with PTAA derivatives as HTLs and (c) statistical data of PCE and Voc were obtained from the devices for each different HTL. Reprinted with permission from Ref. [47] Copyright 2018, John Wiley and Sons. (d) Molecular structure and density functional theory (DFT) calculation results of the stacking model of S-Poly-TPD, and (e) J–V characteristics and (f) maximum power point (MPP) tracking properties of PSCs. Adapted with permission from Ref. [49] Copyright 2022, John Wiley and Sons.
Polymers 17 02607 g003
In addition to modifications to PTAA, new classes of triphenylamine-based polymers have been designed to offer both electronic compatibility and mechanical robustness. For instance, the application of side-chain engineering has proven effective in enhancing polymer packing and interfacial interactions. A notable example is the side-chain-modified triphenylamine polymer (S-Poly-TPD) developed by Xie et al., which replaces linear alkyl chains with branched isobutyl groups [49]. This ostensibly subtle change improves molecular stacking, enhances hole mobility, and reduces nonradiative losses at the perovskite/HTL interface. Devices employing S-Poly-TPD achieved a certified PCE of 21.3% and maintained over 90% of initial performance after prolonged operation under illumination (Figure 3d–f).
These diverse developments reflect a broader trend toward rational molecular engineering of polymeric HTLs, with the aim of addressing the key limitations of scalability, interfacial compatibility, and long-term operational stability. The evolution of polymer HTLs, through means such as fluorination, side-chain engineering, and anchoring strategies, is of critical importance in the context of next-generation PSC technologies.

5. Polymers as Interfacial Layers for PSCs

In addition to their established roles as charge transport layers, polymers serve a variety of interfacial functions in PSCs, such as compatibilization, surface passivation, and morphological control. The presence of interfacial polymer layers has demonstrably regulated film formation, aligned energy levels, and mitigated nonradiative losses at critical junctions. By modifying surface energy and eliciting chemical or physical interactions with the perovskite layer, polymers can suppress trap-assisted recombination, facilitate crystallization, and enhance long-term device stability (Figure 4a). A notable example of interfacial compatibilization was demonstrated using amphiphilic conjugated polyelectrolytes. The introduction of ultrathin poly[(9,9-bis(3′-(N,N-dimethylamino)propyl)-2,7-fluorene)-alt-2,7-(9,9-dioctylfluorene)] (PFN) layers on top of hydrophobic organic transport layers effectively addressed the issue of wettability mismatch with aqueous perovskite precursors, enabling uniform and pinhole-free film formation across large areas. The PFN-modified interface preserved charge transport while promoting better perovskite coverage and stability. Devices incorporating this strategy achieved PCEs exceeding 19% and retained performance under ambient conditions, as shown in Figure 4b [50].
Interface passivation using ultrathin polymer coatings has also been explored to minimize recombination losses. In particular, the combination of PFN at the hole interface with a thin LiF layer at the electron contact suppressed non-radiative recombination at both interfaces, thereby increasing the quasi-Fermi level splitting. This dual-passivation approach has been demonstrated to enable reproducible 1 cm2 devices with certified efficiencies up to 19.83%, while maintaining uniformity and stability over large areas [20]. The inherent trade-off between passivation and charge transport has also been addressed through nanostructured architectures. The patterning of TiO2 nanorods and the subsequent selective coating with an insulating polymer blend (PMMA:PCBM) resulted in the maintenance of charge-selective contact, whilst the remaining surface was passivated. This design allowed for localized extraction and extensive passivation simultaneously, resulting in an outstanding fill factor of 83.9% and a certified PCE of 21.6% [51].
From a fabrication perspective, scalable polymer deposition techniques have shown promise. The plasma polymerization of adamantane-based polymers yielded ultrathin, uniform coatings when applied directly to TiO2 without the use of solvents. The nanometer-scale interlayers improved perovskite coverage and reduced interfacial resistance, resulting in champion efficiencies over 19% and prolonged stability under ambient exposure for over 1000 h without encapsulation [52]. Direct surface passivation using insulating polymers such as PMMA has been particularly effective across various device configurations. The spin-coating of PMMA layers on CsPbI2Br perovskites has been shown to be an effective method of suppressing defect-assisted recombination and enhancing Voc [53]. Furthermore, the deployment of ultrathin PMMA films at both electron- and hole-selective interfaces consistently improved interfacial quality and device stability, as evidenced in wide band-gap PSCs and MAPbI3-based systems [54,55]. Bilateral application of PMMA continued the significant suppression of recombination losses at both contacts, pushing Voc to 1.22 V and yielding a PCE of 20.8%. The carbonyl groups in PMMA were key to this process, coordinating with Pb2+ ions to neutralize trap states (Figure 4c,d) [56].
Notably, both physical and chemical passivation mechanisms have emerged, and their efficacy depends strongly on the functional groups introduced at the perovskite surface. It has been demonstrated that the deposition of extremely dilute PMMA solutions via the spin-coating method results in the preferential accumulation of the solutions in grain boundaries and surface depressions. This phenomenon effectively isolates high-defect regions without forming continuous barriers. This approach resulted in enhanced fill factor and absolute PCE gains that surpassed 1% [57]. Beyond this physical isolation, carbonyl-bearing polymers (e.g., PMMA and related methacrylate/amide derivatives) provide Lewis-base carbonyl oxygen sites that can weakly coordinate undercoordinated Pb2⁺ and reduce surface trap density, complementing the geometrical isolation. Other Lewis base polymers, such as poly(4-vinylpyridine) (PVP), have shown analogous advantageous effects. Through the coordination of pyridyl nitrogen with undercoordinated Pb2+ ions, PVP treatment increased device performances as shown in Figure 4e [58]. Additionally, this PVP enhanced carrier lifetime and suppressed nonradiative pathways, delivering a champion Voc of 1.16 V and enhanced operational durability over 90 days [58,59]. In parallel, polymers such as PCDTBT have been applied as interfacial modifications, contributing to long-term stability and efficient charge extraction when deposited directly on the perovskite surface [60].
Figure 4. (a) Overview of polymer functions at PSC interfaces—passivation, crystallization, protection. (b) Schematic illustrations of perovskite film formation on organic HTLs with and without the interfacial compatibilizer, as well as corresponding photographs and SEM images. Reprinted with permission from Ref. [50] Copyright 2017, John Wiley and Sons. (c) Schematic of the device structure and (d) corresponding SEM cross-sectional image. Reprinted with permission from Ref. [56]. Copyright 2018, John Wiley and Sons. (e) PCE distributions of 24 devices for the devices with and without PVP modification. Adapted with permission from Ref. [58] Copyright 2017, John Wiley and Sons.
Figure 4. (a) Overview of polymer functions at PSC interfaces—passivation, crystallization, protection. (b) Schematic illustrations of perovskite film formation on organic HTLs with and without the interfacial compatibilizer, as well as corresponding photographs and SEM images. Reprinted with permission from Ref. [50] Copyright 2017, John Wiley and Sons. (c) Schematic of the device structure and (d) corresponding SEM cross-sectional image. Reprinted with permission from Ref. [56]. Copyright 2018, John Wiley and Sons. (e) PCE distributions of 24 devices for the devices with and without PVP modification. Adapted with permission from Ref. [58] Copyright 2017, John Wiley and Sons.
Polymers 17 02607 g004
Solvent-free polymer coatings have also been realized using chemical vapor deposition polymerization of poly(p-xylylene), which offered a ~40 mV improvement in Voc and significantly enhanced moisture stability during long-term aging [61]. In a separate study, PVP coatings processed under ambient conditions were found to improve both film uniformity and resistance to moisture ingress. When applied as an ultrathin interfacial layer, PVP enabled PSCs to retain stable operation over 600 h and tolerate repeated mechanical deformation, demonstrating potential for flexible device platforms [62]. Passivation mechanisms can also be classified by their dominant functional chemistry as well as by physical or chemical characteristics. In particular, pyridyl groups (PVP-type) and carbonyl groups(ester/amide) show strong Lewis-base coordination to undercoordinated Pb2⁺ [58,59], while fluorinated side chains (–CF3, –CnF2n+1) provide hydrophobic surface passivation by lowering surface energy and suppressing moisture ingress; the latter can also introduce interfacial dipoles that assist band alignment when properly assembled. Complementary anchoring chemistries used in polymer or polymer-tethered layers (e.g., carboxylates and phosphonates for stronger binding, zwitterionic moieties for electrostatic screening) can be integrated into polymer backbones or side chains to simultaneously reduce defect density and improve environmental robustness. Hybrid strategies—such as zwitterionic interlayers, polymer blends, or SAM-tethered polymer brushes—thus combine electrostatic screening, dipole engineering, selective adhesion, and hydrophobic protection to stabilize large-area films under operational stress [57,58,59,60,61,62].
In summary, polymer-based interfacial layers—whether functioning via coordination bonding, surface energy tuning, or physical barrier formation—offer multifunctional, scalable solutions to reduce interfacial recombination and enhance device longevity. The tunability, ease of deposition, and compatibility with ambient or low-temperature processes of these materials make them essential design elements in both current and future PSC architectures.

