Open Access Article
Farzad Zahedi
a,
Mohsen Ameri
*b,
Mohammad Hossein Rajabi Manshadic,
Mohammad Jafari Pouyac,
Marziyeh Mohebbid,
Maryam Alidaei
e,
Siming Huangf,
Tandis HosseinMirzaiec,
Galyam Sanfof and
Mojtaba Abdi-Jalebi
*f
aPolymer Engineering Department, Amirkabir University of Technology (Tehran Polytechnic), Tehran, Iran
bFirst Solar, Inc., 28101 Cedar Park Blvd., Perrysburg, OH 43551, USA. E-mail: mohsen.ameri@firstsolar.com
cPhysics Department, Faculty of Science, University of Tehran, Iran
dDepartment of Chemistry, University of Isfahan, Isfahan, 81746-73441, Iran
eFaculty of Electrical and Computer Engineering, Tarbiat Modares University, Tehran, Iran
fInstitute for Materials Discovery, University College London, Malet Place, London WC1E 7JE, UK. E-mail: m.jalebi@ucl.ac.uk
First published on 21st November 2025
Polymers have emerged as multifunctional enablers in the evolving architecture of perovskite solar cells (PSCs), addressing key challenges in film formation, interface engineering, defect modification, and device longevity. Their molecular tunability, macromolecular chains, and compatibility with low-temperature solution processing make them ideal candidates for integration at various functional layers within PSCs. This review delineates the strategic roles of polymers as both passive and active components in PSCs and emphasizes their indispensable contribution toward scalable, printable perovskite photovoltaics. In the active layer, polymeric additives modulate the nucleation and crystal growth kinetics of the perovskite phase, leading to adjusted grain size, reduced trap states, and suppressed ion migration. At the interfaces, conjugated polymers serve as efficient charge-transport materials (CTMs), offering favorable energy level alignment and improved mechanical adhesion, while insulating polymers mitigate interface-induced recombination and phase instability. Additionally, polymers have been pivotal in enabling flexible and lightweight PSCs. This review provides a comprehensive overview of perovskite defects and systematically investigates the role of polymers in enhancing the properties of PSCs. We explore the incorporation of polymers into the active layer, charge transport layers (CTLs), and their interfaces, highlighting recent advancements in self-healing, deep-level trap passivation, network formation, the use of hyperbranched polymers for flexible devices, and so on. A detailed analysis of semiconducting polymers, focusing on main-chain and side-chain structures, physical properties, and dopant-free systems, is presented for both the active layer as the bulk heterointerface and CTLs. Finally, addressing the imminent industrialization of PSCs, we examine various printing techniques and propose polymer-based strategies to mitigate structural defects during large-scale fabrication, concluding with a perspective on future scalability.
In PSCs, a perovskite absorber layer is situated between an ETL and a HTL. These layers are designed to extract and direct photo-generated electrons and holes into the circuit. However, this process is hindered by defects within the perovskite structure and during fabrication. These defects lead to issues such as charge recombination and ion migration, which reduce the efficiency and stability of PSCs and make them susceptible to degradation from external factors like moisture, heat, and light.5 Therefore, managing the fabrication process and passivating these defects is essential for achieving high performance and stability.
The morphology of the perovskite layer, both in the bulk and on the surface, plays a critical role in photovoltaic performance.6 Researchers aim to manipulate the composition and interfaces to reduce recombination, enhance charge transport, extend charge carrier lifetime, and improve stability. For this purpose, two goals are running in parallel, one is the engineering of perovskite compounds and the other is the employment of suitable materials to passivate defects.7–9 Perovskite passivation includes a broad range of techniques with various agents from inorganic additives to polymers.
Among the materials, small molecules are beneficial but macromolecules are more efficient. Polymers can be engineered to have a wide range of properties, such as energy level alignment, charge mobility, and solubility, which are crucial for optimizing the performance of solar cells. They tend to form high-quality films with fewer defects, which is essential for producing uniform layers. Polymers can improve the stability of perovskite solar cells under environmental stressors like moisture, heat, and UV light, which is a significant advantage over small molecules. The structures of polymer–perovskite solar cells are of great significance due to their potential for high efficiency and stability.10–12 Fig. 1 depicts, in a categorized manner, some of the most important advantages and limitations of PSCs and the role of polymers in them.
![]() | ||
| Fig. 1 Benefits and challenges of perovskites and synergistic effects of polymers for use in perovskite solar cells. | ||
In this review, we begin by briefly outlining the various types of defects present in perovskite materials. We then explore how polymers can be strategically employed to enhance the performance of perovskite solar cells. To this end, we first touch upon the key factors of polymeric additives and then examine their integration into the active layer, CTLs, and their corresponding interfaces. This discussion highlights cutting-edge research, including self-healing mechanisms, passivation of deep-level traps, network formation, the application of hyperbranched polymers in flexible solar cells, and so on. Subsequently, we offer an in-depth analysis of semiconducting polymers (SCPs), investigating the influence of main-chain and side-chain engineering, physical and crystalline properties, dopant-free systems, and donor–acceptor polymers within the device architecture. Given the impending commercialization of this technology, we then address the challenges of scalable manufacturing by evaluating various printing methods. We identify key structural challenges associated with printing and present polymer-based solutions drawn from both foundational and recent studies. Finally, we conclude with a forward-looking perspective on the large-scale production of polymer-enhanced PSCs (see Fig. 2).
Imperfections typically induce shallow or deep energy levels within the band structure. Shallow-level defects, which can be identified by a redshift in the photoluminescence (PL) peak, cause energy loss due to charge capture.16 These defects are near the band edge and provide potential pathways for trapped electrons to recombine within the conduction or valence bands. On the other hand, deep-level defects capture charges that lead to non-radiative recombination, with most energy dissipated as phonons to the neighboring lattice. This severely impacts the performance of PSCs by reducing carrier density, diffusion length, and lifetime.17
From an energy level perspective, defects create trap states for charge carriers, with the depth of these trap states indicating the potential risk they pose. Shallow defects are less energetically activated and are close to the valence band maximum (VBM) or conduction band minimum (CBM), often allowing trapped carriers to escape.18 Conversely, deep defects are the most destructive, possessing high formation energy and energy levels distant from the VBM and CBM, leading to carrier recombination and reduced PSC efficiency (Fig. 4A). Some researchers introduced the term defect tolerance to explain the unusually long carrier diffusion length in perovskites compared to traditional semiconductors, attributing this characteristic to atomic orbital coupling within the material.19,20 Due to the antibonding nature of the molecular orbitals in perovskites, the material inherently exhibits more shallow defects, making it more defect tolerant. This means that these defects do not severely disrupt charge carriers, leading to reduced recombination and enhanced PCE. Further computational studies expanded on this concept by examining the role of ions in perovskites, providing deeper insights into both deep and shallow defects. Although the sharp absorption edge and unique defect tolerance of perovskites have been acknowledged, recent research studies have shed more light on the nature of these defects. This increased understanding highlights the necessity of defect passivation to mitigate their adverse effects, which is essential for producing high-performance PSCs.21–23
![]() | ||
| Fig. 4 Schematic of defect types in PSCs. (A) The consequences of imperfection effects in the PSCs (B). | ||
| VOC = VradOC − ΔVnon-radOC | (1) |
Since defect formation largely occurs during the fabrication of perovskite, and due to the inherently defective nature of the surface and CTL/perovskite interfaces in a PSC, researchers have generally considered two primary approaches for PSC defect passivation, perovskite bulk and interface passivation. Within the scope of this review, we will sequentially focus on the use of polymers in these two main approaches to passivation.
![]() | ||
| Fig. 6 A summary chart of polymer incorporation methods in perovskites, ex situ (A) and in situ (B). | ||
On the other hand, when non-polar polymers are utilized, they can be introduced via the anti-solvent method.37 During the fabrication of thin perovskite films, an anti-solvent (such as chlorobenzene, toluene, or diethyl ether) is employed to rapidly remove the primary solvent (such as DMF or DMSO) during the spin-coating process. This rapid extraction triggers the immediate nucleation of the perovskite phase. When polymers are incorporated into the anti-solvent, they interact with perovskite precursors, altering the kinetics of crystallization by either decelerating the process or modulating grain formation. This modification ensures more regulated grain growth, thereby minimizing the formation of excessive grain boundaries and structural defects. The incorporation of polymers can influence the surface energy of the perovskite film, thereby improving its compatibility with CTLs. Furthermore, hydrophobic polymers establish a protective layer over the perovskite surface, serving as a barrier against moisture ingress and enhancing the material's overall stability.38 Certain polymers also mitigate residual mechanical stress and strain, reducing the risk of crack formation or pinhole defects within the perovskite film.39
It should be considered that strong polymer–PbI2 interactions can increase nucleation density but reduce grain size, making it difficult to achieve a well-dispersed, homogeneous polymer–perovskite composite. This often leads to suboptimal film quality and diminished performance enhancements.40
Totally, the molecular structure of polymer passivation agents and their interaction with perovskite crystals are key factors that are not fully understood. As a result, selecting an appropriate polymeric surface passivation agent often requires complex trial-and-error processes. Therefore, a guideline for identifying the most effective passivating functional groups within the polymeric structure to neutralize charged defects and extend the device lifespan is greatly needed. Fig. 7 summarizes some of the most commonly used insulating polymers in hybrid PSCs.
The agent used in a study is 4-hydroxybenzoic acid (4-HBA), which acts as a crosslinking agent that polymerizes in situ to form a network, playing multiple critical roles in improving the performance of PSCs, especially flexible versions (Fig. 8).46 The primary role of 4-HBA is to serve as a multifunctional additive that controls the growth of the perovskite film and enhances its final properties. It forms an in situ cross-linking network that acts as a template or formwork for high-quality perovskite crystal growth, and significantly decreases grain boundary defects, lowers the film's residual stress, and reduces its Young's modulus, making it more flexible. The network introduces dynamic hydrogen bonds that give the perovskite film the ability to self-heal after mechanical damage. Unlike many insulating polymer additives, the 4-HBA network improves charge transport and extraction, avoiding the issue of electrical insulation. The hydrophobic nature of 4-HBA enhances the perovskite film's resistance to environmental factors like humidity. The process by which 4-HBA improves the perovskite film occurs in distinct stages during fabrication and operation. During the mild heating in the annealing stage (e.g., 100 °C for 30 min), 4-HBA molecules undergo an esterification reaction, causing them to crosslink with each other. This creates a network of oligomers with an average chain length of about seven units, held together by both covalent ester bonds and intermolecular hydrogen bonds. This network provides an excellent scaffold for the perovskite to grow on. Before and during crystallization, the functional groups on 4-HBA interact strongly with the perovskite precursors. The carbon-oxygen double bond (C
O) forms a coordination bond with uncoordinated Pb2+ ions. This strong interaction limits the number of excessive perovskite nucleation sites and prolongs the overall crystallization process. By slowing down and controlling the crystal growth, this method allows for the formation of a high-quality perovskite film with significantly larger grains (average size of 937 nm vs. 584 nm for the control) and fewer defects.
![]() | ||
Fig. 8 (A) A schematic diagram that illustrates how the 4-hydroxybenzoic acid (4-HBA) crosslinking agent interacts with perovskite precursors to form a high-quality film, compared to the standard process. The C O/C–O groups in 4-HBA form coordination bonds with lead (Pb2+) ions, while the hydroxyl (–OH) groups form hydrogen bonds with iodide (I−) ions. This anchors the agent directly to the perovskite components. The benzene rings of the 4-HBA molecules interact with each other (π–π stacking) and with Pb2+ ions (electrostatic interaction), which helps enhance the film's electrical conductivity. (B) These three figures are a series of SEM images that visually demonstrate the mechanical resilience and remarkable self-healing properties of the perovskite film modified with 4-HBA after being subjected to extreme stress. For this test, the films were bent 4000 times at a tight radius of 2 mm. The left picture (control film): the “before healing” image shows that the standard perovskite film suffers from severe, widespread cracks after bending. The “after healing” image is identical, showing that the film has no ability to self-heal. The middle picture (3-BA modified film): this film, modified with a similar but non-crosslinking molecule, also exhibits serious cracks from the stress and shows no self-healing. The right picture (4-HBA modified film): this film is far more robust. The “before healing” image shows only slight, minor cracks after the same intense bending. Most importantly, the “after healing” image reveals that after a mild heat treatment (60 °C for 1 hour), the cracks have completely disappeared. Reproduced with permission from Yankun Yang, et al., Advanced Science, 2025, Wiley.46 | ||
After the film is formed, the 4-HBA network remains at the grain boundaries, acting as a flexible bridge that connects the rigid perovskite grains. This helps to release residual lattice stress and significantly lowers the film's Young's modulus, making it much more robust against bending and stretching. The network is rich in dynamic hydrogen bonds. When the film is bent and cracks appear, these bonds can break and then rapidly reform with mild heat treatment (e.g., 60 °C), allowing the cracks to heal and the device to recover up to 89% of its original efficiency. The benzene ring in 4-HBA has an electrostatic interaction with Pb2+ ions, and the molecules can arrange to allow for π–π stacking. This creates effective channels for charge extraction and transport, overcoming the insulation problem often caused by other polymer additives. The efficiency of rigid PSCs reached 24.76%, while FPSCs achieved 22.73%, and the PSCs retained 91% of their initial efficiency after 10
000 bending cycles at a 5 mm radius.
Developing complex polymers with three-dimensional (3D) structures and multiple functional groups is a current trend for addressing deep-level traps in various perovskites. One efficient approach is multi-mode passivation. In a study,47 a singular functional group within the polyethyleneimine (PEI) family matrix with various configurations and protonation states was used. The in situ protonation process reduced deep-level traps for minority carriers. The research emphasized the importance of primary amine branches with moderate density for optimal passivation effects, highlighting the limitations of traditional protonation methods in addressing carrier traps. The protonation process of perovskite/PEI fundamentally differs from physisorption or metal-chelation mechanisms. The study employed deep-level transient spectroscopy (DLTS) to investigate variations in the properties of deep-level traps, such as trap category, energy level, carrier lifetime, and cross-section area of captured charges (Fig. 9). The p–i–n device structure consists of ITO/PTAA/perovskite/PEIE/C60/BCP/Ag, with the hole carrier acting as the minority carrier at the perovskite/PEIE junction. The DLTS spectra are illustrated through the linear fitting of DLTS peaks, and the energetic positions of these traps were deduced from the Arrhenius plot. In devices lacking the PEIE layer, two p-type deep-level traps for hole carriers were detected at energy levels of 0.61 and 0.73 eV above the valence band maximum. In contrast, a single p-type trap was identified in devices with the PEIE layer, located at an energy level 0.65 eV above the valence band maximum. The PEIE coating reduced trap density by a factor of ten compared to the standard device, significantly extending carrier lifetime. Another crucial trap parameter that impacts carrier density and distribution is the captured cross-section area. This research demonstrates that in situ protonation is essential for effective interface passivation. The presence of non-protonated PEIE molecules in excess PEIE significantly increases insulation, leading to noticeable decreases in JSC and FF. Additionally, the reduction of the perovskite's work function is not the primary factor driving the performance improvement. Statistical analysis shows that the 15% increase in PCE of PEIE devices compared to control devices is mainly due to enhancements in VOC and FF.
![]() | ||
| Fig. 9 (A) Deep-level transient spectroscopy (DLTS) measurements comparing unmodified perovskite (control, black curve) and PEIE-incorporated perovskite devices (red curve) reveal three distinct minority carrier trap states, denoted as D1, D2, and D3. The spectral analysis demonstrates significant differences in trap state signatures between the bare and surface-passivated perovskite samples. And comparative study of the energy band structure and deep-level trap states was conducted for both control (B) and PEIE-passivated (C) devices. The observed peaks in the trap density profiles correspond to distinct defect states (D1, D2, and D3), where their spectral positions indicate trap depths and their amplitudes reflect relative trap densities. Key electronic parameters—including the conduction band minimum (EC), valence band maximum (EV), and Fermi level (EF)—were experimentally determined through ultraviolet photoelectron spectroscopy (UPS) measurements. This analytical approach provides critical insights into how PEIE passivation modifies the energetic landscape of the devices. Reproduced with permission from Zhu et al., Joule, 2022, Cell Press journal.47 | ||
The JSC remained unchanged, consistent with the overlap of external quantum efficiency (EQE) spectra. Furthermore, the diode ideality factor of the PEIE device (n = 1.03) showed significant improvement compared to the control device (n = 1.27), indicating reduced non-radiative recombination. PEIE devices exhibited remarkable shelf stability, maintaining their original performance when stored in an N2 atmosphere for 5500 hours. Notably, the PEIE devices retained their initial PCE after aging at 85 °C for 1100 hours, in contrast to the control device without PEIE passivation, which experienced a 10% reduction in PCE.
Li et al.48 developed a multifunctional fluorinated additive, capable of polymerizing through hydrogen bonds to control the intermediate phase and influence the preferred orientation of crystal growth. This additive, known as FTPA, was applied individually or simultaneously to the bulk and surface of various cell structures and perovskite compositions (Fig. 10). The FTPA additive includes delocalized aromaticity for charge carrier regulation, fluorine to stabilize the intermediate phase, and vinyl groups for in situ polymerization, which fills grain boundaries in the perovskite. By leveraging the chemical interactions of FTPA's functional groups with solution components, they effectively managed nucleation and crystallization, achieving a PCE of 24.10% (from 22.48%), with a VOC of 1.182 V and a FF of 83.45% for the top-performing planar FA0.95MA0.05Pb(I0.95Br0.05)3-based PSC. This PSC showed significant operational stability, maintaining over 95% of its initial efficiency after 1000 hours of continuous sunlight exposure and 2000 hours in ambient air with ∼50% humidity. A known issue with formamidinium lead iodide (FAPbI3) perovskite is the instability of the α-black-phase, which easily converts to the δ-yellow-phase. In this study, the unsealed FTPA-incorporated perovskite film retained the black phase for more than 5 minutes of immersion in water due to the FTPA network preventing water molecule penetration.
![]() | ||
| Fig. 10 The molecular structure design of FTPA, and schematic diagram of the possible phase evolution of the nucleation and crystallization of FA-based mixed anion perovskites (FA0.95MA0.05Pb(I0.95Br0.05)3) during the film-forming process with (w) or without (w/o) FTPA. Reproduced with permission from ref. 48. Reprinted from Nature Communication, Mubai Li, et al., 2023, under CC BY license. | ||
In another study, the careful design of an elastomer-like polymer with consistent in situ polymerization during perovskite crystal growth led to significant improvements in both PCE and mechanical stability for flexible PSCs. Wu et al.49 demonstrated that incorporating bis((3-methyloxetan-3-yl)methyl)thiophene-2,5-dicarboxylate (OETC) into the PbI2 solution and subsequent deposition created a mesoporous PbI2 scaffold with fewer nucleation sites, allowing cations to penetrate the PbI2 network and form large perovskite grains with a preferential (100) crystal orientation (Fig. 11). Additionally, polymer residues in the grain boundaries impart flexibility to the perovskite layer and increase Young's modulus. Consequently, PCEs of 23.4% and 21.1% were achieved for flexible PSCs with active areas of 0.062 cm2 and 1.004 cm2, respectively, demonstrating the ability to relieve mechanical stress through the flexibility of the perovskite film.
![]() | ||
| Fig. 11 Schematic of the in situ crosslinking-assisted grain growth process. Reproduced with permission from ref. 49. Reproduced with permission from Yeyong Wu, et al., Joule, 2023, Cell. | ||
In a study, researchers used Rb-functionalized poly(acrylic acid) (Rb-PAA) as a multifunctional alkali-functionalized polymer to control the crystallization of FAPbI3 perovskite films. The synthesis of the Rb-PAA complex involves a simple acid–base neutralization reaction between Rb2CO3 and PAA (poly(acrylic acid)), resulting in Rb-PAA held together by ionic interactions. When dissolved in chlorobenzene, these ionic interactions may weaken. During the anti-solvent dropping process, chlorobenzene removes DMF and DMSO from the wet perovskite films, ensuring even dispersion of the Rb-PAA complex within the film, which allows for further diffusion and distribution of Rb ions during post-annealing at 150 °C. However, due to the large size and steric hindrance of the PAA polymer, it primarily remains on the film's surface rather than penetrating the bulk. Small angle X-ray diffraction showed that the signals of excess PbI2 in PAA and Rb-PAA treated films were significantly suppressed. SEM analysis revealed uniform films with compact textures and grain sizes in the hundreds of nanometers range, free of visible pinholes. Additionally, the Rb-PAA-treated film had a slightly larger average grain size (909 nm) compared to the control film (847 nm), similar to the PAA-treated film (907 nm), indicating fewer grain boundaries and defects. The UV-vis absorption spectra show absorption edges at around 800 nm, indicating a favorable light absorption range and suggesting that the optical properties (such as the optical bandgap) of the perovskite films remain unchanged after Rb-PAA and PAA treatments. This result implies that adding PAA does not affect the crystallization of perovskite, and Rb+ does not integrate into the crystal lattice to form Rb-based perovskite, which would otherwise result in a wider bandgap structure. The PL spectra indicate that the emission intensity of films treated with Rb-PAA is about twice that of films treated only with PAA or left untreated, suggesting that Rb+ doping effectively reduces defect density and minimizes non-radiative recombination. The recombination in time-resolved photoluminescence (TRPL) is mainly governed by first-order trap-assisted recombination. The fitted lifetime of perovskite films has significantly increased from 470 to 1059 ns due to the passivation effect of Rb+ and PAA. The enhanced PL intensity and significantly extended charge carrier lifetime of the treated perovskite film are clearly due to the reduced defects. Additionally, the extended charge-carrier lifetime of the perovskite film treated with Rb+ corresponds to the presence of micron-sized crystal grains with fewer grain boundaries, which act as recombination centers and charge carrier traps. The top-performing Rb-PAA polymer-treated device showed a PCE of 24.93% for the reverse scan, a PCE of 24.69% for the forward scan, and a stabilized efficiency of 24.74%. After 500 hours of operation, the PSC treated with Rb-PAA retained 83% of its initial efficiency, compared to 30 ± 5% for the PSC without Rb-PAA treatment.50
Table 1 lists the photovoltaic characteristics of perovskite solar cells that use more conventional polymers in their active layer.