6. Polymers as Transparent Electrodes for PSCs

Despite significant improvements in light absorption and charge transport layers in PSCs, the reliance on brittle and expensive transparent conductive oxides (TCOs) such as indium tin oxide (ITO) and costly noble metal electrodes like gold and silver remains a critical bottleneck to overcome for large-scale commercialization [63,64,65,66]. In recent years, conducting polymers have gained increasing attention as alternative electrodes for PSCs, providing a combination of flexibility, transparency, and processability through low-temperature solution techniques. When polymers are used as transparent electrodes in PSCs, as shown in Figure 5a, their performance is primarily assessed through three critical parameters: optical transmittance, sheet resistance, and work function alignment [67,68]. High optical transmittance (typically >85% in the visible range) ensures efficient photon delivery to the photoactive layer, while low sheet resistance (ideally <100 Ω/sq) is essential for efficient lateral charge transport, which is evaluated using a figure of merit to quantify the balance between optical transmittance and electrical conductivity [68].
Especially for the scale-up of PSCs, the lower the sheet resistance, the greater the minimization of energy loss during lateral charge transport. Notably, conducting polymers offer tunable work functions that facilitate effective energy level alignment with adjacent charge transport layers, minimizing interfacial losses [67]. Additionally, their flexibility and compatibility with low-temperature processing are key considerations, especially for flexible or roll-to-roll processed PSCs [68]. Recent advances in polymer composites, such as those blended with silver nanowires, carbon nanotubes, or graphene, have enabled simultaneous improvements in conductivity and transparency, approaching the performance benchmarks of traditional ITO electrodes while offering superior mechanical resilience, which are well reviewed in other papers and will not be discussed in this review [69,70,71]. Here, we mainly focus on only polymer-based electrodes for PSCs [67,72].
Conducting polymers such as PEDOT:PSS, PANI, polypyrrole (PPy), and newly developed n-type polymers such as poly(benzodifurandione) (n-PBDF) have demonstrated their potential to function as transparent electrodes for PSCs. In particular, PEDOT:PSS has been extensively studied since the development of OSCs, as its conductivity can be significantly enhanced through acid doping or solvent doping (Figure 5b) [73]. For instance, Zhu et al. demonstrated the feasibility of FAI-treated PEDOT:PSS as a transparent electrode for flexible PSCs, where FAI refers to formamidinium iodide [74]. The FAI-treated PEDOT:PSS electrode exhibited high optical transparency (over 85% transmittance from 350 nm to 550 nm and around 75% transmittance from 550 nm to 900 nm) and moderate electrical conductivity (1562 ± 36 S cm−1), enabling its application in both rigid and flexible PSC architectures. Devices fabricated on the polymer/glass substrate achieved a PCE of 13.36%, compared to 16.60% on conventional ITO/glass, while flexible devices using polymer-coated PET substrates reached a PCE of 10.16%. The reduced efficiency was attributed primarily to the lower conductivity of the polymer electrode and interfacial charge recombination. Despite these limitations, the work highlights a scalable, low-temperature processing route for transparent and flexible electrodes, suggesting future potential through conductivity optimization and interface engineering.
Moreover, by introducing fluorosurfactant to regulate the phase separation in PEDOT:PSS formulation, Hu et al. developed a mechanically robust PEDOT:PSS electrode to bridge the performance gap between flexible and rigid PSCs [75]. The resulting polymer network achieved exceptional electrical conductivity (>4000 S cm−1), high optical transmittance (>80% from 400 nm to 900 nm), and strong mechanical endurance. Flexible PSCs fabricated with this electrode exhibited high PCEs of 19.0% (0.1 cm2) and 10.9% (25 cm2), closely matching rigid device performance. Additionally, when employed as the top electrode in semi-transparent PSCs, the device achieved 12.5% PCE with an average visible transmittance of 30.6%. The PSCs demonstrated excellent mechanical stability, retaining over 80–90% of their initial efficiencies after 5000 bending cycles at a 3 mm bending radius, validating the electrode’s suitability for wearable and rollable photovoltaic applications.
Unlike p-type conducting polymers such as PEDOT:PSS, which are limited to inverted device structures, n-doped conducting polymers enable the fabrication of conventional planar PSC architectures. Kumar et al. reported the development of transparent conducting polymer electrodes based on n-doped conducting polymer (n-PBDF), offering a viable alternative to conventional ITO in PSCs (Figure 5c) [76]. The polymer electrodes with the optimal thickness range of 65–75 nm (90.9–78.8 Ohm sq−1) provide a suitable balance between optical transparency and electrical conductivity. Integration of n-PBDF into both rigid and flexible PSCs resulted in PCEs of 12.70% and 11.23%, respectively, with minimal impact on perovskite film quality, morphology, or optoelectronic properties. This work confirms the potential of n-type conducting polymers for developing ITO-free, flexible, and low-cost PSCs, addressing a critical bottleneck in transparent electrode technology.
Figure 5. (a) Schematic PSC device structure with transparent conducting polymer as a bottom electrode. (b) PEDOT:PSS nanofibrils (right) via a charge-separated transition mechanism (middle) via a concentrated H2SO4 treatment. Adapted with permission from Ref. [73] Copyright 2014, John Wiley and Sons. (c) Schematic overview of the preparation of n-type conducting polymer (n-PBDF) film. Adapted with permission from Ref. [76] Copyright 2025, Royal Society of Chemistry.
Figure 5. (a) Schematic PSC device structure with transparent conducting polymer as a bottom electrode. (b) PEDOT:PSS nanofibrils (right) via a charge-separated transition mechanism (middle) via a concentrated H2SO4 treatment. Adapted with permission from Ref. [73] Copyright 2014, John Wiley and Sons. (c) Schematic overview of the preparation of n-type conducting polymer (n-PBDF) film. Adapted with permission from Ref. [76] Copyright 2025, Royal Society of Chemistry.
Polymers 17 02607 g005
Despite recent progress, the widespread adoption of conducting polymers as transparent electrodes in PSCs still faces several critical challenges: further improving electrical conductivity and optical transparency by advanced doping strategies and nano-structuring approaches—such as incorporating metal grid, metal nanowires or carbon-based nanomaterials; ensuring mechanical robustness and environmental stability by developing cross-linked or self-healing polymer networks; the work function tunability of n-type and p-type conducting polymers to enable better energy-level alignment with transport layers in both conventional and inverted PSC architectures; scalable, low-temperature, and solvent-compatible processing techniques are essential to ensure compatibility with large-area, roll-to-roll manufacturing. Addressing these limitations could accelerate the commercial viability of polymer-based transparent electrodes in next-generation, flexible, and light weight photovoltaic technologies.

7. Polymers as Encapsulation Layers for PSCs

One of the most pressing challenges impeding the commercial deployment of PSCs is poor long-term environmental stability. Perovskite materials are highly sensitive to moisture, oxygen, heat, and ultraviolet (UV) light, which induce rapid degradation, particularly under ambient and outdoor conditions [77,78,79,80]. To address this, advanced encapsulation strategies are crucial, and polymeric materials have recently emerged as highly effective candidates due to their flexibility, processability, and tunable barrier properties [81,82,83,84,85,86,87]. The fundamental role of a polymer encapsulant is to isolate the photoactive layers from environmental stressors such as moisture, oxygen, dust, and UV radiation. Thus, effective polymer encapsulants must form a conformal and pinhole-free film that prevents the diffusion of ambient air into the device while maintaining optical transparency and mechanical compliance. Moreover, certain polymers, such as fluoropolymers, polyurethane (PU), and poly(methyl methacrylate) (PMMA), can be engineered to exhibit low water vapor transmission rate and low oxygen transmission rate (OTR), ideally below 10−5 g·m−2·day−1 and 10−3 cm3·m−2·day−1·atm−1, respectively, which makes them suitable for both rigid and flexible PSC architectures. Advanced polymer formulations may also incorporate self-healing, cross-linking, or lead-sequestration functionalities, further enhancing long-term durability and environmental safety. Therefore, the strategic integration of polymer encapsulants plays a critical role in advancing PSCs toward real-world stability and commercialization.
Among the latest advancements, Zagorovskaia et al. demonstrated the effectiveness of poly-para-xylylene (parylene N) as a robust encapsulation material, providing hydrophobic characteristics and excellent barrier performance (Figure 6a) [88]. PSCs encapsulated with parylene N and covered glass exhibited impressive operational stability exceeding 3800 h in dark ambient conditions, retaining 92% of their initial PCE. Furthermore, the polymer’s compatibility with low-temperature deposition processes makes it suitable for flexible substrates and tandem devices. Interestingly, Raman et al. reported a simple and scalable bilayer polymer encapsulation strategy to enhance the environmental stability of PSCs [89]. By combining a hydrophilic PMMA layer with a hydrophobic PU layer, the authors developed a protective barrier that effectively blocks moisture and oxygen infiltration without damaging the device structure (Figure 6b). Under high humidity conditions (80 ± 5% RH), the encapsulated PSCs showed no degradation for over 28 days and retained more than 92% of their initial PCE after 1500 h. In contrast, unencapsulated devices degraded rapidly within 48 h. The approach demonstrates the strong potential of polymer encapsulants for application in flexible and scalable optoelectronic devices.
Yang et al. investigated the role of hermetic encapsulation in enhancing the operational stability of PSCs, as shown in Figure 6c [90]. Unsealed or poorly sealed devices exhibited rapid degradation under continuous 1-sun illumination, showing a PCE drop of 20% within 300 h, accompanied by morphological changes such as the formation of coral-like crystal structures due to a decomposition–nucleation–regrowth process. In contrast, PSCs encapsulated with pressure-applied barrier films retained more than 90% of their initial PCE after 1000 h of operation. The hermetically sealed devices also suppressed ion migration and PbI2 reformation, key factors driving long-term degradation. This study emphasizes that encapsulation is not merely a protective covering but a functional layer essential for preserving both the morphological integrity and optoelectronic performance of PSCs, particularly under heat, light, and moisture stress. Unlike traditional encapsulants, the fluorosilicone gel offers simplified processing, effective thermal dissipation, and excellent suppression of lead leakage. Li et al. developed a room-temperature, nondestructive encapsulation method using a self-crosslinked fluorosilicone polymer gel to enhance the environmental and thermal stability of PSCs (Figure 6d) [91]. Devices encapsulated using this method retained 98% of their PCE after 1000 h under damp heat (85 °C/85% RH) and 95% after 220 thermal cycles, fully meeting the IEC 61,215 stability criteria. Additionally, the encapsulated cells achieved > 98% lead containment efficiency in immersion and rain exposure tests. This encapsulation strategy provides a universal and scalable path toward stable, sustainable, and environmentally safe perovskite photovoltaics. Table 1 summarizes the details of the encapsulated device performance.
Taken together, these studies reflect a significant shift toward multifunctional polymer encapsulation technologies that serve not only as protective barriers but also as performance enhancers. The future of PSC encapsulation will likely involve hybrid strategies combining self-healing, UV filtering, moisture barrier, and flexible properties into a single polymeric system—paving the way for efficient, durable, and scalable perovskite solar modules.
Table 1. The summary of the encapsulation strategies and the encapsulated device performance.
Table 1. The summary of the encapsulation strategies and the encapsulated device performance.
Device StructureEncapsulation MaterialsStability Test ConditionPerformanceRef.
Glass/ITO/SnO2/PCBA/MAPbI3/PTAA/VOx/AgParylene-N + cover glassDark ambient (22–26 °C, 30–40% RH)PCE retained 92% of initial value after 3800 h[88]
Glass/ITO/SnOx/FA0.9Cs0.1PbI3/spiro-OMeTAD/AgPMMA/PU bilayer Dark ambient (25 ± 3 °C, RH 80 ± 5%)PCE retained 92% after 1500 h[89]
FTO/SnO2/Cs0.05(FA0.85MA0.15)0.95Pb(Br0.15I0.85)3/Spiro-OMeTAD/Auhermetic (Surlyn + glass) Ambient air (25 °C, RH 55%), dark storagePCE retained 95.3% after 1036 h[90]
Glass/ITO/NiOx/Perovskite/PCBM/C60/BCP/Au+CrCFDP + cover glassThermal cycling (−40 °C to 85 °C, 220 cycles)Retained 95% of initial PCE[91]
Glass/FTO/c-TiO2/mp-TiO2/Cs0.10FA0.90Pb(I0.83Br0.17)3/HTL/AgLaser-assisted glass frit70 thermal cycles (−40 °C to 85 °C) + 50 h damp heat (85 °C, 85% RH)Retained 97% of initial PCE after 50 h[92]
Glass/ITO/NiOx/Cs0.17FA0.83Pb(I0.83Br0.17)3/PC60BM/SnO2/ZnSnO2/ITO/AgEthylene vinyl acetate (EVA) polymer or ionomer Surlyn 5400 + glassthermal cycling 200 cycles (−40 °C to 85 °C)Retained 90% of initial PCE[93]
Glass/ITO/NiO/Cs0.17FA0.83Pb(Br0.17I0.83)3/LiF/PCBM/SnO2/ZTO/ITO/AgEVA + glass, butyl rubber (edge seal)damp heat test: 85 °C, 85% RH, 1000 hPCE maintains over 90% after 1000 h[94]