| Device | Polymer | Device structure | VOC (V) | JSC (mA cm−2) | FF (%) | PCE (%) | Ref. |
|---|---|---|---|---|---|---|---|
| 1 | EVA | ITO/NiOx/Cs0.1FA0.7MA0.2PbIxBr3−x:EVA/PC61BM/BCP/Ag | 1.06 | 22.39 | 81.0 | 19.31 | 51 |
| 2 | PVA | FTO/TiO2/FAxMA1−xPbBryI3−y:PVA/spiro-OMeTAD/Au | 1.07 | 22.14 | 72.69 | 17.22 | 52 |
| 3 | PVB | FTO/TiO2/FAxMA1−xPbBryI3−y:PVB/spiro-OMeTAD/Au | 1.09 | 23.09 | 75.89 | 19.10 | 52 |
| 4 | 4Tm | ITO/SnO2/Cs0.05(FA0.87MA0.13)0.95Pb(I0.87Br0.13)3:4Tm/spiro-OMeTAD/Au | 1.17 | 23.79 | 79.05 | 22.06 | 53 |
| 5 | F4TCNQ | ITO/SnO2/PCBM/Cs0.09FA0.27MA0.64PbI3:F4TCNQ/P3HT/carbon | 1.07 | 22.23 | 63.68 | 15.1 | 54 |
| 6 | PTMA | FTO/TiO2/CsxFAyMA1−x−yPbBrzI3−z/PTAA/Au | 1.09 | 22.8 | 75.0 | 18.8 | 55 |
| 7 | PECL | ITO/NiOx/Cs0.05(MA0.15FA0.85)0.95Pb(I0.85Br0.15)3:PECL/PCBM + C60/BCP/Cr/Au | 1.15 | 23.98 | 83.79 | 23.11 | 56 |
| 8 | FTPA | FTO/SnO2/FA0.95MA0.05Pb(I0.95Br0.05)3:FTPA/FTPA/spiro-OMeTAD/Au | 1.18 | 24.43 | 83.45 | 24.10 | 57 |
| 9 | PEG | FTO/TiO2/MAPbI3:PEG/spiro-OMeTAD/Au | 0.98 | 22.5 | 72.0 | 16.00 | 58 |
| 10 | Rb-PAA | FTO/c-TiO2/m-TiO2/(FAPbI3)0.97(MAPbBr3)0.03:Rb-PAA/spiro-OMeTAD/Au | 1.18 | 25.42 | 82.8 | 24.93 | 59 |
| 11 | poly[Se-MI][BF4] | ITO/SnO2/MA0.08FA0.92PbI3:poly[Se-MI][BF4]/PEAI/spiro-OMeTAD/Au | 1.16 | 26.04 | 82.66 | 25.10 | 60 |
| 12 | PAA | FTO/SnO2/FAxMA1−xPbBryClzI3−y−z:PAA/spiro-OMeTAD/Ag | 1.16 | 25.20 | 82.46 | 24.19 | 61 |
| 13 | D2 | FTO/SnO2/Cs0.05(MA0.17FA0.83)0.95Pb(I0.83Br0.17):D2/spiro-OMeTAD/Au | 1.12 | 18.3 | 74.0 | 15.1 | 62 |
| 14 | D3 | FTO/SnO2/Cs0.05(MA0.17FA0.83)0.95Pb(I0.83Br0.17):D3/spiro-OMeTAD/Au | 1.07 | 18.2 | 74.0 | 14.4 | 62 |
| 15 | PEI | ITO/PEDOT:PSS/MAPbI3−xClx:PEI/PCBM/LiF/Ag | 0.97 | 22.63 | 64.4 | 14.07 | 63 |
| 16 | PCDTBT | ITO/PEDOT:PSS/MAPbIxCl3−x:PCDTBT/PCBM/Bphen/Ag | 0.94 | 21.71 | 77.0 | 15.76 | 64 |
| 17 | PVP | ITO/NiOx/MAPbI3:PVP/PC60BM/BCP/Ag | 1.08 | 21.57 | 75.0 | 17.5 | 65 |
| 18 | J71 | FTO/TiO2/MAPbI3:J71/spiro-OMeTAD/Au | 1.11 | 22.31 | 77.6 | 19.19 | 66 |
| 19 | PPC | ITO/SnO2/MAPbI3:PPC/spiro-OMeTAD/Ag | 1.13 | 22.81 | 77.8 | 20.06 | 67 |
| 20 | PANI | FTO/TiO2/(FAPbI3)0.85(MAPbBr3)0.15:PANI/spiro-OMeTAD/Au | 1.10 | 22.50 | 77.13 | 19.09 | 68 |
| 21 | F1 | FTO/TiO2/MAPbI3:F1/spiro-OMeTAD/Au | 1.09 | 20.7 | 73.0 | 16.41 | 69 |
| 22 | F2 | FTO/TiO2/MAPbI3:F2/spiro-OMeTAD/Au | 1.06 | 17.8 | 76.0 | 15.07 | 69 |
| 23 | F3 | FTO/TiO2/MAPbI3:F3/spiro-OMeTAD/Au | 1.07 | 18.5 | 78.0 | 16.37 | 69 |
| 24 | PMMA-AM | FTO/c-TiO2/SnO2/FA0.90MA0.03Cs0.07Pb(I0.92Br0.08)3/PMMA-AM/spiro-OMeTAD/Au | 1.22 | 24.12 | 78.99 | 23.24 | 70 |
| 25 | P–Si | ITO/SnO2/PCBA-(FAPbI3)x(MAPbBr3)1−x:P–Si/spiro-OMeTAD/Au | 1.13 | 25.1 | 77.0 | 21.5 | 71 |
| 26 | PBAT | ITO/NiOx/CsxMAyFA1−x−yPbBrzI3−z:PBAT/PCBM + C60/BCP/Cr/Au | 1.13 | 23.53 | 83.1 | 22.07 | 72 |
| 27 | C60-PU | FTO/SnO2/FAPbI3:C60-PU/spiro-OMeTAD/Ag | 1.15 | 22.44 | 82.66 | 21.36 | 73 |
| 28 | PTB7 | ITO/SnO2/Cs0.05FA0.79MA0.16Pb(Br0.17I0.83)3:PTB7/spiro-OMeTAD/Au | 1.16 | 21.40 | 71.49 | 17.75 | 74 |
| 29 | PBTI | ITO/Cu:NiOx/CsFAMA perovskite:PBTI/PCBM/ZrAcac/Ag | 1.13 | 22.91 | 79.8 | 20.67 | 75 |
| 30 | PEA | FTO/TiO2/MAPbI3:PEA/spiro-OMeTAD/Au | 1.08 | 22.89 | 76.3 | 18.87 | 76 |
| 31 | PEA | ITO/SnO2/(FAPbI3)1−x(MAPbBr3)x:PEA/spiro-OMeTAD/Au | 1.15 | 24.42 | 76.94 | 21.60 | 76 |
| 32 | PHIA | FTO/NiOx/MAPbI3:PHIA/PCBM/Rhodamine 101/Ag | 1.07 | 23.92 | 78.40 | 20.17 | 77 |
| 33 | N2200 | FTO/TiO2/MAPbI3:N2200/spiro-OMeTAD/Au | 1.05 | 22.4 | 76.0 | 17.9 | 78 |
| 34 | F-N2200 | FTO/TiO2/MAPbI3:F-N2200/spiro-OMeTAD/Au | 1.06 | 22.7 | 76.0 | 18.4 | 78 |
| 35 | PF-0 | FTO/TiO2/MAPbI3:PF-0/spiro-OMeTAD/Au | 1.07 | 22.8 | 75.0 | 18.1 | 78 |
| 36 | PF-1 | FTO/TiO2/MAPbI3:PF-1/spiro-OMeTAD/Au | 1.08 | 22.8 | 76.0 | 18.7 | 78 |
| 37 | MCE | ITO/SnO2/MAPbI3:MCE/spiro-OMeTAD/MoO3/Ag | 1.17 | 23.0 | 78.1 | 21.0 | 79 |
| 38 | S-Polymer | FTO/c-TiO2/m-TiO2/FA0.65MA0.35PbI3−xClx:S-polymer/spiro-OMeTAD/Au | 1.11 | 25.57 | 83.14 | 23.52 | 80 |
| 39 | s-PU | PDMS/hc-PEDOT:PSS/PEDOT:PSS Al4083/FAxMA1−xPbBryI3−y:s-PU/PCBM/PEI/hc-PEDOT:PSS/PDMS | 1.09 | 22.34 | 78.65 | 19.15 | 81 |
| 40 | PU-PMDS-IU | ITO/SnO2/MA0.25FA0.75PbCl0.6I2.4:PU-PMDS-IU/spiro-OMeTAD/Au | 1.15 | 25.34 | 79.5 | 23.25 | 82 |
| 41 | P(VDF-TrFE) | FTO/TiO2/MAPbI3:P(VDF-TrFE)/spiro-OMeTAD/MoO3/Ag | 1.13 | 22.95 | 77.0 | 20.04 | 83 |
| 42 | P(VDF-TrFE-CFE) | FTO/TiO2/MAPbI3:P(VDF-TrFE-CFE)/spiro-OMeTAD/MoO3/Ag | 1.13 | 22.84 | 75.0 | 19.14 | 83 |
| 43 | P(VDF-TrFE-CTFE) | FTO/TiO2/MAPbI3:P(VDF-TrFE-CTFE)/spiro-OMeTAD/MoO3/Ag | 1.12 | 22.59 | 75.0 | 19.08 | 83 |
| 44 | TET | PEN/ITO/PTAA/CsxFAyMA1−x−yPbBrzI3−z:TET/C60/BCP/Cu | 1.06 | 22.92 | 83.1 | 20.32 | 84 |
| 45 | DI | ITO/SnO2/FAxMA1−x PbI3:DI/spiro-OMeTAD/Ag | 1.14 | 24.9 | 80.8 | 23.0 | 85 |
| 46 | β-FV2F | ITO/MeO-2PACz/Cs0.05(FA0.98MA0.02)0.95 Pb(I0.98Br0.02)3:β-FV2F/PC61BM/BCP/Ag | 1.17 | 24.8 | 84.3 | 24.6 | 86 |
| 47 | |||||||
| 48 | TUEG3 | Glass/ITO/PTAA/MAPbI3:TUEG3/PC61BM/BCP/Ag | 1.10 | 20.61 | 76.7 | 17.42 | 87 |
| 49 | TUEG3 | PET/ITO/PTAA/MAPbI3:TUEG3/PC61BM/BCP/Ag | 0.99 | 20.04 | 69.0 | 13.64 | 87 |
| 50 | POSP | ITO/NiOx/CsxFAyMA1−x−yPbBrzI3−z:POSP/PCBM + C60/BCP/Cr/Au | 1.14 | 23.56 | 84.8 | 22.74 | 88 |
| 51 | PCL | Glass/ITO/PEDOT:PSS/MAPbI3:PCL/PC61BM/BCP/Ag | 1.01 | 19.35 | 74.14 | 14.49 | 89 |
| 52 | PS | ITO/TiO2/Cs0.05FA0.81MA0.14PbI2.55Br0.45:PS/spiro-OMeTAD/Au | 1.15 | 21.5 | 85.0 | 21.0 | 90 |
| 53 | PU | ITO/NiOx/MAPbI3:PU/PCBM/BCP/Ag | 1.05 | 22.12 | 80.3 | 18.7 | 91 |
| 54 | PMMA | ITO/NiOx:PDA/MAPbI3:PMMA/PCBM/BCP/Ag | 1.07 | 22.97 | 82.0 | 20.12 | 92 |
Fig. 12 shows the PCE change after addition of polymers in the bulk perovskite according to Table 1.
![]() | ||
| Fig. 12 Comparative analysis of PCE changes induced by selected conventional polymer integrations within the perovskite active layer, based on the device performance metrics presented in Table 1. | ||
Also, the researchers developed a two-step method involving pre-polymerization and secondary polymerization.97 In pre-polymerization, a solution of the monomer, such as acrylonitrile, is first heated for a short period (e.g., 60 seconds) before being applied to the solar cell. This initial step forms low molecular weight (LMW) polyacrylonitrile, which is less volatile than the original monomer. In secondary polymerization, this LMW polymer solution is then spin-coated onto the perovskite film's surface and heated again. This second heating completes the polymerization process, creating a stable, thin (∼3 nm) polymer network directly on the top interface. The resulting polyacrylonitrile layer passivates the perovskite surface through the formation of strong coordination bonds between its cyano groups (C
N) and the uncoordinated Pb2+ defects. This top surface passivation significantly reduces defect density, which leads to enhanced photoluminescence, longer carrier lifetimes, and ultimately, a more efficient and stable solar cell with a higher open-circuit voltage.
Using polymers for this purpose offers unique advantages in improving both the chemical and mechanical properties of this interface. For instance, one strategy involves adding hexamethylene diisocyanate (HDI) to the perovskite precursor solution and coating the underlying ETL with ethylene glycol (EG).99 During the annealing process, the HDI and EG react in situ to form polyurethane at the interface. The resulting polyurethane fastens the perovskite to the ETL through a network of hydrogen and coordination bonds. This strong adhesion improves the mechanical robustness of the device and mitigates residual stress in the perovskite film. The process also regulates crystallization by inhibiting the formation of complex intermediate phases, leading to a more uniform and higher-quality perovskite film (Fig. 13).
In another study,100 a pre-existing polymer is coated on the CTL before the perovskite is deposited. A HMW PVP is used as a multifunctional interlayer at the buried interface. The functional groups on the polymer (e.g., C
O in PVP) interact strongly with the ions in the perovskite precursor solution. Specifically, PVP forms hydrogen bonds with the ammonium cations (FA+ and MA+), creating polymer–ammonium intermediates. This interaction is stronger than that with lead ions and effectively slows down, or retards, the perovskite crystallization process. This slower, more controlled crystallization leads to a perovskite film with improved crystallinity and fewer defects. The polymer interlayer also helps create a better energy level alignment for more efficient charge extraction.
The key features of these two methods are summarised in Table 2. The most efficient PSCs often use both top and buried passivation. In this way a buried interface improves crystallization and charge extraction, and the top interface passivates surface defects and enhances stability.101
| Top interfacial passivation | Buried interfacial passivation | |
|---|---|---|
| a ALD: atomic layer deposition.b SAM: self-assembled monolayers. | ||
| Application timing | After perovskite formation | Before perovskite deposition |
| Main target | Surface defects | Bulk/interface defects |
| Processing method | Spin-coating, ALD,a and vapor-phase | SAMs,b interlayers, and substrate mod |
| Most impact on the device | Improves stability | Enhances charge transport |
| Limitations | May block charge extraction | Requires precise control |
Common defect passivation methods involve pre- or post-treatment of the perovskite film using solvents or ammonium salts, which can lead to uncontrolled changes in film morphology or bulk crystal phase, complicating the preparation process and increasing costs. Consequently, in situ passivation of defects at dual interfaces during perovskite film preparation has become a research hotspot. This approach enhances device performance by reducing defect density at both interfaces and improving perovskite crystallization, resulting in higher crystalline quality and improved PCEs.101
Research has shown that using the fluorinated polymer poly(vinylidene fluoride) (PVDF) as an additive in spiro-OMeTAD is beneficial for PSCs.102 PVDF copolymers have been effective in PSC development, serving as additives in perovskite films and materials for interface passivation. Specifically, P(VDF-TRFE) has demonstrated potential in enhancing PSC resilience when used as a passivation material at perovskite–transport layer interfaces. The hydrophobic nature of P(VDF-TRFE) as an additive to transport layers like spiro-OMeTAD is still to be explored. The composite of spiro-OMeTAD and PVDF improves HTL conductivity, reduces non-radiative recombination, and enhances FF and VOC compared to a spiro-based HTL. PSCs with P(VDF-TRFE) achieve an efficiency of 24.1%, surpassing that of the control PSCs at 21.4%. Additionally, these PSCs show impressive ambient and operational stability, retaining over 90% of their initial efficiency after 45 days without encapsulation and maintaining 94% of their initial efficiency after 1080 hours under nitrogen conditions. The SEM images reveal that the control HTL material has voids, indicating poor film quality, which could lead to non-radiative recombination. In contrast, the target HTL material exhibits a uniform film morphology, suggesting better quality and potential performance improvements. Nyquist plots show lower series and transfer resistance values in the target PSC, indicating improved charge transfer and extraction mechanisms. The higher recombination resistance suggests reduced recombination at the perovskite–HTL interface, aligning with enhanced FF and VOC. Dark I–V characteristics analysed using the SCLC model reveal smaller VTFL values for hole-only devices in the target PSC compared to the control PSC.