8. Polymers as Anti-Reflection Layers for PSCs

Although the reported PCE of PSCs has surpassed 27%, minimizing optical and recombination losses remains essential to approach the fundamental theoretical limit of over 30% [95]. Thus, light management is a critical component in the design of high-efficiency PSCs [96,97,98]. Due to inherent optical losses from surface reflection, the integration of anti-reflection (AR) coatings has proven effective in improving light harvesting, particularly across the visible spectrum. Polymers have recently emerged as versatile AR materials due to their tunable refractive indices, solution processability, and mechanical flexibility, all of which are essential for next-generation solar technologies, including flexible and tandem architectures [99,100].
The principle of AR is fundamentally based on the destructive interference of reflected light waves at material interfaces. When a thin dielectric layer with an appropriate refractive index (n) is coated onto a substrate, partial reflections from the air/film and film/substrate boundaries interfere destructively if the optical path difference equals half the wavelength (λ/2). For maximum cancellation, the n of the AR layer is ideally chosen as the geometric mean of the refractive indices of air and the substrate ( n AR = n air   n substrate ), and the film thickness is typically a quarter of the target wavelength (λ/4n) [99,101]. This approach minimizes reflection and enhances transmission within the photoactive layers of optoelectronic devices, resulting in improved photon harvesting by reducing Fresnel losses at the air/glass or air/encapsulation interfaces, leading to enhanced short-circuit current density (Jsc) and thereby overall PCE.
Almuqoddas et al. theoretically simulated the effect of single-layer and double-layer AR coatings (SL-ARC and DL-ARC) on the performance of an inverted p–i–n PSC (Figure 7a) [102]. Initially, by optimizing the photoactive layer thickness, charge carrier mobility, and recombination rates, the maximum attainable efficiency under ideal internal conditions was predicted to be 28.4%. However, this value still falls short of the theoretical limit due to optical losses caused by surface reflection. To address this, SL-ARC and DL-ARC were introduced. Application of a 70 nm thick PMMA-based SL-ARC increased the PCE to 29.7%, while a dual-layer PMMA/polydimethylsiloxane (PDMS) DL-ARC further enhanced the PCE beyond 30%, effectively reaching the theoretical efficiency limit. These simulated results clearly show the necessity of AR coatings for high-performance PSCs.
Interestingly, Ma et al. developed a biomimetic AR film based on PDMS to enhance the optical management of PSCs [103]. Inspired by the hierarchical surface structures of plumeria flower petals, the authors replicated micro- and nano-patterns onto PDMS substrates using a soft-molding technique, as illustrated in Figure 7b. These films with specific patterns exhibited strong light-scattering capabilities, resulting in significantly reduced surface reflectance and broadband transmission enhancement in the visible range. Integration of the biomimetic AR film led to a PCE of 23.56%, with a notable Jsc of 26.54 mA cm−2, corresponding to an average efficiency improvement of 6.17%. The performance enhancement was validated under various lighting conditions, including AM 1.5G solar irradiation at different angles and light-emitting diode (LED) illumination of varying intensities. Optical simulations further confirmed the correlation between microstructure geometry and light-trapping efficiency. This study highlights the potential of nature-inspired polymer encapsulation films in advancing PSC performance through scalable, low-cost AR strategies.
In addition, Choi et al. proposed a novel anti-reflective film that was inspired by the nano-micro hierarchical structure of glasswing butterfly wings, fabricating a multilayer AR coating with gradient refractive index properties (Figure 7c) [104]. The resulting film minimized front-side optical reflection and simultaneously provided mechanical resilience compatible with flexible substrates. When applied to flexible PSCs, the AR film exhibited a substantial enhancement in light absorption and device stability under repetitive bending stress. These studies demonstrate a promising route toward high-efficiency, flexible PSCs via the integration of biomimetic AR functionality with structural robustness.
Polymer-based AR coatings represent a rapidly developing frontier in PSC technology. Their multifunctionality, structural tunability, and compatibility with low-cost fabrication methods make them promising candidates for both high-performance and scalable photovoltaic applications. Ongoing research is expected to further integrate these AR layers with encapsulation and passivation functionalities.

9. Polymers as Interconnecting Layers for Perovskite Tandem Cells

The use of perovskite/perovskite tandem solar cells has emerged as a promising strategy to surpass the theoretical efficiency limit of single-junction perovskite devices. By stacking wide-bandgap and narrow-bandgap perovskite absorbers, tandem structures enable a broader solar spectrum to be harvested. However, efficient charge recombination, energy level alignment, and interface passivation between the sub-cells are essential to achieving high device performance and long-term operational stability. Notably, polymer-based interconnecting layers (ICLs) have attracted significant attention as multifunctional interface engineering materials due to their solution processability, tunable optoelectronic properties, and mechanical flexibility [34,37]. Yang et al. introduced SnO2/ITO/PEDOT:PSS as ICL in all-perovskite tandem solar cells by showing a monolithic all-perovskite tandem cell PCE of 22.7% (Figure 8a) [105]. A crucial enabler of the tandem architecture was the use of a PEDOT:PSS-based ICL, which facilitated efficient hole recombination and preserved the optical and electrical integrity between the top and bottom sub-cells. The PEDOT:PSS ICL’s conformal coverage, transparency, and energetic alignment were key to minimizing series resistance and maximizing tandem performance.
Likewise, Abdollahi Nejand et al. successfully demonstrated scalable two-terminal all-perovskite tandem solar modules implementing an ICL architecture using SnOx/ITO or Au/PEDOT:PSS (Figure 8a), where PEDOT:PSS served as the hole-transport and recombination layer between the wide-bandgap top and narrow-bandgap bottom sub-cells [36]. The resulting tandem module achieved an aperture-area PCE of 19.1% over a large area of 12.25 cm2 with a geometric fill factor of 94.7%. Additionally, the PEDOT:PSS ICL was compatible with laser scribing and did not compromise module uniformity, as confirmed by electroluminescence and laser-beam-induced current (LBIC) mapping. Lin et al. developed high-efficiency monolithic all-perovskite tandem solar cells by addressing the key instability issue in tin-based narrow-bandgap perovskites: the oxidation of Sn(II) to Sn(IV) [38]. They introduced an ICL architecture of ALD-SnO2/Au/PEDOT:PSS in all perovskite tandem solar cells, achieving a certified small-area PCE of 24.8% (0.049 cm2) and 22.1% for larger area (1.05 cm2) devices. The PEDOT:PSS layer in ICL served as an efficient hole recombination layer, facilitating energy level alignment, maintaining optical transparency, and allowing low-temperature processing compatible with the perovskite layers. Its inclusion was critical in ensuring minimal series resistance and in maintaining the structural and operational integrity of the tandem stack, which retained over 90% of its initial performance after 463 h under 1-sun continuous illumination at the maximum power point. Consequently, these studies elucidate the critical role of conductive polymers like PEDOT:PSS in achieving scalable, efficient, and low-cost all-perovskite tandem photovoltaics.
Xu et al. developed a metal-free ICL for monolithic perovskite/organic tandem solar cells, consisting of atomic layer-deposited tin oxide (SnO2) combined with diluted PEDOT:PSS (Figure 8b) [35]. The authors fully eliminated the need for parasitically absorbing metallic recombination layers typically used in tandem architectures. This configuration provided a highly transparent, chemically stable, and energetically favorable interface between the perovskite and organic sub-cells. Importantly, the all-metal-free tandem device achieved a comparable PCE to the metal-based counterpart while exhibiting significantly enhanced photostability under prolonged light exposure. Outdoor stability tests over 2 weeks of real sunlight exposure showed that the monolithic tandem device outperformed both single-junction perovskite and OSCs in terms of retained PCE and thermal/photo durability, highlighting the practical advantage of this metal-free interface strategy for scalable tandem photovoltaics.
For scalable production of PSCs, a roll-to-roll printable polymer interlayer system with uniform film formation and good interfacial adhesion needs to be developed. This technology paves the way for the commercial viability of large-area tandem modules. Overall, polymers in interconnecting layers play a pivotal role in enhancing the optoelectronic coupling and mechanical/chemical stability of perovskite/perovskite tandem cells. The next generation of tandem photovoltaics will surely depend on further development of multifunctional, scalable polymer interconnecting layers tailored for both interface and device-wide performance optimization.