Point defects such as surface under-coordinated lead cations and iodine (I) anions of the PbI6−4 octahedron, ionic vacancies, interstitial ions, and anti-site defects are primary sources of non-radiative Shockley–Read–Hall recombination. This recombination results in the loss of VOC and accelerates the decomposition of PSCs. The presence of both ester and amide groups aids in the coordination of Pb2+ ions at the interface and the anchoring of I− ions, thus suppressing defects like interstitial I and Pb–I anti-site defects. In this context, a novel polymer, poly(methyl methacrylate-co-acrylamide) (PMMA-AM), was developed as an interfacial passivation layer. The interfacial passivation layer demonstrated exceptional stability during HTL deposition due to its limited solubility in chlorobenzene. This approach effectively mitigated defects in the perovskite layer, significantly increasing the VOC of the device from 1.12 V to 1.22 V, achieving a PCE of 23.24%. Additionally, a PCE of 20.64% was attained for a large-area perovskite module (14 cm2). The device modified with PMMA-AM also showed remarkable long-term stability, retaining 95% of its initial PCE after 1000 hours of continuous operation under illumination.103
In a study, the linear polymer heparin sodium (HS) is introduced as a multifunctional heterointerface bridge between the SnO2 ETL and the perovskite film in n–i–p solar cells104 (Fig. 14). The long-chain structure of HS allows it to simultaneously interact with both layers, leading to a combination of chemical, electronic, and mechanical improvements. The HS polymer has multiple functional groups (COO− and SO3−) and ions (Na+) distributed along its backbone. These groups form robust chemical bonds that neutralize defects at the interface. The COO− and SO3− groups form strong covalent coordination bonds with undercoordinated lead (Pb2+) defects on the perovskite surface and with tin (Sn4+) on the SnO2 surface. The Na+ ions interact with iodide, forming Na–I bonds that further passivate defects. This passivation is confirmed by a reduction in the trap density of the perovskite film. The HS layer optimizes the energy landscape for more efficient electron extraction. It raises the conduction band of the SnO2 layer from −4.35 eV to −4.16 eV, creating a better energy level alignment with the perovskite film. The HS layer acts as a template for perovskite crystallization, resulting in films with larger grain sizes and improved crystallinity. The HS modification effectively alleviates the inherent tensile stress found in perovskite films, converting it to a minimal compressive stress. This reduction in mechanical strain is critical for long-term stability. The polymer acts as a molecular bridge that chemically bonds the ETL and perovskite layers together. This dramatically improves mechanical toughness, as demonstrated by the fracture energy, which more than doubled for the HS-modified interface compared to the control. The optimized interface results in a PCE of 26.61% for rigid devices (certified at 26.54%) and 25.23% for flexible devices. They retain 94.9% of their initial efficiency after 1800 hours of continuous operation under simulated sunlight. They maintain 95.2% of their initial PCE after being aged at 85 °C for 1800 hours. Flexible devices retain over 95.5% of their efficiency after 1000 bending cycles.
![]() | ||
| Fig. 14 The explanation of the mechanism and impact of using an HS polymer as a bridge at the interface between the tin oxide SnO2 ETL and the perovskite film. A schematic that illustrates the core concept of the “heterointerface bridge”. The long-chain HS polymer is shown sandwiched between the SnO2 and perovskite layers. Its structure contains functional groups and ions (COO−, SO3−, and Na+) that are distributed along its backbone, allowing it to form robust chemical bonds with both the underlying SnO2 and the overlying perovskite. This dual interaction serves to both physically connect the layers and chemically passivate interfacial defects. This results in the successful formation of the intended bridge layer. Redrawn based on concept reported in Nature Photonics, Xiaodan Tang, 2025, under CC BY license.104 | ||
Hyperbranched polymers (HBPs) are unique polymers with a 3D branched molecular structure, numerous functional end groups, and nanometer-scale intramolecular voids.105 By manipulating their molecular configuration and functional end groups, the physical and chemical properties of HBPs can be precisely controlled. Incorporating complementary hydrogen bond donor and acceptor moieties within HBP branches creates an internal dynamic bonding network that absorbs mechanical energy and facilitates micro-fracture repair. The spherical shape and intramolecular voids of HBPs enhance branch mobility, enabling significant molecular deformation.106 These unique structures give HBP materials exceptional mechanical strength and resilience. Furthermore, the polar end groups of HBPs can form strong interactions with other substances, providing excellent adhesive properties. As a result, elastic adhesive HBPs have gained significant interest as promising materials for reinforcing interfaces in various optoelectronic devices (Fig. 15).105 HBPs with Pb2+ binding groups can be specifically designed to function as internal encapsulants, reducing lead leakage and creating a safer environment for daily use. A research study involved the synthesis of polyamide–amine-based HBPs with a range of branch lengths.105 These HBPs are rich in amide, primary amine, secondary amine, and carboxyl groups on their spherical surface and within nanometer-scale cavities. These polar groups form strong hydrogen bonds with SnO2 and perovskite, enhancing adhesion at the delicate ETL/perovskite interface. The flexible branches of HBPs allow them to rearrange hydrogen bond donors and acceptors, creating a dynamic, self-adjusting hydrogen bond network. The intramolecular cavities can also reorganize during bending, absorbing deformation energy. These features make HBPs excellent at damping and mitigating deformation energy, improving the mechanical resilience of flexible PSCs. This study found that rigid PSCs with HAD (hexane-diamine)–HBP modified SnO2 achieved an impressive PCE of up to 25.05%. Flexible PSCs (1 cm2) exhibited a PCE of up to 23.86% and demonstrated outstanding mechanical flexibility, retaining 88.9% of their initial PCE even after 10
000 bending cycles at a 3 mm radius. Furthermore, the carboxyl and amine groups within HBPs effectively interact with Pb2+, significantly reducing the risk of lead leakage even if the flexible PSCs experience structural failure. To examine how chain flexibility affects the mechanical properties of HBPs, researchers utilized four different alkyl spacers with varying chain lengths. These HBPs were labeled as EDA (ethylenediamine)–HBPs, BDA (butanediamine)–HBPs, HDA–HBPs, and ODA (octanediamine)–HBPs, corresponding to ethylenediamine, 1,4-butanediamine, 1,6-hexanediamine, and 1,8-octanediamine derived HBPs, respectively. The polar –NH2 terminal groups in HBPs can form high-density multiple hydrogen bonds with SnO2 and perovskite layers, enhancing adhesion. The inter- and intramolecular hydrogen bond interactions involving –NH–, –NH2, and C
O groups create a dynamic hydrogen bond network with strong mechanical properties, preventing crack propagation and increasing the toughness of the adhesive interface. HBPs with longer branches have a more relaxed structure, characterized by greater elasticity and flexibility. This is due to the extended aliphatic chains allowing ample single-bond rotation, reducing steric hindrance during synthesis, and resulting in a larger, more relaxed molecular configuration. Enhancing mechanical and adhesive properties in materials can be achieved by incorporating crystalline micro-domains in HDA–HBPs. Among the different HBPs tested, HDA–HBPs showed the best device performance, likely due to the strong interaction between the HBPs and adjacent SnO2 and perovskite layers.
![]() | ||
| Fig. 15 Interfacial bonding mechanism of hyperbranched polymers (HBPs) as an adhesive layer between SnO2 and perovskite layers. (A) Molecular architecture of HBPs. (B) Schematic illustration of HBP-mediated bonding at the SnO2/perovskite interface.105 Reprinted from Nature Communication, Zhihao Li, et al., 2023, under CC BY license. | ||
An HBP with dopamine adhesive (HPDA) is designed to enhance the mechanical durability of flexible perovskite solar cells (FPSCs), particularly under high humidity conditions.107 The polymer's effects are inspired by marine mussels, which use dopamine groups to achieve strong adhesion in wet environments. The key to the polymer's success is its dopamine end-groups. Lap-shear tests show that HPDA maintains excellent adhesive strength even when washed with water, a feat that a control polymer without dopamine could not achieve (Fig. 16). Due to its 3D hyperbranched structure, the HPDA polymer forms a vertical scaffold that permeates the entire perovskite film. This scaffold physically connects the bottom ETL to the top HTL through the bulk of the perovskite. This internal scaffold acts as a reinforcement network, making the perovskite film significantly tougher. The fracture strength increases from 13.21 MPa to 43.73 MPa, and the fracture energy nearly doubles. Crucially, after aging in a humid environment, the control film loses ∼80% of its fracture strength, while the HPDA-modified film loses only 10%. This prevents cracks from forming and propagating when the device is bent, especially in humid air. The HPDA scaffold effectively releases the residual tensile strain that typically builds up in perovskite films during crystallization. Measurements show a ∼73% reduction in this strain, which helps to prevent the formation of micro-cracks that can degrade the device. The oxygen-containing functional groups in HPDA coordinate with and passivate uncoordinated Pb2+ defects within the perovskite grain boundaries. At the bottom interface, the dopamine's catechol groups form strong bidentate hydrogen bonds with the SnO2 ETL. At the top interface, the benzene rings in the dopamine structure create π–π interactions with the spiro-OMeTAD HTL. The HPDA-modified flexible PSCs achieve a champion PCE of 24.43%. The polymer's strong affinity for lead ions also acts as a self-encapsulation mechanism, preventing 99% of lead leakage if the device is damaged.
![]() | ||
| Fig. 16 (A) The results of a lap-shear test, which measures the force required to pull apart two ITO substrates that have been glued together with HPDA. The test was conducted in various solvent environments to simulate the conditions present during PSC fabrication and operation. The curves show that while solvents like DMF, IPA, and CB slightly reduce the polymer's adhesion, HPDA still maintains a decent bonding strength. Most importantly, the curve for water (H2O) shows an excellent adhesive strength, comparable to the pristine (dry) condition. This demonstrates the polymer's remarkable ability to function effectively as an adhesive in a wet environment, a key feature inspired by marine mussels. (B) A bar chart to compare the underwater adhesion strength of HPDA with a control polymer, HPBA, which has a similar hyperbranched structure but lacks the critical dopamine end-groups. The results clearly show that after being washed with water, the adhesion strength of HPDA remains very high. This direct comparison proves that the dopamine terminal groups are essential for achieving strong underwater adhesion, highlighting the success of the bio-inspired design. Reprinted from nature communication, Z. Li, et al., 2025, under CC BY license.107 | ||
Tables 3 and 4 show the photovoltaic characteristics of perovskite solar cells that use more conventional polymers in the ETL and the HTL, respectively. Fig. 17A and B show the PCE change after the addition of polymers in the ETL and HTL according to Tables 3 and 4, respectively.