10. Challenges and Considerations for Large-Area Module Integration

Integrating polymer functional layers into large-area PSC modules presents significant hurdles, primarily in controlling film morphology and ensuring long-term device stability. Transitioning from lab-scale spin-coating to scalable methods such as slot-die or blade coating is essential but difficult, as altered fluid dynamics and drying kinetics often induce morphological defects. These issues, including non-uniform thickness, pinholes, and the coffee-ring effect, create shunt pathways that critically degrade the device’s fill factor and efficiency. Key research strategies to mitigate these problems involve optimizing ink formulations through solvent engineering, precisely controlling processing parameters, and designing new polymers tailored for high-throughput deposition. Stability issues become even more pronounced at the module level. Large-area devices are prone to heat accumulation and hot spots that can thermally degrade the polymer layers. In flexible roll-to-roll devices, mechanical stress can cause cracking or delamination, leading to electrical failure. Furthermore, as the difficulty of achieving complete encapsulation of large modules increases, the risk of moisture and oxygen ingress rises, accelerating chemical degradation at sensitive internal interfaces. Therefore, resolving the interrelated challenges of morphology and stability during the scale-up process remains a key challenge for the commercialization of PSCs with polymer functional layers.

11. Computational Modeling in Polymer Design

Beyond traditional experimental synthesis and characterization, the integration of computational modeling is emerging as a powerful strategy to accelerate the discovery of novel polymer functional layers. Theoretical methods offer profound insights into the properties that govern device performance, enabling a more rational design approach. For instance, Density Functional Theory (DFT) calculations can accurately predict the frontier molecular orbital energies (HOMO/LUMO) of yet-to-be-synthesized polymer candidates, allowing for an initial screening based on the prerequisite energy level alignment with the perovskite active layer.
Furthermore, as highlighted in recent computational studies, theoretical models can effectively simulate charge transport phenomena and thermal behavior at complex interfaces, including those between polymers and perovskites [106]. Molecular Dynamics (MD) simulations complement these electronic structure calculations by predicting polymer chain packing, bulk morphology, and the dynamic interactions at the polymer–perovskite junction, which are critical for both charge mobility and long-term stability. The significance of this approach lies in its potential for high-throughput virtual screening, where vast libraries of candidate molecules can be computationally evaluated in silico. This allows researchers to prioritize the most promising candidates for synthesis, creating an efficient feedback loop between theory and experimentation and significantly accelerating the development of next-generation materials for PSCs.

12. Summary

In summary, we reviewed the applications of polymers as functional layers and analyzed the commercialization challenges of PSCs. As charge transport layers, such as ETLs and HTLs, various polymers have appropriate energy levels and mobility for both normal and inverted structured PSCs. In particular, broadly utilized materials such as PTAA demonstrate performance comparable to that of SAMs or other small molecules that were originally developed for OLED applications before being adapted for use in other devices. In addition to the charge transport functional layers, polymers have been broadly applied as an interfacial layer for better wettability and the passivation of perovskite. They have also been applied as an encapsulation layer to prevent oxygen or moisture penetration. Conducting polymers such as PEDOT:PSS, PANI, and PPy have been adopted as electrodes for semitransparent and flexible PSCs. Thus far, polymers originally utilized in organic solar cells (OSCs), OLEDs, and DSSCs have been adapted and further optimized for integration into PSCs.
For more efficient and stable device operation, the advancement of novel dopant technologies or the development of new materials is likely required. In particular, under outdoor operating conditions, polymer functional layers in PSCs must withstand multiple stress factors, including UV exposure, thermal cycling, and ambient humidity for competitive outdoor performance. Polymer layers should maintain over 90% of their initial functionality after 1000 h of damp heat testing (85 °C/85% RH) and resist UV-induced degradation (e.g., under 1-sun AM 1.5G for 500–1000 h) without significant chemical or morphological changes. Glass transition temperatures (Tg) above 150 °C, UV absorption thresholds above 350 nm, and water vapor transmission rates (WVTR) < 10−5 g·m−2·day−1 are desirable targets for encapsulating or barrier-type polymer layers. Advances in crosslinked networks, fluorinated backbones, or UV-stabilized side chains help polymers close the stability gap with inorganics, and hybrid designs incorporating organic–inorganic nanocomposites may offer the best of both worlds. We hope this review will accelerate the development of polymer functional layers that can contribute to the commercialization of PSCs.

Author Contributions

Conceptualization, J.L., J.-H.L. and S.H.; investigation, J.L., J.K., J.-H.L. and S.H.; resources, J.L., J.-H.L. and S.H.; writing—original draft preparation, J.L., J.-H.L. and S.H.; writing—review and editing, J.L., J.-H.L. and S.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Korean government (MOTIE) under Grant No. 20223030010360. This work was also supported by a grant from the KRICT Core project (KS2522-30). This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (RS-2023-00278803), and this work was supported by the Technology development Program (RS-2025-02313936) funded by the Ministry of SMEs and Startups (MSS, Korea).

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
PSCPerovskite solar cells
PCEPower conversion efficiency
JscShort-circuit current density
VocOpen-circuit voltage
DSSCDye-sensitized solar cells
HTLHole transport layer
ETLElectron transport layer
Spiro-OMeTAD2,2′,7,7′-Tetrakis(N,N-di-p-methoxyphenylamine)-9,9′-spirobifluorene
PEDOT:PSSPoly(3,4-ethylenedioxythiophene):poly(styrenesulfonate)
PTAAPoly[bis(4-phenyl)(2,4,6-trimethylphenyl)amine]
OSCOrganic solar cells
PANIPolyaniline
PMMAPoly(methyl methacrylate)
ARAnti-reflection
PDMSPolydimethylsiloxane
LEDLight-emitting diode
ICLInterconnecting layer
ALDAtomic layer deposition
P3HTPoly(3-hexylthiophene)
PEIPolyethylenimine
PVPPoly(4-vinylpyridine)
PAAPolyallylamine
PUPolyurethane
n-PBDFPoly(benzodifurandione)