| Device | Polymer | Device structure | VOC (V) | JSC (mA cm−2) | FF (%) | PCE (%) | Ref. |
|---|---|---|---|---|---|---|---|
| 1 | HP | ITO/SnO2:HP/Cs0.05FA0.85MA0.10Pb(I0.97Br0.03)3/spiro-OMeTAD/Au | 1.162 | 25.0 | 79.2 | 23.03 | 108 |
| 2 | PEI | ITO/ZnO:PEI/PCBM/MAPbI3/spiro-OMeTAD/Ag | 1.04 | 19.77 | 75.0 | 15.38 | 109 |
| 3 | PEG | ITO/SnO2:PEG/Cs0.05FA0.81MA0.14Pb2.55IBr0.45/spiro-OMeTAD/Au | 1.12 | 22.67 | 81.9 | 20.8 | 110 |
| 4 | PEGDA | ITO/SnO2:PEGDA/FAPbI3/spiro-OMeTAD/MoO3/Ag | 1.14 | 25.24 | 81.0 | 23.31 | 111 |
| 5 | PVC | FTO/TiO2/PVC/MAPbI3/spiro-OMeTAD/AgAl | 1.07 | 23.65 | 74.75 | 19.02 | 112 |
| 6 | PMMA | FTO/c-In-TiO2/m-TiO2/PCBM:PMMA/Cs0.07Rb0.03FA0.765MA0.135PbI2.55Br0.45/spiro-OMeTAD/Au | 1.170 | 22.75 | 75.8 | 20.20 | 113 |
| 7 | PY-IT | ITO/2PACz/Cs0.05(FA0.98MA0.02)0.95Pb(I0.98Br0.02)3/PY-IT/PCBM/BCP/Ag | 1.15 | 24.37 | 84.0 | 23.57 | 114 |
| 8 | F8TBT | FTO/PEDOT:PSS/MAPbI3/PC61BM:F8TBT/Ag | 1.12 | 22.43 | 82.0 | 20.6 | 115 |
| 9 | PMMA | ITO/c-TiO2/TiO2 nanorods/PMMA:PCBM/Cs0.05FA0.88MA0.07PbI2.56Br0.44/PMMA/P3HT:CuPc/Au | 1.240 | 22.112 | 84.5 | 23.168 | 116 |
| 10 | PN4N | ITO/SnO2/PN4N/CsPbI2Br/PDCBT/MoO3/Ag | 1.3 | 15.3 | 81.5 | 16.2 | 117 |
| 11 | PMMA | FTO/TiO2:PMMA/MAPbI3/spiro-OMeTAD/Au | 0.979 | 18.70 | 76.3 | 14.0 | 118 |
| 12 | PFO | ITO/NiOx/MAPbI3/PC61BM:PFO/BCP/Ag | 1.03 | 16.3 | 64 | 10.8 | 119 |
| 13 | F8BT | ITO/NiOx/MAPbI3/PC61BM:F8BT/BCP/Ag | 1.04 | 19.28 | 74.7 | 15.0 | 119 |
| 14 | Parylene C | FTO/NiMgLiO/(Cs0.15FA0.85)Pb (I0.95Br0.05)3/parylene C/PCBM/BCP/Ag | 1.130 | 23.30 | 82.80 | 21.81 | 120 |
| 15 | PFNOX | ITO/PEDOT:PSS/MAPbCl3−XIX/PC61BM:PFNOX/Ag | 0.94 | 20.4 | 72.9 | 14.0 | 121 |
| Device | Polymer | Device structure | VOC (V) | JSC (mA cm−2) | FF (%) | PCE (%) | Ref. |
|---|---|---|---|---|---|---|---|
| 1 | PTAA | ITO/TiO2–Cl/Cs0.05MA0.05FA0.9PbI2.85Br0.15/PTAA/spiro-OMeTAD/Au | 1.126 | 23.2 | 79.8 | 20.9 | 122 |
| 2 | PTPD | ITO/TiO2–Cl/Cs0.05MA0.05FA0.9PbI2.85Br0.15/PTPD/spiro-OMeTAD/Au | 1.137 | 23.1 | 83.2 | 21.9 | 122 |
| 3 | polyTPD | FTO/c-SnO2/FA0.85MA0.15Pb(I0.85Br0.15)3/polyTPD/spiro-OMeTAD/Au | 1.167 | 23.3 | 79.0 | 21.37 | 123 |
| 4 | P3 | ITO/SnO2/PCBA/MAPbI3/P3/MoO3/Ag | 1.080 | 20.9 | 77.0 | 17.4 | 124 |
| 5 | 2DP-F | ITO/SnO2/p-FPhFA-incorporated MA0.16FA0.84PbI3/2DP-F/spiro-OMeTAD/MoO3/Ag | 1.162 | 24.95 | 80.40 | 23.31 | 125 |
| 6 | 2DP-O | ITO/SnO2/p-FPhFA-incorporated MA0.16FA0.84PbI3/2DP-O/spiro-OMeTAD/MoO3/Ag | 1.184 | 24.91 | 81.60 | 24.08 | 125 |
| 7 | PCDTBT | FTO/TiO2/MAPbI3/p-DTS(FBTTh2)2:PCDTBT/Au | 1.1 | 20.6 | 79.4 | 18.0 | 126 |
| 8 | PTAA | ITO/NPB:PTAA/MAPbI3/PC61BM/Al | 1.06 | 20.80 | 80.72 | 17.79 | 127 |
| 9 | PVDF-HFP | ITO/TiO2–Cl/Cs0.05MA0.05FA0.9PbI2.85Br0.15/PVDF-HFP/spiro-OMeTAD/Au | 1.131 | 22.4 | 79.3 | 20.1 | 122 |
| 10 | PVA | ITO/SnO2/FAyMA1−yPbClxBrzI3−x−z/PVA/spiro-OMeTAD/Au | 1.149 | 25.24 | 80.0 | 23.20 | 128 |
| 11 | PEG | ITO/SnO2/FAyMA1−yPbClxBrzI3−x−z/PEG/spiro-OMeTAD/Au | 1.126 | 24.94 | 79.6 | 22.35 | 128 |
| 12 | PVK | ITO/SnO2/FAyMA1−yPbClxBrzI3−x−z/PVK/spiro-OMeTAD/Au | 1.106 | 25.27 | 78.6 | 21.97 | 128 |
| 13 | PS | ITO/PTAA:F4-TCNQ/PS/MAPbI3/PCBM/Bphen/Al | 1.11 | 23.58 | 76.7 | 20.09 | 129 |
| 14 | PMMA | ITO/poly-TPD:F4-TCNQ/PMMA/MAPbI3/PCBM/Bphen/Al | 1.11 | 23.52 | 76.8 | 20.12 | 129 |
| 15 | PS | FTO/c-TiO2 + m-TiO2/Cs0.05(FA0.83MA0.17)0.95Pb(I0.83Br0.17)3/PS/spiro-OMeTAD/Ag | 1.158 | 22.32 | 79.2 | 20.46 | 130 |
| 16 | PFN-I | ITO/poly-TPD/PFN-I (double layer)/Cs0.05FA0.79MA0.16PbI2.4Br0.6/PCBM/BCP/Ag | 1.13 | 22.47 | 81.0 | 20.47 | 131 |
| 17 | PTQ10 | ITO/SnO2/FAPbI3/PTQ10/PTAA/Ag | 1.12 | 23.15 | 81.57 | 21.21 | 132 |
| 18 | PVBI-TFSI | FTO/c-TiO2/mp-TiO2/K0.05(MA0.15FA0.85)0.95PbI2.55Br0.45/spiro-OMeTAD:PVBI-TFSI/Au | 1.16 | 22.99 | 76.0 | 20.33 | 133 |
| 19 | BNo-F | ITO/BNo-F/MAPbI3/PCBM/BCP/Ag | 1.065 | 22.31 | 80.2 | 19.07 | 134 |
| 20 | p-NP-E | FTO/c-TiO2/m-TiO2/CsxFAyMA1−x−yPbBrzI3−z/p-NP-E/Au | 1.130 | 24.5 | 78.3 | 21.7 | 135 |
| 21 | PBDB-O | ITO/SnO2/CsxFAyMA1−x−yPbBrzI3−z/PBDB-O/MoO3/Ag | 1.072 | 16.98 | 56.24 | 10.23 | 136 |
| 22 | PBDB-T | ITO/SnO2/CsxFAyMA1−x−yPbBrzI3−z/PBDB-T/MoO3/Ag | 1.122 | 22.75 | 74.70 | 19.07 | 136 |
| 23 | PBDB-Cz | ITO/SnO2/CsxFAyMA1−x−yPbBrzI3−z/PBDB-Cz/MoO3/Ag | 1.144 | 24.41 | 79.03 | 22.06 | 136 |
| 24 | Asy-PSeDTS | FTO/TiO2/CsPbI3 PQDs/Asy-PSeDTS/MoOx/Ag | 1.25 | 15.8 | 77.0 | 15.2 | 137 |
| 25 | PTAA-P1 | ITO/PTAA-P1/Cs0.05(FA0.98MA0.02)0.95Pb(I0.98Br0.02)3/C60/bathocuproine/Ag | 1.17 | 25.50 | 83.28 | 24.89 | 138 |
| 26 | PFBTI | ITO/SnO2/Cs0.05FA0.95PbI3/PFBTI/MoO3/Ag | 1.16 | 24.6 | 80.8 | 23.1 | 139 |
| 27 | RP33 | FTO/SnO2/Cs0.05(FA0.83MA0.17)0.95Pb(I0.83Br0.17)3/RP33/Au | 1.02 | 23.5 | 68.7 | 16.44 | 140 |
| 28 | RP-OR | FTO/SnO2/Cs0.05(FA0.83MA0.17)0.95Pb(I0.83Br0.17)3/RP-OR/Au | 1.14 | 22.9 | 72.7 | 19.06 | 140 |
| 29 | CzAn | ITO/CzAn/FA0.8Cs0.2Sn0.5Pb0.5I3/PMMA/C60/BCP/Cu | 0.870 | 32.64 | 79.62 | 22.61 | 141 |
| 30 | TABT | ITO/SnO2/MAPbI3/Mes-TABT/Au | 1.15 | 23.8 | 77.0 | 21.3 | 142 |
| 31 | BNs | ITO/BNs/MAPbI3/PCBM/BCP/Ag | 1.099 | 21.75 | 79.6 | 19.03 | 134 |
| 32 | BNs-F | ITO/BNs-F/MAPbI3/PCBM/BCP/Ag | 1.083 | 22.78 | 83.40 | 20.56 | 134 |
| 33 | PEDOT:EVA | PET/ITO/PEDOT:EVA/MAxFA1−xPbBryI3−y/PCBM:BCP/Ag | 1.18 | 21.26 | 79.0 | 19.87 | 143 |
| 34 | PEDOT:EVA | Glass/ITO/PEDOT:EVA/MAxFA1−xPbBryI3−y/PCBM:BCP/Ag | 1.18 | 22.91 | 82.0 | 22.16 | 143 |
| 35 | PASQ-IDT | ITO/PTAA:PASQ-IDT/MAPbI3−xClx/C60/BCP/Ag | 1.14 | 22.41 | 82.47 | 21.07 | 144 |
| 36 | PS | ITO/PTAA:PS/MAPbI3(Cl)/PCBM/BCP/Ag | 1.138 | 22.5 | 83.5 | 20.8 | 145 |
| 37 | m-PFICZ | ITO/m-PFICZ/Cs0.15FA0.85PbI3/PCBM/Bphen/Ag | 1.062 | 24.21 | 77.30 | 19.87 | 146 |
| 38 | p-PFICZ | ITO/p-PFICZ/Cs0.15FA0.85PbI3/PCBM/Bphen/Ag | 1.091 | 24.48 | 80.1 | 21.39 | 146 |
| 39 | PBTI | ITO/SnO2/Cs0.05FA0.95PbI3/PBTI/MoO3/Ag | 1.14 | 24.7 | 80.4 | 22.6 | 139 |
![]() | ||
| Fig. 17 Comparative analysis of PCE changes induced by selected conventional polymer integrations within the (A) ETL and (B) HTL, based on the device performance metrics presented in Tables 3 and 4, respectively. | ||
Based on Fig. 17 the addition of PMMA and PEI in the TiO2 and ZnO layers, respectively, has had an impressive effect on increasing the PCE of devices. On the other hand, the addition of the PEDOT:EVA in the ITO and also PASQ-IDT in the PTAA have had a good effect on increasing PCE.
The alignment of energy bands at the heterojunctions creates the built-in electric fields and energetic pathways that control all aspects of charge carrier dynamics. A fundamental understanding of these interfacial physics is therefore an essential prerequisite for designing high-efficiency devices. The optimization of the Conduction Band Offset (CBO) and Valence Band Offset (VBO) is a critical determinant of the power conversion efficiency in PSCs, as these offsets directly govern charge carrier transport and recombination dynamics at the interfaces. The alignment can be broadly categorized into two types, cliff and spike, each with distinct consequences for device performance.147 A cliff alignment occurs when the energy level of the transport layer provides a downward step for the majority carrier. In contrast, a spike alignment presents a small energy barrier to the majority carrier. Research has shown that a moderate spike is highly beneficial for device performance. This small barrier does not significantly impede the forward extraction of majority carriers (e.g., electrons into the ETL), but it serves two crucial functions. First, it effectively blocks the back-injection of electrons from the ETL into the perovskite. Second, and more importantly, it creates an energy barrier that repels minority carriers (e.g., holes in the perovskite) from the interface, thereby suppressing interfacial recombination. However, the magnitude of the spike is critical.148 The VOC is fundamentally limited by recombination. Since interfacial recombination is a dominant loss mechanism in many PSCs, minimizing it is crucial for achieving high voltages. A moderate spike alignment at both the ETL and HTL interfaces is one of the most effective strategies for suppressing this recombination channel, thereby maximizing the achievable VOC, and a defective interface with high recombination rates can lower shunt resistance and this leads to a reduced FF.149
In insulating polymers, the absence of π-conjugation results in a very large energy gap between the HOMO and LUMO, typically exceeding 4 eV. This wide bandgap makes these materials electrical insulators, or dielectrics. Instead, their utility in PSCs stems from their dielectric nature and chemical functionality, which allow them to serve as passive interfacial layers. Their primary role is not to transport charge but to mitigate performance losses by neutralizing defects and electrically isolating problematic regions of the perovskite film. The principal function of insulating polymers at the perovskite interface is to suppress non-radiative recombination by passivating defects. This is achieved through two distinct but complementary mechanisms that leverage both their chemical and physical properties. Chemical passivation targets molecular-level point defects, while physical passivation is driven by the mesoscopic topography of the film, suggesting that the optimal approach may depend on the dominant defect type in a given perovskite film.150 The combined effect of chemical and physical passivation by insulating polymer interlayers leads to significant improvements in key photovoltaic parameters, primarily the VOC and FF.151 A crucial design parameter for this strategy is the thickness of the polymer layer. The layer must be sufficiently thin, ideally a sub-monolayer or monolayer, to allow charge carriers to efficiently tunnel through it to reach the adjacent CTL.