References

  1. Dastgeer, G.; Nisar, S.; Zulfiqar, M.W.; Eom, J.; Imran, M.; Akbar, K. A Review on Recent Progress and Challenges in High-Efficiency Perovskite Solar Cells. Nano Energy 2024, 132, 110401. [Google Scholar] [CrossRef]
  2. Seyisi, T.; Fouda-Mbanga, B.G.; Mnyango, J.I.; Nthwane, Y.B.; Nyoni, B.; Mhlanga, S.; Hlangothi, S.P.; Tywabi-Ngeva, Z. Major Challenges for Commercialization of Perovskite Solar Cells: A critical review. Energy Rep. 2025, 13, 1400–1415. [Google Scholar] [CrossRef]
  3. Kumar, N.S.; Naidu, K.C.B. A Review on Perovskite Solar Cells (PSCs), Materials and Applications. J. Mater. 2021, 7, 940–956. [Google Scholar]
  4. Green, M.A.; Ho-Baillie, A.; Snaith, H.J. The Emergence of Perovskite Solar Cells. Nat. Photon. 2014, 8, 506–514. [Google Scholar] [CrossRef]
  5. 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]
  6. Green, M.A.; Dunlop, E.D.; Yoshita, M.; Kopidakis, N.; Bothe, K.; Siefer, G.; Hao, X.; Jiang, J.Y. Solar Cell Efficiency Tables (Version 66). Prog. Photovolt. 2025, 33, 795–810. [Google Scholar] [CrossRef]
  7. Tang, X.; Yang, C.; Xu, Y.; Xia, J.; Li, B.; Li, M.; Zhou, Y.; Jiang, L.; Liu, H.; Ma, K.; et al. Enhancing the Efficiency and Stability of Perovskite Solar Cells via a Polymer Heterointerface Bridge. Nat. Photon. 2024, 19, 701–708. [Google Scholar] [CrossRef]
  8. Liu, X.; Zhang, J.; Wang, B.; Tong, G.; Yang, J.; Shi, Y.; Chen, Z.; Ono, L.K.; Jiang, Y.; Qi, Y. Perovskite Solar Modules with High Efficiency Exceeding 20%: From Laboratory to Industrial Community. Joule 2025, 9, 102056. [Google Scholar] [CrossRef]
  9. Kim, H.S.; Lee, C.R.; Im, J.-H.; Lee, K.-B.; Moehl, T.; Marchioro, A.; Moon, S.-J.; Humphry-Baker, R.; Yum, J.H.; Moser, J.E.; et al. Lead Iodide Perovskite Sensitized All-Solid-State Submicron Thin Film Mesoscopic Solar Cell with Efficiency Exceeding 9%. Sci. Rep. 2012, 2, 591. [Google Scholar] [CrossRef]
  10. Heo, J.H.; Im, S.H.; Noh, J.H.; Mandal, T.N.; Lim, C.-S.; Chang, J.A.; Lee, Y.H.; Kim, H.-j.; Sarkar, A.; Nazeeruddin, M.K.; et al. Efficient Inorganic–Organic Hybrid Heterojunction Solar Cells Containing Perovskite Compound and Polymeric Hole Conductors. Nat. Photon. 2013, 7, 486–491. [Google Scholar] [CrossRef]
  11. Jung, H.S.; Park, N.-G. Perovskite Solar Cells: From Materials to Devices. Small 2015, 11, 10–25. [Google Scholar] [CrossRef]
  12. Zhang, H.; Park, N.-G. Progress and Issues in P-I-N Type Perovskite Solar Cells. DeCarbon 2024, 3, 100025. [Google Scholar] [CrossRef]
  13. Le, T.-H.; Driscoll, H.; Hou, C.-H.; Montgomery, A.; Li, W.; Stein, J.S.; Nie, W. Perovskite Solar Module: Promise and Challenges in Efficiency, Meta-Stability, and Operational Lifetime. Adv. Electron. Mater. 2023, 9, 2300093. [Google Scholar] [CrossRef]
  14. Liu, Y.; Xiao, Y.; Jia, J.; Wang, H.; Yan, W.; Zhu, M. Perovskite Solar Cells: From Planar Designs to Fiber-Based Innovations. Wearable Electron. 2024, 1, 150–159. [Google Scholar] [CrossRef]
  15. Duan, L.; Liu, S.; Wang, X.; Zhang, Z.; Luo, J. Interfacial Crosslinking for Efficient and Stable Planar TiO2 Perovskite Solar Cells. Adv. Sci. 2024, 11, 2402796. [Google Scholar] [CrossRef]
  16. Yang, L.; Cappel, U.B.; Unger, E.L.; Karlsson, M.; Karlsson, K.M.; Gabrielsson, E.; Sun, L.; Boschloo, G.; Hagfeldt, A.; Johansson, E.M.J. Comparing Spiro-OMeTAD and P3HT Hole Conductors in Efficient Solid State Dye-Sensitized Solar Cells. Phys. Chem. Chem. Phys. 2012, 14, 779–789. [Google Scholar] [CrossRef]
  17. Marinova, N.; Valero, S.; Delgado, J.L. Organic and perovskite solar cells: Working Principles, Materials and Interfaces. J. Colloid Interface Sci. 2017, 488, 373–389. [Google Scholar] [CrossRef]
  18. Wang, Y.; Duan, L.; Zhang, M.; Hameiri, Z.; Liu, X.; Bai, Y.; Hao, X. PTAA as Efficient Hole Transport Materials in Perovskite Solar Cells: A Review. Sol. RRL 2022, 6, 2200234. [Google Scholar] [CrossRef]
  19. Yi, J.; Zhang, G.; Yu, H.; Yan, H. Advantages, Challenges and Molecular Design of Different Material Types Used in Organic Solar Cells. Nat. Rev. Mater. 2024, 9, 46–62. [Google Scholar] [CrossRef]
  20. Stolterfoht, M.; Wolff, C.M.; Márquez, J.A.; Zhang, S.; Hages, C.J.; Rothhardt, D.; Albrecht, S.; Burn, P.L.; Meredith, P.; Unold, T.; et al. Visualization and Suppression of Interfacial Recombination for High-Efficiency Large-Area pin Perovskite Solar Cells. Nat. Energy 2018, 3, 847–854. [Google Scholar] [CrossRef]
  21. Zhang, G.; Zheng, Y.; Wang, H.; Ding, G.; Yang, F.; Xu, Y.; Yu, J.; Shao, Y. Shellac Protects Perovskite Solar Cell Modules under Real-World Conditions. Joule 2024, 8, 496–508. [Google Scholar] [CrossRef]
  22. Weerasinghe, H.C.; Macadam, N.; Kim, J.-E.; Sutherland, L.J.; Angmo, D.; Ng, L.W.T.; Scully, A.D.; Glenn, F.; Chantler, R.; Chang, N.L.; et al. The First Demonstration of Entirely Roll-to-Roll Fabricated Perovskite Solar Cell Modules under Ambient Room Conditions. Nat. Commun. 2024, 15, 1656. [Google Scholar] [CrossRef]
  23. Kang, B.; Yan, F. Emerging Strategies for the Large-Scale Fabrication of Perovskite Solar Modules: From Design to Process. Energy Environ. Sci. 2025, 18, 3917–3954. [Google Scholar] [CrossRef]
  24. Dai, X.; Deng, Y.; Brackle, C.H.V.; Chen, S.; Rudd, P.N.; Xiao, X.; Lin, Y.; Chen, B.; Huang, J. Scalable Fabrication of Efficient Perovskite Solar Modules on Flexible Glass Substrates. Adv. Energy Mater. 2020, 10, 1903108. [Google Scholar] [CrossRef]
  25. Jung, H.S.; Han, G.S.; Park, N.-G.; Ko, M.J. Flexible Perovskite Solar Cells. Joule 2019, 3, 1850–1880. [Google Scholar] [CrossRef]
  26. Lee, D.-K.; Park, N.-G. Materials and Methods for High-Efficiency Perovskite Solar Modules. Sol. RRL 2022, 6, 2100455. [Google Scholar] [CrossRef]
  27. Pathak, C.S.; Choi, H.; Kim, H.; Lim, J.; Cho, S.-K.; Ham, D.S.; Song, S. Recent Progress in Coating Methods for Large-Area Perovskite Solar Module Fabrication. Sol. RRL 2024, 8, 2300860. [Google Scholar] [CrossRef]
  28. Isakova, A.; Topham, P.D. Polymer Strategies in Perovskite Solar Cells. J. Polym. Sci. B Polym. Phys. 2017, 55, 549–568. [Google Scholar] [CrossRef]
  29. Ma, Y.; Ge, J.; Jen, A.K.-Y.; You, J.; Liu, S. Polymer Boosts High Performance Perovskite Solar Cells: A Review. Adv. Opt. Mater. 2024, 12, 2301623. [Google Scholar] [CrossRef]
  30. Wang, S.; Gong, X.Y.; Li, M.-X.; Li, M.-H.; Hu, J.-S. Polymers for Perovskite Solar Cells. J. Am. Chem. Soc. Au. 2024, 4, 3400–3412. [Google Scholar] [CrossRef] [PubMed]
  31. Nam, J.-S.; Choi, J.-M.; Lee, J.W.; Han, J.; Jeon, I.; Kim, H.D. Decoding Polymeric Additive-Driven Self-Healing Processes in Perovskite Solar Cells from Chemical and Physical Bonding Perspectives. Adv. Energy Mater. 2024, 14, 2304062. [Google Scholar] [CrossRef]
  32. Kim, K.; Han, J.; Maruyama, S.; Balaban, M.; Jeon, I. Role and Contribution of Polymeric Additives in Perovskite Solar Cells: Crystal Growth Templates and Grain Boundary Passivators. Sol. RRL 2021, 5, 2000783. [Google Scholar] [CrossRef]
  33. Fan, X.; Nie, W.; Tsai, H.; Wang, N.; Huang, H.; Cheng, Y.; Wen, R.; Ma, L.; Yan, F.; Xia, Y. PEDOT:PSS for Flexible and Stretchable Electronics: Modifications, Strategies, and Applications. Adv. Sci. 2019, 6, 190081376. [Google Scholar] [CrossRef] [PubMed]
  34. Lim, J.; Park, N.-G.; Seok, S.I.; Saliba, M. All-Perovskite Tandem Solar Cells: From Fundamentals to Technological Progress. Energy Environ. Sci. 2024, 17, 4390–4425. [Google Scholar] [CrossRef] [PubMed]
  35. Xu, H.; Merino, L.T.; Koc, M.; Aydin, E.; Zhumagali, S.; Haque, M.A.; Yazmaciyan, A.; Sharma, A.; Villalva, D.R.; Hernandez, L.H.; et al. Metal-Free Interconnecting Layer for Monolithic Perovskite/Organic Tandem Solar Cells with Enhanced Outdoor Stability. ACS Appl. Energy Mater. 2022, 5, 14035–14044. [Google Scholar] [CrossRef]