The primary role of conjugated polymers in PSCs is to function as active charge transport layers. They are designed to provide efficient, selective pathways for either holes or electrons to be extracted from the perovskite absorber and transported to the appropriate electrode. Many conjugated polymers exhibit high intrinsic charge carrier mobility, often surpassing that of commonly used small-molecule transport materials like spiro-OMeTAD, which is a significant advantage for minimizing transport losses.
The fundamental difference between insulating and semiconducting polymers at the interface lies in their mechanism of action. The former primarily acts by passivating defects, whereas the latter enhances the VOC not only through defect passivation but also by facilitating efficient charge transport and improving the energy level alignment between the perovskite and the CTL.
| Polymer name | Bandgap (eV) | Charge mobility (cm2 V−1 s−1) | Primary applications |
|---|---|---|---|
| Polyacetylene (PA) | 1.4–1.8 | 10−5–10−1 (undoped–doped) | Early conductive polymers |
| Poly(3-hexylthiophene) (P3HT) | 1.9–2.1 | 0.01–0.1 (hole) | OPVs and OFETs |
| Polypyrrole (PPy) | 2.7–3.2 | 10−4–10−1 (hole, doped) | Sensors and antistatic coatings |
| Poly(p-phenylene vinylene) (PPV) | 2.4–2.5 | 10−5–10−3 (hole) | OLEDs (green emission) |
| Polyfluorene (PFO) | 2.8–3.2 | 10−3–10−2 (hole) | Blue OLEDs and lasers |
| Polyaniline (PANI) | ∼3.5 (base) | 10−2–100 (hole, doped) | Conductive films and corrosion protection |
| Polythiophene (PT) | 2.0–2.2 | 10−5–10−4 (hole) | OFETs and electrochromics |
| Polycarbazole (PCz) | ∼3.5 (PVK) | 10−6–10−5 (hole) | Perovskite solar cells and OLEDs |
| Polymer name | Structure type | Bandgap (eV) | Charge mobility (cm2 V−1 s−1) | Key properties | Primary applications |
|---|---|---|---|---|---|
| PAMAM dendrimers | Polyamidoamine (core–shell) | N/A | Insulating (σ < 10−9) | High surface functionality | Drug delivery and gene therapy |
| Boltorn® hyperbranched polyester | Aliphatic polyester (random branches) | N/A | Insulating (σ ∼ 10−12) | Low viscosity and solubility | Coatings and adhesives |
| Conjugated dendrimers (e.g., PPP-Ph) | Polyphenylene branches | 2.5–3.0 | 10−4–10−3 (hole) | Light-harvesting and energy transfer | OLEDs and sensors |
| Hyperbranched polythiophene (hb-PT) | Thiophene branches | 1.8–2.2 | 10−3–10−2 (hole) | High solubility and film-forming | OPVs and OFETs |
| Dendritic porphyrin polymers | Porphyrin core + conjugated arms | 1.5–2.0 | 10−4–10−3 (ambipolar) | Photocatalysis and NIR absorption | Photodynamic therapy and solar cells |
| Hyperbranched PPV (hb-PPV) | PPV with 3D branches | 2.0–2.5 | 10−4–10−3 (hole) | Enhanced light emission | LEDs and bioimaging |
| Carbazole dendrimers | Star-shaped carbazole cores | 2.8–3.2 | 10−5–10−4 (hole) | High triplet energy | Host materials in OLEDs |
| Dendritic triarylamines | Star-shaped triarylamine cores | 2.5–3.0 | 10−3–10−2 (hole) | Hole transport and thermal stability | OLED hole transport layers |
| Hyperbranched DPP polymers | DPP-core with conjugated arms | 1.3–1.6 | 10−2–10−1 (ambipolar) | Low bandgap and NIR absorption | OPVs and photodetectors |
| Donor (D) unit | Acceptor (A) unit | Polymer name | Bandgap (eV) | Charge mobility (cm2 V−1 s−1) | Primary applications |
|---|---|---|---|---|---|
| Carbazole | Benzothiadiazole (BT) | PCDTBT | 1.7–1.9 | ∼10−3 (hole) | OPVs (PCE ∼7%) |
| Cyclopentadithiophene (CPDT) | Benzothiadiazole (BT) | PCPDTBT | 1.4–1.6 | ∼0.1 (hole) | Low-bandgap OPVs |
| Fluorene | Benzothiadiazole (BT) | F8BT | 2.1–2.3 | ∼10−3 (ambipolar) | OLEDs (green emitter) |
| Dithienosilole (DTS) | Benzothiadiazole (BT) | PSBTBT | 1.5–1.7 | ∼0.1 (hole) | High-efficiency OPVs |
| Thienothiophene (TT) | Thiazolo[5,4-d]thiazole (TTz) | PTTz | 1.3–1.5 | ∼0.5 (hole) | NIR photodetectors |
| Benzodithiophene (BDT) | Thieno[3,4-b]thiophene (TT) | PTB7 | 1.6–1.8 | ∼10−2–10−1 (hole) | OPVs (PCE ∼9–10%) |
| Benzodithiophene (BDT) | Fluorinated thieno[3,4-b]thiophene | PTB7-Th | 1.5–1.7 | ∼0.1–0.2 (hole) | OPVs (PCE > 10%) |
| Indacenodithiophene (IDT) | DPP (diketopyrrolopyrrole) | PIDT-DPP | 1.2–1.4 | ∼1–5 (ambipolar) | OFETs and flexible electronics |
| Naphthodithiophene (NDT) | DPP | PNDT-DPP | 1.3–1.5 | ∼0.5–2 (hole) | High-mobility OFETs |
| Carbazole | Benzothiadiazole (BT) | PCDTBT1 | 1.8–2.0 | ∼10−3 (hole) | OLEDs and OPVs |
| Thiophene | Quinoxaline | TQ1 | ∼2.0 | ∼10−2 (electron) | n-Type OFETs |
| Benzodithiophene (BDT) | Naphthalenediimide (NDI) | P(BDT-NDI) | 1.4–1.6 | ∼0.1–0.3 (electron) | All-polymer solar cells |
| Selenophene | DPP | PSeDPP | 1.3–1.5 | ∼0.5–1 (hole) | High-performance OFETs |
| Bithiophene | Isoindigo (IID) | PBTIID | 1.5–1.7 | ∼0.1–0.5 (ambipolar) | Flexible transistors and sensors |
| Dithienogermole (DTG) | DPP | PDTG-DPP | 1.2–1.4 | ∼1–3 (hole) | Stretchable electronics |
| Polymer | HOMO (eV) |
|---|---|
| Poly(3-hexylthiophene) (P3HT) | −4.9 to −5.2 |
| Polypyrrole (PPy) | −4.9 to −5.2 |
| Polyaniline (PANI) | −4.7 to −5.1 |
| PBTTT (poly(2,5-bis(3-alkylthiophen-2-yl)thieno[3,2-b]thiophene)) | −5.0 to −5.2 |
| DPP-based polymers (e.g., PDPP3T) | −5.1 to −5.3 |
| IDT-BT (indacenodithiophene-benzothiadiazole) | −5.2 to −5.4 |
| PTB7-Th | −5.2 |
| PCDTBT | −5.5 |
| PM6 (e.g., PBDB-T-2F) | −5.4 |
![]() | ||
| Fig. 18 Some important p-type conjugated polymers used in solar cells and their properties.159–165 | ||
![]() | ||
| Fig. 19 Some important n-type conjugated polymers used in solar cells and their properties.166–170 | ||
| Polymer | LUMO (eV) |
|---|---|
| Poly(benzimidazobenzophenanthroline) (BBL) | −4.3 |
| Poly(perylene diimide) (PPDI) | −4.0 |
| Poly(thiazole) (PTz) | −3.8 |
| P(NDI2OD-T2) (N2200) | −4.0 |
| P(PDI2OD-T2) | −3.9 |
| PFT | −3.9 |
| N2200 | −4.0 |
| PZ1 | −3.7 |
| P(NDI-T2) | −3.9 to −4.1 |
| Polymer | Structure | µh (cm2 V−1 s−1) | µe (cm2 V−1 s−1) |
|---|---|---|---|
| DPP-TT-T | Diketopyrrolopyrrole + terthiophene | 0.1–1.3 | 0.05–0.8 |
| PBTIID | Bithiophene + isoindigo | 0.1–0.4 | 0.1–0.3 |
| PCDTPT | Cyclopentadithiophene + pyridal thiadiazole | 0.2–0.5 | 0.1–0.4 |
SCPs are used as charge transporting layers,176,177 interfacial layers,178 and additives179,180 in perovskite solar cells. These polymers can be used as electron or hole transport materials. They facilitate the movement of charge carriers from the perovskite layer to the respective electrodes, which is essential for the conversion of light into electricity. Adjusting the LUMO level of SCPs with the LUMO level of the perovskite is necessary for polymer selection as a ETM, and suitable alignment of the HOMO level of the polymer with the HOMO level of the perovskite makes the polymer the HTM. The benefits of SCPs compared to traditional HTMs and ETMs are categorized in Fig. 20 and 21, respectively. In addition, they can act as interlayers between the active perovskite material and the charge-transporting layers. These polymer interlayers can improve the interface's electronic properties, leading to better charge extraction and reduced recombination. Also, polymers can be added to the perovskite layer to adjust its nucleation and crystallization processes. This can lead to improved film formation, increased grain size, reduced defects, enhanced charge transport, increased stability, and optoelectronic property tuning, which are beneficial for the device's efficiency and stability.
| Side chain type | Example polymers | Impact on performance | Trade-offs |
|---|---|---|---|
| Linear alkyl (e.g., –C6H13) | P3HT | Enhances π–π stacking | May reduce crystallinity if too long |
| Branched alkyl (e.g., –2-ethylhexyl) | PTB7-Th | Prevents excessive aggregation | Can disrupt backbone packing |
| Short alkyl (e.g., –C4H9) | P3BT | Promotes strong intermolecular ordering due to less steric hindrance | Limited solubility in many common organic solvents, making the polymer difficult to process |
| Side chain type | Example polymers | Impact on performance |
|---|---|---|
| Alkoxy (e.g., –OC6H13) | PBDTTT-EFT | Improves donor–acceptor mixing |
| Ethylene glycol (e.g., –OEG) | PEG-modified P3HT | Enhances water dispersibility (for eco-friendly processing) |
| Side chain type | Example polymers | Impact on performance |
|---|---|---|
| Fluoroalkyl (e.g., –CF3) | PTB7-Th (fluorinated BT) | Lowers HOMO and increases VOC |
| Perfluoroarene (e.g., –C6F5) | PF-based acceptors | Enhances air stability |
| Side chain type | Example polymers | Impact on performance |
|---|---|---|
| Polydimethylsiloxane (PDMS) | PDMS-grafted DPP polymers | Improves mechanical flexibility |
| Side chain type | Example polymers | Impact on performance |
|---|---|---|
| Sulfonate (–SO3−) | P3HT-SO3Na | Allows aqueous processing |
| Ammonium (–NR3+) | Cationic PF derivatives | Useful for layer-by-layer deposition |
| Side chain type | Example polymers | Impact on performance |
|---|---|---|
| Oligoether-thiophene | p(g42T-T) | Balances solubility and ordering |
| Alkyl-phenyl | PffBT4T-2DT | Optimizes morphology in BHJs |
The structures of some SCPs are illustrated in Fig. 22.