  36. Nejand, B.A.; Ritzer, D.B.; Hu, H.; Schackmar, F.; Moghadamzadeh, S.; Feeney, T.; Singh, R.; Laufer, F.; Schmager, R.; Azmi, R.; et al. Scalable Two-Terminal All-Perovskite Tandem Solar Modules with a 19.1% Efficiency. Nat. Energy 2022, 7, 620–630. [Google Scholar] [CrossRef]
  37. Li, C.; Wang, Y.; Choy, W.C.H. Efficient Interconnection in Perovskite Tandem Solar Cells. Small Methods 2020, 4, 2000093. [Google Scholar] [CrossRef]
  38. Lin, R.; Xiao, K.; Qin, Z.; Han, Q.; Zhang, C.; Wei, M.; Saidaminov, M.I.; Gao, Y.; Xu, J.; Xiao, M.; et al. Monolithic All-Perovskite Tandem Solar Cells with 24.8% Efficiency Exploiting Comproportionation to Suppress Sn(ii) Oxidation in Precursor Ink. Nat. Energy 2019, 4, 864–873. [Google Scholar] [CrossRef]
  39. Chen, W.; Shi, Y.; Wang, Y.; Feng, X.; Djurišić, A.B.; Woo, H.Y.; Guo, X.; He, Z. N-Type Conjugated Polymer as Efficient Electron Transport Layer for Planar Inverted Perovskite Solar Cells with Power Conversion Efficiency of 20.86%. Nano Energy 2020, 68, 104363. [Google Scholar] [CrossRef]
  40. Deng, C.; Wan, L.; Li, S.; Tao, L.; Wang, S.-n.; Zhang, W.; Fang, J.; Fu, Z.; Song, W. Naphthalene Diimide Based Polymer as Electron Transport Layer in Inverted Perovskite Solar Cells. Org. Electron. 2020, 87, 105959. [Google Scholar] [CrossRef]
  41. Li, D.; Sun, C.; Li, H.; Shi, H.; Shai, X.; Sun, Q.; Han, J.; Shen, Y.; Yip, H.-L.; Huang, F.; et al. Amino-Functionalized Conjugated Polymer Electron Transport Layers Enhance the UV-Photostability of Planar Heterojunction Perovskite Solar Cells. Chem. Sci. 2017, 8, 4587–4594. [Google Scholar] [CrossRef]
  42. Xu, Z.; Ng, C.H.; Zhou, X.; Li, X.; Zhang, P.; Teo, S.H. Polymer-Complexed SnO2 Electron Transport Layer for High-Efficiency n-i-p Perovskite Solar Cells. Nanoscale 2022, 14, 12090–12098. [Google Scholar] [CrossRef] [PubMed]
  43. Sun, C.; Wu, Z.; Yip, H.-L.; Zhang, H.; Jiang, X.-F.; Xue, Q.; Hu, Z.; Hu, Z.; Shen, Y.; Wang, M.; et al. Amino-Functionalized Conjugated Polymer as an Efficient Electron Transport Layer for High-Performance Planar-Heterojunction Perovskite Solar Cells. Adv. Energy Mater. 2015, 6, 1501534. [Google Scholar] [CrossRef]
  44. Jeon, N.J.; Noh, J.H.; Kim, Y.C.; Yang, W.S.; Ryu, S.; Seok, S.I. Solvent Engineering for High-Performance Inorganic-Organic Hybrid Perovskite Solar Cells. Nat. Mater. 2014, 13, 897–903. [Google Scholar] [CrossRef]
  45. Ryu, S.; Noh, J.H.; Jeon, N.J.; Kim, Y.C.; Yang, W.S.; Seo, J.; Seok, S.I. Voltage Output of Efficient Perovskite Solar Cells with High Open-Circuit Voltage and Fill Factor. Energy Environ. Sci. 2014, 7, 2614–2618. [Google Scholar] [CrossRef]
  46. Zhang, C.; Yu, Z.; Li, B.; Li, X.; Gao, D.; Wu, X.; Zhu, Z. Exploring the Potential and Hurdles of Perovskite Solar Cells with p-i-n Structure. ACS Nano 2024, 18, 32299–32314. [Google Scholar] [CrossRef]
  47. Kim, Y.; Jung, E.H.; Kim, G.; Kim, D.; Kim, B.J.; Seo, J. Sequentially Fluorinated PTAA Polymers for Enhancing VOC of High-Performance Perovskite Solar Cells. Adv. Energy Mater. 2018, 8, 1801668. [Google Scholar] [CrossRef]
  48. Wang, Z.; Fan, P.; Zhang, D.; Yang, G.; Yu, J. Enhanced Efficiency and Stability of p-i-n Perovskite Solar Cells Using PMMA Doped PTAA as Hole Transport Layers. Synth. Met. 2020, 265, 116428. [Google Scholar] [CrossRef]
  49. Xie, Y.-M.; Yao, Q.; Xue, Q.; Zeng, Z.; Niu, T.; Zhou, Y.; Zhuo, M.-P.; Tsang, S.-W.; Yip, H.-L.; Cao, Y. Subtle Side Chain Modification of Triphenylamine-Based Polymer Hole-Transport Layer Materials Produces Efficient and Stable Inverted Perovskite Solar Cells. Interdiscip. Mater. 2022, 1, 281–293. [Google Scholar] [CrossRef]
  50. Lee, J.; Kang, H.; Kim, G.; Back, H.; Kim, J.; Hong, S.; Park, B.; Lee, E.; Lee, K. Achieving Large-Area Planar Perovskite Solar Cells by Introducing an Interfacial Compatibilizer. Adv. Mater. 2017, 29, 1606363. [Google Scholar] [CrossRef] [PubMed]
  51. Peng, J.; Walter, D.; Ren, Y.; Tebyetekerwa, M.; Wu, Y.; Duong, T.; Lin, Q.; Li, J.; Lu, T.; Mahmud, M.A.; et al. Nanoscale Localized Contacts for High Fill Factors in Polymer-Passivated Perovskite Solar Cells. Science 2021, 371, 390–395. [Google Scholar] [CrossRef] [PubMed]
  52. Obrero-Perez, J.M.; Contreras-Bernal, L.; Nuñez-Galvez, F.; Castillo-Seoane, J.; Valadez-Villalobos, K.; Aparicio, F.J.; Anta, J.A.; Borras, A.; Sanchez-Valencia, J.R.; Barranco, A. Ultrathin Plasma Polymer Passivation of Perovskite Solar Cells for Improved Stability and Reproducibility. Adv. Energy Mater. 2022, 12, 2200812. [Google Scholar] [CrossRef]
  53. Yuan, B.; Li, C.; Yi, W.; Juan, F.; Yu, H.; Xu, F.; Li, C.; Cao, B. PMMA Passivated CsPbI2Br Perovskite Film for Highly Efficient and Stable Solar Cells. J. Phys. Chem. Solids 2021, 153, 110000. [Google Scholar] [CrossRef]
  54. Ferdowsi, P.; Ochoa-Martinez, E.; Alonso, S.S.; Steiner, U.; Saliba, M. Ultrathin Polymeric Films for Interfacial Passivation in Wide Band-Gap Perovskite Solar Cells. Sci. Rep. 2020, 10, 22260. [Google Scholar] [CrossRef]
  55. Kim, W.; Park, J.; Aggarwal, Y.; Sharma, S.; Choi, E.H.; Park, B. Highly Efficient and Stable Self-Powered Perovskite Photodiode by Cathode-Side Interfacial Passivation with Poly(Methyl Methacrylate). Nanomaterials 2023, 13, 619. [Google Scholar] [CrossRef]
  56. Peng, J.; Khan, J.I.; Liu, W.; Ugur, E.; Duong, T.; Wu, Y.; Shen, H.; Wang, K.; Dang, H.; Aydin, E.; et al. A Universal Double-Side Passivation for High Open-Circuit Voltage in Perovskite Solar Cells: Role of Carbonyl Groups in Poly(methyl methacrylate). Adv. Energy Mater. 2018, 8, 1801208. [Google Scholar] [CrossRef]
  57. Ochoa-Martinez, E.; Ochoa, M.; Ortuso, R.D.; Ferdowsi, P.; Carron, R.; Tiwari, A.N.; Steiner, U.; Saliba, M. Physical Passivation of Grain Boundaries and Defects in Perovskite Solar Cells by an Isolating Thin Polymer. ACS Energy Lett. 2021, 6, 2626–2634. [Google Scholar] [CrossRef]
  58. Chaudhary, B.; Kulkarni, A.; Jena, A.K.; Ikegami, M.; Udagawa, Y.; Kunugita, H.; Ema, K.; Miyasaka, T. Poly(4-Vinylpyridine)-Based Interfacial Passivation to Enhance Voltage and Moisture Stability of Lead Halide Perovskite Solar Cells. ChemSusChem 2017, 10, 2473–2479. [Google Scholar] [CrossRef]
  59. Zuo, L.; Guo, H.; Dequilettes, D.W.; Jariwala, S.; Marco, N.D.; Dong, S.; DeBlock, R.; Ginger, D.S.; Dunn, B.; Wang, M.; et al. Polymer-Modified Halide Perovskite Films for Efficient and Stable Planar Heterojunction Solar Cells. Sci. Adv. 2017, 3, e1700106. [Google Scholar] [CrossRef]
  60. Zhang, C.-C.; Li, M.; Wang, Z.-K.; Jiang, Y.-R.; Liu, H.-R.; Yang, Y.-G.; Gao, X.-Y.; Ma, H. Passivated Perovskite Crystallization and Stability in Organic-Inorganic Halide Solar Cells by Doping a Donor Polymer. J. Mater. Chem. A 2017, 5, 2572–2579. [Google Scholar] [CrossRef]
  61. Byranvand, M.M.; Behboodi-Sadabad, F.; Alrhman, E.A.; Trouillet, V.; Welle, A.; Ternes, S.; Hossain, I.M.; Khan, M.R.; Schwenzer, J.A.; Farooq, A.; et al. Chemical Vapor Deposited Polymer Layer for Efficient Passivation of Planar Perovskite Solar Cells. J. Mater. Chem. A 2020, 8, 20122–20132. [Google Scholar] [CrossRef]
  62. Xiong, H.; DeLuca, G.; Rui, Y.; Zhang, B.; Li, Y.; Zhang, Q.; Wang, H.; Reichmanis, E. Modifying Perovskite Films with Polyvinylpyrrolidone for Ambient-Air-Stable Highly Bendable Solar Cells. ACS Appl. Mater. Interfaces 2018, 10, 35385–35394. [Google Scholar] [CrossRef]
  63. Xu, Y.; Lin, Z.; Wei, W.; Hao, Y.; Liu, S.; Ouyang, J.; Chang, J. Recent Progress of Electrode Materials for Flexible Perovskite Solar Cells. Nano-Micro Lett. 2022, 14, 117. [Google Scholar] [CrossRef]
  64. Guo, H.; Hou, F.; Ren, X.; Ning, X.; Wang, Y.; Yang, H.; Wei, J.; Guo, J.; Li, T.; Zhu, C.; et al. Recent Progresses on Transparent Electrodes and Active Layers Toward Neutral, Color Semitransparent Perovskite Solar Cells. Sol. RRL 2023, 7, 2300333. [Google Scholar] [CrossRef]
  65. Mujahid, M.; Chen, C.; Zhang, J.; Li, C.; Duan, Y. Recent advances in semitransparent perovskite solar cells. InfoMat 2020, 3, 101–124. [Google Scholar] [CrossRef]
  66. Wang, T.; Lu, K.; Xu, Z.; Lin, Z.; Ning, H.; Qiu, T.; Yang, Z.; Zheng, H.; Yao, R.; Peng, J. Recent Developments in Flexible Transparent Electrode. Crystals 2021, 11, 511. [Google Scholar] [CrossRef]
  67. Yelshibay, A.; Bukari, S.D.; Baptayev, B.; Balanay, M.P. Conducting Polymers in Solar Cells: Insights, Innovations, and Challenges. Organics 2024, 5, 640–669. [Google Scholar] [CrossRef]
  68. Kim, U.; Lee, J.; Lee, Y.S.; Choi, M. Mechanically Robust Transparent Conducting Electrodes for Flexible Perovskite Solar Cells. Korean J. Chem. Eng. 2024, 41, 3677–3692. [Google Scholar] [CrossRef]
  69. Chen, R.; Zhang, W.; Guan, X.; Raza, H.; Zhang, S.; Zhang, Y.; Troshin, P.A.; Kuklin, S.A.; Liu, Z.; Chen, W. Rear Electrode Materials for Perovskite Solar Cells. Adv. Funct. Mater. 2022, 32, 2200651. [Google Scholar] [CrossRef]
  70. Gan, Y.; Sun, J.; Guo, P.; Jiang, H.; Li, J.; Zhu, H.; Fan, X.; Huang, L.; Wang, Y. Advances in the Research of Carbon Electrodes for Perovskite Solar Cells. Dalton Trans. 2023, 52, 16558. [Google Scholar] [CrossRef] [PubMed]
  71. Chae, E.C.; Seo, Y.-H.; Kang, B.J.; Oh, J.H.; Jung, Y.; Jang, J.; Kim, T.; Jo, Y.-R.; Kim, D.J.; Kim, T.-S.; et al. PTAA-Infiltrated Thin-Walled Carbon Nanotube Electrode with Hidden Encapsulation for Perovskite Solar Cells. EcoMat 2024, 6, e12495. [Google Scholar] [CrossRef]
  72. Nguyen, V.H.; Papanastasiou, D.T.; Resende, J.; Bardet, L.; Sannicolo, T.; Jimenez, C.; Munoz-Rojas, D.; Nguyen, N.D.; Bellet, D. Advances in Flexible Metallic Transparent Electrodes. Small 2022, 18, e2106006. [Google Scholar] [CrossRef]
  73. Kim, N.; Kee, S.; Lee, S.H.; Lee, B.H.; Kahng, Y.H.; Jo, Y.-R.; Kim, B.-J.; Lee, K. Highly Conductive PEDOT:PSS Nanofibrils Induced by Solution-Processed Crystallization. Adv. Mater. 2014, 26, 2268–2272. [Google Scholar] [CrossRef]
  74. Zhu, T.; Yang, Y.; Yao, X.; Huang, Z.; Liu, L.; Hu, W.; Gong, X. Solution-Processed Polymeric Thin Film as the Transparent Electrode for Flexible Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2020, 12, 15456–15463. [Google Scholar] [CrossRef] [PubMed]
  75. Hu, X.; Meng, X.; Zhang, L.; Zhang, Y.; Cai, Z.; Huang, Z.; Su, M.; Wang, Y.; Li, M.; Li, F.; et al. A Mechanically Robust Conducting Polymer Network Electrode for Efficient Flexible Perovskite Solar Cells. Joule 2019, 3, 2205–2218. [Google Scholar] [CrossRef]
  76. Kumar, P.; Lee, W.-J.; Tang, Y.; Yang, H.; Xu, W.; Rout, P.; You, L.; Romo, J.A.; Pradhan, B.; Mei, J.; et al. Enabling ITO-Free Perovskite Solar Cells through N-Doped Poly(benzodifurandione) (n-PBDF) Electrodes. EES Sol. 2025, 1, 529–535. [Google Scholar] [CrossRef]
  77. Chowdhury, T.A.; Zafar, M.A.B.; Islam, M.S.-U.; Shahinuzzaman, M.; Islam, M.A.; Khandaker, M.U. Stability of Perovskite Solar Cells: Issues and Prospects. RSC Adv. 2023, 13, 1787–1810. [Google Scholar] [CrossRef]
  78. Kahandal, S.S.; Tupke, R.S.; Bobade, D.S.; Kim, H.; Piao, G.; Sankapal, B.R.; Said, Z.; Pagar, B.P.; Pawar, A.C.; Kim, J.M.; et al. Perovskite Solar Cells: Fundamental Aspects, Stability Challenges, and Future Prospects. Prog. Solid State Chem. 2024, 74, 100463. [Google Scholar] [CrossRef]
  79. Dipta, S.S.; Uddin, A. Stability Issues of Perovskite Solar Cells: A Critical Review. Energy Technol. 2021, 9, 2100560. [Google Scholar] [CrossRef]
  80. Wang, D.; Wright, M.; Elumalai, N.K.; Uddin, A. Stability of Perovskite Solar Cells. Sol. Energy Mater. Sol. Cells 2016, 147, 255–275. [Google Scholar] [CrossRef]
  81. Uddin, A.; Upama, M.B.; Yi, H.; Duan, L. Encapsulation of Organic and Perovskite Solar Cells: A Review. Coatings 2019, 9, 65. [Google Scholar] [CrossRef]
  82. Ma, S.; Yuan, G.; Zhang, Y.; Yang, N.; Li, Y.; Chen, Q. Development of Encapsulation Strategies Towards the Commercialization of Perovskite Solar Cells. Energy Environ. Sci. 2022, 15, 13–55. [Google Scholar] [CrossRef]
  83. Lu, Q.; Yang, Z.; Meng, X.; Yue, Y.; Ahmad, M.A.; Zhang, W.; Zhang, S.; Zhang, Y.; Liu, Z.; Chen, W. A Review on Encapsulation Technology from Organic Light Emitting Diodes to Organic and Perovskite Solar Cells. Adv. Funct. Mater. 2021, 31, 2100151. [Google Scholar] [CrossRef]
  84. Chu, Q.-Q.; Sun, Z.; Wang, D.; Cheng, B.; Wang, H.; Wong, C.-P.; Fang, B. Encapsulation: The Path to Commercialization of Stable Perovskite Solar Cells. Matter 2023, 6, 3838–3863. [Google Scholar] [CrossRef]
  85. Corsini, F.; Griffini, G. Recent Progress in Encapsulation Strategies to Enhance the Stability of Organometal Halide Perovskite Solar Cells. J. Phys. Energy 2020, 2, 31001. [Google Scholar] [CrossRef]
  86. Sutherland, L.J.; Weerasinghe, H.C.; Simon, G.P. A Review on Emerging Barrier Materials and Encapsulation Strategies for Flexible Perovskite and Organic Photovoltaics. Adv. Energy Mater. 2021, 11, 2101383. [Google Scholar] [CrossRef]
  87. Mu, L.; Wang, S.; Liu, H.; Li, W.; Zhu, L.; Wang, H.; Chen, H. Innovative Materials for Lamination Encapsulation in Perovskite Solar Cells. Adv. Funct. Mater. 2024, 35, 2415353. [Google Scholar] [CrossRef]
  88. Zagorovskaia, K.; Trepalin, O.; Krasionov, I.; Luchkin, S.; Krasnikov, D.V.; Nasibulin, A.G.; Boldyreva, A.G. Improved Operational Lifetime of MAPbI3 Solar Cells Encapsulated with Parylene-N. Sol. RRL 2025, 9, 2400833. [Google Scholar] [CrossRef]
  89. Raman, R.K.; Ganesan, S.; Alagumalai, A.; Menon, V.S.; Krishnan, S.; Thangavelu, S.A.G.; Krishnamoorthy, A. Facile and Scalable Bilayer Polymer Encapsulation to Achieve Long-Term Stability of Perovskite Solar Cells under Harsh Humidity Conditions. Sustain. Energ. Fuels 2024, 8, 1953–1965. [Google Scholar] [CrossRef]
  90. Yang, Y.; Song, X.; Liu, H.; Zhang, A.; Cao, J.; Wu, C. Encapsulation-Driven Stability in Perovskite Solar Cells: Suppressing Degradation through Hermetic Sealing. ACS Appl. Mater. Interfaces 2025, 17, 34119–34128. [Google Scholar] [CrossRef]
  91. Wang, T.; Yang, J.; Cao, Q.; Pu, X.; Li, Y.; Chen, H.; Zhao, J.; Zhang, Y.; Chen, X.; Li, X. Room Temperature Nondestructive Encapsulation via Self-Crosslinked Fluorosilicone Polymer Enables Damp Heat-Stable Sustainable Perovskite Solar Cells. Nat. Commun. 2023, 14, 1342. [Google Scholar] [CrossRef]
  92. Emami, S.; Martins, J.; Ivanou, D.; Mendes, A. Advanced Hermetic Encapsulation of Perovskite Solar Cells: The Route to Commercialization. J. Mater. Chem. A 2020, 8, 2654–2662. [Google Scholar] [CrossRef]
  93. Cheacharoen, R.; Rolston, N.; Harwood, D.; Bush, K.A.; Dauskardt, R.H.; McGehee, M.D. Design and Understanding of Encapsulated Perovskite Solar Cells to Withstand Temperature Cycling. Energy Environ. Sci. 2018, 11, 144–150. [Google Scholar] [CrossRef]
  94. Bush, K.A.; Palmstrom, A.F.; Yu, Z.J.; Boccard, M.; Cheacharoen, P.; Mailoa, J.P.; McMeekin, D.P.; Hoye, R.L.Z.; Bailie, C.D.; Leijtens, T.; et al. 23.6%-Efficient Monolithic Perovskite/Silicon Tandem Solar Cells with Improved Stability. Nat. Energy 2017, 2, 17009. [Google Scholar] [CrossRef]
  95. Ma, C.; Park, N.-G. A Realistic Methodology for 30% Efficient Perovskite Solar Cells. Chem 2020, 6, 1254–1264. [Google Scholar] [CrossRef]
  96. Chen, C.; Zheng, S.; Song, H. Photon Management to Reduce Energy Loss in Perovskite Solar Cells. Chem. Soc. Rev. 2021, 50, 7250–7329. [Google Scholar] [CrossRef]
  97. Jung, S.-K.; Park, N.-G.; Lee, J.-W. Light Management in Perovskite Solar Cells. Mater. Today Energy 2023, 37, 101401. [Google Scholar] [CrossRef]
  98. Lei, Y.; Li, Y.; Jin, Z. Photon Energy Loss and Management in Perovskite Solar Cells. Energy Rev. 2022, 1, 100003. [Google Scholar] [CrossRef]
  99. Ji, C.; Liu, W.; Bao, Y.; Chen, X.; Yang, G.; Wei, B.; Yang, F.; Wang, X. Recent Applications of Antireflection Coatings in Solar Cells. Photonics 2022, 9, 906. [Google Scholar] [CrossRef]