Nanowire P3HT with the advantage of higher carrier mobility along both the vertical and parallel directions is a choice for facilitating hole extraction from the perovskite.189,219
![]() | ||
| Fig. 24 Major energy loss mechanisms in standard single-junction solar cells: (1) photon absorption and excitation – initial conversion of sunlight into electron–hole pairs. (2) Thermalization of hot carriers – energy loss as excited electrons relax to the band edge. (3) Recombination losses – wasted energy from electron–hole pair annihilation (radiative/non-radiative). (4) Contact resistive losses – efficiency reduction due to imperfect charge collection at electrodes. Reprinted from Nature, Irfan Ahmed, et al., 2021, under CC BY license.222 | ||
The miscibility diagram of SCP and perovskite solutions and blends is a graphical representation that shows the compatibility of the two materials when mixed.227,228 It typically displays the phase behaviour of the blend across different compositions and temperatures, indicating regions where the materials are fully miscible, partially miscible, or immiscible. In the context of SCPs and perovskites, miscibility is crucial for the performance of devices like solar cells, where the microstructure and phase morphology of the active layer can significantly impact efficiency. In layered perovskite solar cells, the operation mechanism involves energy transfer from layered to 3D-like perovskite networks. The morphology of the interpenetrating network can affect this energy transfer, thus impacting the efficiency and stability of the solar cell.
A heterojunction is the interface formed between two different semiconductor materials. This interface is crucial for the operation of the solar cell, as it drives the separation of charge carriers. A research study229 investigates and compares two types of heterojunctions: the BHJ and the pseudo-planar heterojunction (PPHJ). The primary difference lies in how the donor and acceptor layers are arranged. In the BHJ, the donor and acceptor materials are mixed into a single layer, creating a distributed heterojunction throughout the bulk of the material. The PPHJ structure is created by sequentially depositing the acceptor (N3) and then the donor (D18-Cl) materials in distinct layers. This results in a more ordered, vertically separated structure, creating a sharper, more defined interface between the layers. Unlike the mixed BHJ, the PPHJ has a clear vertical distribution with the acceptor material (N3) concentrated at the bottom and the donor material (D18-Cl) at the top (Fig. 25). This well-defined separation is crucial for efficient charge transport, as it creates clear pathways for electrons and holes to move to their respective electrodes with less chance of recombination. The ordered PPHJ structure suppresses charge recombination and improves carrier transportation compared to the BHJ. It exhibits higher electron and hole mobility and a more balanced charge transport, which contributes to a higher JSC and FF. The PPHJ structure has less energetic disorder, which means there are fewer electronic traps that can impede carrier movement. This leads to a higher (VOC). The PPHJ structure, with its hydrophobic long alkyl chains and distinct fiber network, provides a more effective barrier against moisture. This leads to significantly improved long-term humidity and operational stability compared to devices based on the BHJ structure. The optimized integrated perovskite/PPHJ solar cell achieved a champion PCE of 23.25%. This is a significant increase from the control device's PCE of 20.14%. In a high-humidity environment (≈25 °C, ≈80% relative humidity), the optimized device maintained 81% of its initial PCE after 1200 hours, and retained over 89% of its initial efficiency after 1000 hours of continuous illumination.
![]() | ||
| Fig. 25 Donor: the D18-Cl molecule is a polymer that functions as the electron donor in the heterojunction. The researchers chose this specific donor to pair with the acceptor N3 due to its optimal electronic and morphological properties, which are beneficial for achieving a larger (VOC). Acceptor: N3 is a non-fullerene acceptor (NFA) that acts as the electron acceptor. It's designed to absorb light in the near-infrared spectrum, extending the overall light-harvesting range of the solar cell. DIB, which stands for 1,4-diiodobenzene, is used as an additive. Its purpose is to improve the crystallinity and morphology of the N3 acceptor layer during the fabrication process. This helps reduce material damage when the subsequent donor layer is deposited on top. Reproduced with permission from Li He, et al., Advanced Functional Materials, 2024, Wiley.229 | ||
Using P3HT frequently results in a reduced VOC in PSC devices. This is due to the inadequate physical interaction between P3HT and PVK, hindering effective carrier transport. The benefits of this polymer encompass its ability to avoid the need for moisture-attracting additives, deliver high performance, ensure prolonged durability, and facilitate processing over large areas.161,198 Yuqian Yang et al.227 use the spinodal or metastable decomposition of P3HT in the perovskite layer that created a hetero-interface bi-continuous morphology with an interpenetrating network structure. In order to lower the system's Gibbs free energy (ΔG), spontaneous uphill diffusion (spinodal decomposition) took place. Molecules of P3HT clustered together creating a phase rich in P3HT, while PVK molecules also aggregated to establish a PVK-rich phase (Fig. 26). As both the PVK and P3HT phases materialized simultaneously and within the same area, they became entwined, resulting in the formation of a network structure that interpenetrates. Elemental analysis indicates that in the top layer, there is a decrease in Pb content originating from the perovskite and an increase in sulphur (S) content, primarily from P3HT. These findings imply the existence of a P3HT/PVK hetero-interface.
![]() | ||
| Fig. 26 This describes the interfacial structure created between P3HT and perovskite (PVK) when blended together. The two polymers form a phase-separated boundary that significantly influences charge transport properties in organic electronic devices. Reproduced with permission from Yuqian Yang, et al., Advanced Materials, 2024, Wiley.227 | ||
This method also brings gradual precipitation of the P3HT phase mitigating the energy barrier and increasing hole extraction and transportation through the P3HT/perovskite hetero-interface; thus carrier and energy loss can be effectively reduced. Following the modification with P3HT, the observed red shift in both the PL spectra and the UV-visible absorption spectra matched the data from PL mapping. This redshift is primarily due to the formation of the P3HT/PVK hetero-interface, which aids in the extraction of holes from the perovskite, a process enhanced by the p-type characteristics of P3HT. The PCE of the control device was calculated to be about 20.34%, and the addition of an optimal amount of P3HT soluble in chlorobenzene as an anti-solvent of perovskite brought up the PCE to about 23.02%. In addition, the water contact angles of the films are 58.96° and 96.73° for PVK and P3HT/PVK films, respectively.
While PTAA is a popular choice for an HTL in PSCs, it has several drawbacks that hinder device performance. PTAA has a poor interaction with the perovskite material it is in contact with. This can lead to a lower-quality perovskite film growing on top of it. The inherent structure of PTAA limits its ability to efficiently transport positive charge carriers (holes) from the perovskite to the electrode. To address PTAA's weaknesses, the researchers employed a side-chain modification strategy called cyclic alkoxylation.232 This involves attaching oxygen-containing ring structures to the side benzene groups of the PTAA backbone. The core of their design is the insight that the size and geometry of the attached ring are critical. Based on theoretical calculations, they determined that a six-membered ring offers the optimal balance between molecular stability and the desired electronic energy levels for an HTL. Following this strategy, they synthesized two new polymers for comparison, PTAAO5 (modified with a five-membered benzo[d][1,3]dioxole ring), and PTAAO6 (modified with an optimized six-membered dihydrobenzo[b][1,4]dioxine ring) (Fig. 27). The study found that the six-membered ring in PTAAO6 provides significant advantages over both the original PTAA and the five-membered ring version, PTAAO5. The specific bond angles within the six-membered ring of PTAAO6 allow for better molecular orbital alignment, leading to extended π-conjugation in the side groups. PTAAO6 has a deeper HOMO energy level (−5.12 eV) compared to PTAA (−5.07 eV), which creates a more favourable alignment with the perovskite layer for efficient hole extraction. The improved conjugation and more rigid molecular structure enhance charge transport, giving PTAAO6 a higher hole mobility (6.42 × 10−5 cm2 V−1 s−1) than PTAA (4.57 × 10−5 cm2 V−1 s−1). Ultimately, using PTAAO6 as the HTL resulted in a solar cell with an outstanding efficiency of 25.19% and excellent operational stability.
![]() | ||
| Fig. 27 (A) NICS distribution: this panel shows maps of the Nucleus-Independent Chemical Shift (NICS), a computational metric where lower values signify stronger aromatic conjugation. The side group in PTAAO6 exhibits a region of very strong shielding (NICS value ≈−30 ppm), indicating that the specific bond angles in its six-membered ring effectively extend the π-conjugation. In contrast, the conjugation in PTAAO5's side group is disrupted. (B) Natural Adaptive Orbitals (NAdOs): these images visualize the molecular orbitals. For PTAAO6, the orbitals are shown to be well-aligned and parallel along the C–O bonds, which confirms the molecular orbital alignment responsible for the extended conjugation. (C) Reorganization energy: this chart shows the calculated reorganization energies, which represent the energy required for a molecule to adjust its geometry during charge transport. According to Marcus theory, a lower reorganization energy typically leads to higher charge mobility. The results show a clear trend where PTAAO6 has the lowest reorganization energy of the three polymers, predicting that it should have the best charge transport capability. This is attributed to its rigid structure, which requires minimal reorganization when charged. (D) Experimental hole mobility: this panel presents the experimental verification of the theoretical predictions using the SCLC method. The measurements confirm that PTAAO6 has the highest hole mobility at 6.42 × 10−5 cm2 V−1 s−1, a significant improvement over standard PTAA (4.57 × 10−5). This result directly links PTAAO6's enhanced conjugation and low reorganization energy to its superior performance as a hole transport material. Reproduced with permission from Yuqian Sen Yin, et al., Advanced Energy Materials, 2025, Wiley.232 | ||
The modification of macromolecule chains can be effective for higher-performance hybrid PSCs. Rai et al.233 investigated the effect of polymer chains consisting of ester and thiophene groups in the format of a thin layer between FA0.92MA0.08Pb(I0.92Br0.08)3 perovskite and spiro-OMeTAD as the HTM (Fig. 28). For this purpose, two poly(bithiophene ester) (PBTE) and poly(terthiophene diester) (PTTDE) were used. More functional groups have worked more successfully; in this way the PL intensity from the perovskite film was quenched two-fold and 3.5-fold upon applying poly(bithiophene ester) and poly(terthiophene diester), respectively. In the same way, EIS indicates lower charge transport resistance and higher charge recombination resistance that can be attributed to higher concentration of the ester functionalities per thiophene unit. The findings indicate that the creation of Lewis acid–base complexes, involving Pb2+ ions with insufficient coordination in the perovskite structure and oxygen atoms within the ester groups, could potentially reduce recombination and enhance the efficiency of charge extraction. Furthermore, functional groups of polymers especially [C
O] can inhibit the formation of Pb0 even at 85 °C and about 45% relative humidity. Without any interface modification, the PSCs showed a PCE of 17.8 ± 0.2%. The incorporation of PBTE led to a PCE of 19.2 ± 0.2%, and the use of PTTDE further increased the PCE to 20.0 ± 0.1%, along with improvements in the ISC, VOC, and FF. Elevated VOC values suggest that the ester-functionalized thiophene polymer interlayers are effective in reducing interfacial recombination. Meanwhile, enhancements in FF indicate improved charge transfer efficiency between the perovskite layer and the HTM. The faster-decaying photoluminescence component's lifetime (τ1) commonly associated with the quenching of charge carriers through trap states or interfacial charge transfer, was reduced from 0.40 µs in the uncoated perovskite film to 0.23 µs with PBTE coating and further to 0.03 µs with PTTDE coating. The marked reduction in τ1 aligns with the onset of a swift charge-transfer mechanism at the interface between the perovskite and polymer, which facilitates the extraction of holes from the perovskite into the PBTE or PTTDE layers. This enhancement is probably linked to the polythiophene structure of the polymers and may be further supported by the creation of Lewis acid–base adducts between the perovskite's under-coordinated Pb2+ ions and the oxygen atoms present in the ester groups. In an ambient lab setting with about 45% humidity over two weeks, nonencapsulated devices showed a decline in PCE after an initial increase.
![]() | ||
| Fig. 28 (A) Chemical architectures of PBTE and PTTDE. Illustrated molecular configurations of the conjugated polymers PBTE (poly(benzodithiophene-ethylene)) and PTTDE (poly(thienothiophene-diketopyrrolopyrrole-ethylene)), highlighting their distinct structural motifs. (B) Dual-functional interface mechanism in PSCs. The proposed operational model demonstrates how these polymeric interlayers function at the perovskite/HTM junction in PSCs: passivation pathway (black arrows): the polymers mitigate interfacial defects and non-radiative recombination through coordinated chemical interactions with undercoordinated lead ions at the perovskite surface. Charge transport pathway (red arrows): optimized energy level alignment and π-conjugated backbones facilitate efficient hole extraction from the perovskite absorber layer to the HTM while blocking electron backflow. Reprinted from Advanced MaterialsTechnologies, Nitish Rai, et al., 2024, under the Creative Commons CC BY-NC-ND license.233 | ||
Xie et al. designed a donor–π–acceptor HTM: poly[4-(5-(4,8-bis(5-(6-((2-hexyldecyl)oxy)naphthalen-2-yl)thiophen-2-yl)benzo[1,2-b:4,5-b′]dithiophen-2-yl)-6-undecylthieno[3,2-b]thiophen-2-yl)-5,6-difluoro-2-(6(1,1,1,3,5,5,5-heptamethyltrisiloxan-3-yl)-hexyl)-7-(6-undecylthieno[3,2-b]thiophen-2-yl)-2H-benzo[d]-[1,2,3]triazole] (BDT-TA-BTASi) (Fig. 29).236 This polymer has a co-planar structure with π–π strong stacking and consequently a high charge mobility that is approximately close to that of doped spiro-OMeTAD. Furthermore, the presence of siloxane and alkyl side chains promotes solubility, hydrophobicity, and molecular packing. The PCE of this device approaches 21.53%. The device maintained 92% of its performance following a 1000-hour storage period in a nitrogen environment. Additionally, it demonstrated considerable stability when exposed to elevated temperatures of 60 °C and a relative humidity level of 80 ± 10%, preserving 92% and 82% of its initial efficiency after 400 hours, respectively.