  100. Shanmugam, N.; Pugazhendhi, R.; Elavarasan, R.M.; Kasiviswanathan, P.; Das, N. Anti-Reflective Coating Materials: A Holistic Review from PV Perspective. Energies 2020, 13, 2631. [Google Scholar] [CrossRef]
  101. Raut, H.K.; Ganesh, V.A.; Nair, A.S.; Ramakrishna, S. Anti-Reflective Coatings: A Critical, In-Depth Review. Energy Environ. Sci. 2011, 4, 779–3804. [Google Scholar] [CrossRef]
  102. Almuqoddas, E.; Budiawan, W.; Paramudita, I.; Shobih; Yuliarto, B.; Firdaus, Y. Near Fundamental Limit Performance of Inverted Perovskite Solar Cells with Anti-Reflective Coating Integration. Results Opt. 2024, 16, 100670. [Google Scholar] [CrossRef]
  103. Ma, S.; Tao, L.; Zhang, H.; Qi, C.; Ye, T.; Zhu, X.; Li, B.; Yun, D.-Q.; Sun, X.; Zheng, L. Optical Management of Perovskite Solar Cells by Biomimetic-Structured Polydimethylsiloxane Films. Adv. Eng. Mater. 2024, 26, 2400065. [Google Scholar] [CrossRef]
  104. Choi, J.S.; Kim, U.; Lee, J.; Lee, Y.S.; Choi, M.; Kang, S.M. Bioinspired Sticker-Type Multilayer Anti-Reflective Film for Flexible Perovskite Solar Cells. J. Energy Chem. 2025, 107, 540–547. [Google Scholar] [CrossRef]
  105. Yang, Z.; Yu, Z.; Wei, H.; Xiao, X.; Ni, Z.; Chen, B.; Deng, Y.; Habisreutinger, S.N.; Chen, X.; Wang, K.; et al. Enhancing Electron Diffusion Length in Narrow-Bandgap Perovskites for Efficient Monolithic Perovskite Tandem Solar Cells. Nat. Commun. 2019, 10, 4498. [Google Scholar] [CrossRef]
  106. Koufi, A.; Ziat, Y.; Belkhanchi, H. First-Principles DFT and BoltzTraP Investigation of Multifunctional Properties of XNiH3 (X = Li, Na, K) Perovskite Hydrides: Thermoelectric and Hydrogen Storage Potential. Next Energy 2025, 9, 100402. [Google Scholar] [CrossRef]
Figure 1. A schematic of PSC structure that includes representative polymer-based functional layers such as ETLs, HTLs, interfacial layers, electrodes, encapsulation layers, anti-reflection layers.
Figure 1. A schematic of PSC structure that includes representative polymer-based functional layers such as ETLs, HTLs, interfacial layers, electrodes, encapsulation layers, anti-reflection layers.
Polymers 17 02607 g001
Figure 6. (a) Parylene-N encapsulation of MAPbI3 PSCs via CVD improves shelf-life stability without compromising device performance. Adapted with permission from Ref. [88] Copyright 2025, John Wiley and Sons. (b) Cross-sectional structure and long-term operational stability of FACsPbI3 PSCs with different encapsulation strategies. The bilayer PMMA/PU structure significantly retains both PCE and Voc over 1500 h under ambient storage (25 °C, RH ~50%, dark), compared to single-layer PMMA, PU, and unencapsulated devices. Adapted with permission from Ref. [89] Copyright 2024, Royal Society of Chemistry. (c) Comparison of hermetic and non-hermetic encapsulation on PSCs under heat and light stress, showing improved morphology retention, structural stability, and device performance with hermetic sealing. Adapted with permission from Ref. [90] Copyright 2025, American Chemical Society. (d) Schematic illustration of UV resin-encapsulated, and CFDP-based target-encapsulated PSCs, including the molecular structure of the CFDP composite, and their J–V characteristics and long-term stability under damp heat (85 °C, 85% RH) and thermal cycling (−40 °C to 85 °C) conditions. The molecular structure of a self-crosslinked fluorosilicone polymer gel is in Red dotted lines. Adapted with permission from Ref. [91] Copyright 2023, Springer Nature.
Figure 6. (a) Parylene-N encapsulation of MAPbI3 PSCs via CVD improves shelf-life stability without compromising device performance. Adapted with permission from Ref. [88] Copyright 2025, John Wiley and Sons. (b) Cross-sectional structure and long-term operational stability of FACsPbI3 PSCs with different encapsulation strategies. The bilayer PMMA/PU structure significantly retains both PCE and Voc over 1500 h under ambient storage (25 °C, RH ~50%, dark), compared to single-layer PMMA, PU, and unencapsulated devices. Adapted with permission from Ref. [89] Copyright 2024, Royal Society of Chemistry. (c) Comparison of hermetic and non-hermetic encapsulation on PSCs under heat and light stress, showing improved morphology retention, structural stability, and device performance with hermetic sealing. Adapted with permission from Ref. [90] Copyright 2025, American Chemical Society. (d) Schematic illustration of UV resin-encapsulated, and CFDP-based target-encapsulated PSCs, including the molecular structure of the CFDP composite, and their J–V characteristics and long-term stability under damp heat (85 °C, 85% RH) and thermal cycling (−40 °C to 85 °C) conditions. The molecular structure of a self-crosslinked fluorosilicone polymer gel is in Red dotted lines. Adapted with permission from Ref. [91] Copyright 2023, Springer Nature.
Polymers 17 02607 g006
Figure 7. (a) Device architecture of PSCs incorporating a PMMA/PDMS-based single-layer anti-reflective coating (SL-ARC and DL-ARC) and their chemical structures, along with the simulated J-V curves with SL-ARC (left) and DL-ARC (right). Adapted with permission from Ref. [102] Copyright 2024, Elsevier. (b) Fabrication process and surface morphology of biomimetic PDMS films, device architecture of PDMS-integrated PSCs, and corresponding photovoltaic parameters (Jsc, PCE) for different PDMS film types. Adapted with permission from Ref. [103] Copyright 2024, John Wiley and Sons. (c) Design and fabrication process of GSMA film with actual wing structure of the glasswing butterfly, showing improved photovoltaic performance, especially Jsc compared to the control device. Adapted with permission from Ref. [104]. Copyright 2025, Elsevier.
Figure 7. (a) Device architecture of PSCs incorporating a PMMA/PDMS-based single-layer anti-reflective coating (SL-ARC and DL-ARC) and their chemical structures, along with the simulated J-V curves with SL-ARC (left) and DL-ARC (right). Adapted with permission from Ref. [102] Copyright 2024, Elsevier. (b) Fabrication process and surface morphology of biomimetic PDMS films, device architecture of PDMS-integrated PSCs, and corresponding photovoltaic parameters (Jsc, PCE) for different PDMS film types. Adapted with permission from Ref. [103] Copyright 2024, John Wiley and Sons. (c) Design and fabrication process of GSMA film with actual wing structure of the glasswing butterfly, showing improved photovoltaic performance, especially Jsc compared to the control device. Adapted with permission from Ref. [104]. Copyright 2025, Elsevier.
Polymers 17 02607 g007
Figure 8. Schematic device architectures of perovskite tandem solar cells employing PEDOT:PSS-based interconnecting layers (ICL) (a) with and (b) without metal layers. Red boxes are recombination layers.
Figure 8. Schematic device architectures of perovskite tandem solar cells employing PEDOT:PSS-based interconnecting layers (ICL) (a) with and (b) without metal layers. Red boxes are recombination layers.
Polymers 17 02607 g008
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

Lee, J.; Kang, J.; Lee, J.-H.; Hong, S. Polymer Functional Layers for Perovskite Solar Cells. Polymers 2025, 17, 2607. https://doi.org/10.3390/polym17192607

AMA Style

Lee J, Kang J, Lee J-H, Hong S. Polymer Functional Layers for Perovskite Solar Cells. Polymers. 2025; 17(19):2607. https://doi.org/10.3390/polym17192607

Chicago/Turabian Style

Lee, Jinho, Jaehyeok Kang, Jong-Hoon Lee, and Soonil Hong. 2025. "Polymer Functional Layers for Perovskite Solar Cells" Polymers 17, no. 19: 2607. https://doi.org/10.3390/polym17192607

APA Style

Lee, J., Kang, J., Lee, J.-H., & Hong, S. (2025). Polymer Functional Layers for Perovskite Solar Cells. Polymers, 17(19), 2607. https://doi.org/10.3390/polym17192607

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