![]() | ||
| Fig. 29 The chemical structure of poly (BDT-TA-BTASi). Reproduced with permission from Zhiqing Xie, et al., ACS Applied Materials & Interfaces, 2024, ACS.236 | ||
In another study,237 in a Double-layer Halide Architecture (DHA) free dopant-P3HT was used as the HTM to tackle the poor contact between perovskite and P3HT layers. DHA is formed by depositing a slim layer of wide-bandgap halide (WBH) perovskite onto a narrow-bandgap light-absorbing layer. This is achieved through an in situ reaction involving n-hexyl trimethyl ammonium bromide (HTAB) directly on the surface of the perovskite to produce HTAB0.3(FAPbI3)0.95(MAPbBr3)0.05. Upon applying P3HT to the DHA surface, the intertwining of its alkyl chains with those of HTAB (C6H13−) is expected to facilitate P3HT's self-organization. Additionally, the presence of the N+(CH3)3− group makes the perovskite surface less susceptible to damage in moist environments. To assess the function of DHA, the increased VOC in HTM-free DHA-based PSCs, along with the extended charge duration in the DHA device, indicate that the WBH layer successfully neutralizes charge traps on the perovskite's surface. In the sample containing DHA, P3HT shows a dense nanofiber morphology. While amorphous P3HT exhibits minimal charge mobility, with values around 10−5 cm2 V−1 s−1 the self-organized P3HT nanofibrils, which have robust π–π interactions, display a significantly enhanced mobility rate of 0.1 cm2 V−1 s−1 in the absence of any doping agents. The Raman test indicates that more of the P3HT polymers on the DHA surface are oriented along the preferred direction. The DHA device demonstrates a significant enhancement in PCE, about 23.3% for the best sample, linked to an increased VOC and FF, along with the elimination of hysteresis. Furthermore, the DHA device that incorporates P3HT exhibits a reduced series resistance (Rs) compared to the control device also using P3HT. Conversely, the DHA film presents a higher Rs relative to the control perovskite film, which can be attributed to the elevated aliphatic composition of the WBH. DHA devices without encapsulation and exposed to 85% humidity at ambient temperature preserved about 80% of their original efficiency after 1008 hours. In contrast, the control device deteriorated entirely in just 200 hours.
| Challenge | Solution |
|---|---|
| Film defects (e.g., pinholes) | Optimize ink formulation (additives and solvents) |
| Crystallization control | Use gas quenching or anti-solvent dripping |
| Edge effects | Adjust die design (e.g., multi-slot dies) |
| Reproducibility | Automated feedback systems for speed/flow control |
| Challenge | Solution |
|---|---|
| Slow reaction kinetics | Optimize temperature/vapor pressure |
| Limited thickness control | Layer-by-layer deposition |
| Equipment cost | Shared vacuum tools (e.g., with OLED industry) |
| PbI2 residuals | Precise stoichiometry tuning |
| Challenge | Solution |
|---|---|
| Non-uniform films | Optimize nozzle design, substrate heating, and spray distance |
| Pinholes/cracks | Use solvent engineering (e.g., mixed solvents) or additives |
| Low efficiency (vs. spin coating) | Post-treatment (e.g., gas quenching and solvent annealing) |
| Clogging of nozzles | Filter inks and use pulsed spraying |
Perovskite precursors (PbI2 + MAI/FAI) dissolved in mixed solvents (DMF/DMSO with viscosity modifiers) were optimized for a surface tension of 28–35 mN m−1, viscosity of 8–20 cP, and particle size of <1% of the nozzle diameter. Piezoelectric actuators create pressure waves to eject droplets. Typical drop velocity is 5–10 m s−1, and drop spacing is 20–50 µm. Droplets merge on the substrate via controlled coffee-ring effects, followed by multi-pass printing to achieve the desired thickness (typically 300–500 nm), and post-annealing at 100–150 °C for crystallization. Future directions include hybrid printing combined with laser scribing for monolithic interconnections, AI-driven optimization using machine learning for droplet placement correction, multi-material printing to enable simultaneous deposition of all device layers, and flexible electronics via direct printing on curved surfaces for wearable photovoltaic applications. Important challenges and proposed solutions for this approach are summarized in Table 20.
| Challenge | Solution |
|---|---|
| Nozzle clogging | Ink filtration (0.2 µm) and solvent additives |
| Coffee-ring effect | Mixed solvents and substrate heating (40–60 °C) |
| Film roughness | Multi-pass printing with drying intervals |
| Resolution limits | Electrohydrodynamic (EHD) printing for <5 µm features |
| Parameter | Typical range | Effect |
|---|---|---|
| Blade-substrate gap | 100–300 µm | Directly controls wet film thickness |
| Coating speed | 5–50 mm s−1 | Faster = thinner films, but affects crystallization |
| Substrate temp. | 30–80 °C | Higher = faster drying and smaller grains |
| Ink viscosity | 5–50 cP | Higher viscosity = thicker films |
| Challenge | Solution |
|---|---|
| Streaks/defects | Optimize blade edge roughness (<50 nm) |
| Non-uniform drying | Gradient heating or gas flow control |
| Coffee-ring effect | Solvent engineering (high boiling point additives) |
| Large grain growth | Anti-solvent dripping or vapor-assisted crystallization |
| Challenge | Innovation |
|---|---|
| Crystallization control | Hybrid thermal/vapor annealing stations |
| Edge definition | Laser-etched plates (5 µm precision) |
| Layer registration | CCD camera alignment systems |
| Challenge | Solution |
|---|---|
| Slow process | Multi-zone heating for faster vapor diffusion |
| PbI2 residuals | Optimize vapor pressure/temperature |
| Limited to flat substrates | Conformal vapor delivery systems |
| High equipment cost | Shared vacuum tools (e.g., with OLED production) |
Both VAD and HCVD are two-step deposition techniques for fabricating perovskite thin films, but they differ in their process details, precursor application, and scalability. Crucial differences are summarized in Table 25.
| Feature | VAD | HCVD |
|---|---|---|
| Crystallinity | May have lower crystallinity due to limited interdiffusion | Tends to yield higher crystallinity and larger grains due to better precursor interaction |
| Defect density | Higher defects due to limited conversion efficiency of PbI2 to perovskite | Lower defects as solution interaction enhances perovskite formation |
| Surface morphology | May suffer from grain boundary defects and pinholes | Produces smoother, pinhole-free films |
| Scalability | Can be scaled but requires vacuum-based evaporation, increasing cost | More scalable and compatible with R2R processing |
For a comparison between the methods described, see Tables 26 and 27.
| Method | Thickness control | Scalability | Speed | Material use | Film quality | Best for |
|---|---|---|---|---|---|---|
| Spin coating | High (10–500 nm) | None | Slow | <10% | Good | Lab R&D |
| Blade coating | Excellent (50–500 nm) | High | Medium | >90% | Very good | Pilot lines |
| Slot-die coating | Good (100–1000 nm) | Very high | Fast | >95% | Good | Mass production |
| Spray coating | Moderate | High | Medium | 80–90% | Moderate | BIPV, flexible |
| Inkjet printing | High (20–500 nm) | Medium | Slow-Med | >95% | Good | Custom patterns |
| Vapor deposition | Excellent (10–300 nm) | Medium | Slow | >98% | Excellent | Tandems |
| HCVD | Excellent (50–400 nm) | Medium–high | Slow | >95% | Excellent | High-end apps |
| Flexographic | Moderate (200–800 nm) | Very high | Fast | >90% | Good | R2R electrodes |
| Metric | Spin | Blade | Slot-die | Spray | Inkjet | Vapor | HCVD | Flexo |
|---|---|---|---|---|---|---|---|---|
| a TRL (Technology Readiness Level): scale from 1 (basic research) to 9 (full production).b CAPEX (capital expenditure): equipment/installation costs per MW capacity.c OPEX (operational expenditure): per square meter production costs. | ||||||||
| TRL levela | 4 | 6 | 8 | 5 | 4 | 6 | 5 | 5 |
| CAPEXb ($ per MW) | — | 0.8 M | 0.5 M | 1.2 M | 2 M | 5 M | 3 M | 0.3 M |
| OPEXc ($ per m2) | — | 15 | 10 | 20 | 40 | 80 | 50 | 12 |
| Yield potential (%) | — | 95 | 97 | 90 | 85 | 99 | 98% | 92% |
| Energy use (kWh m−2) | — | 0.5 | 0.3 | 1.2 | 2 | 5 | 3 | 0.8 |
Point defects create mid-gap trap states that capture free electrons/holes,263 and thus reduce VOC and charge carrier lifetime. Mobile defects allow halide ions to migrate under electric fields and cause current–voltage hysteresis and long-term degradation. Deep-level defects dissipate energy as heat instead of light and lower photoluminescence quantum yield (PLQY). Also, these defects accelerate moisture/oxygen penetration. These defects can be mitigated by the following methods. Halide compensation, so that excess I− (e.g., MAI) fills iodine vacancies. Lewis base additives that bind to undercoordinated Pb2+ reducing trap states. Polymer additives improve film uniformity and passivate grain boundaries. Mixed solvents (e.g., DMF:DMSO) lead to better crystallization. Thermal annealing heals defects but must be controlled to avoid decomposition. Light soaking helps in defect annihilation under illumination.264
| Roughness type | Electrical effect | Optical effect | Stability effect |
|---|---|---|---|
| Nanoscale (<50 nm) | Increased trap-assisted recombination | Minimal scattering | Localized degradation |
| Sub-micron (50–500 nm) | Poor interfacial contact → higher Rs | Haze losses (∼3–5% JSC) | Crack initiation sites |
| Micron-scale (>500 nm) | Shunting pathways → low FF | Severe scattering (>10% JSC loss) | Delamination risks |
Rapid drying in techniques like slot-die coating leads to island growth rather than smooth films. Non-Newtonian behaviour of ink rheology causes uneven material deposition. Solute migration during drying creates raised edges and depressed centres (coffee-ring effect). High vapor pressure solvents (e.g., chlorobenzene) cause abrupt phase separation. Poor ink-substrate adhesion causes dewetting patterns. Ether-based treatments yield smoother films than chlorobenzene.274 Perovskite-quantum dot blends show very smooth surfaces.275,276 The RMS roughness measurements showed values of 36.8 nm for the control perovskite film, 30.9 nm for the PVA-treated film, and 34.1 nm for the PEG-treated film. The reduction in roughness observed in the polymer-modified films is likely due to the development of a thin polymeric coating on the surface.277–279
| Temperature range | Primary degradation mode | Efficiency loss mechanism |
|---|---|---|
| 50–70 °C | Ion migration begins | Increased J–V hysteresis |
| 70–85 °C | MA+ cation evaporation | VOC drop (∼100 mV) |
| 85–100 °C | Halide segregation | Bandgap widening → JSC loss |
| 100–150 °C | PbI2 crystallization | Shunting paths → FF collapse |
| >150 °C | Complete decomposition | Device failure |
For large-scale production methods like blade-coating, the wettability of the surface is critical for depositing a uniform perovskite layer. Polymeric HTLs have shown distinct advantages over their small-molecule counterparts. According a study, polymeric HTLs such as poly-DBPP287 and poly-2PACz exhibit better surface wettability compared to conventional self-assembled monolayer (SAM) HTLs.288 This is because the polymer structure can present more hydrophilic groups (like phosphonic acid) on its surface, which helps the perovskite ink to spread more evenly during coating. The improved wettability and higher conductance of polymers like poly-2PACz, and poly-DBPP lead to better deposition homogeneity, which is confirmed by electroluminescence (EL) mapping that shows more uniform emission compared to devices using small-molecule HTLs (Fig. 32). This uniformity is crucial for large-area modules, where performance is often limited by the lowest-performing sub-cell.
![]() | ||
| Fig. 32 EL intensity mapping of the blade-coated PSCs based on (A) small-molecule DBPP as the HTL. The map displays significant colour variation, with distinct patches of lower (greener) and higher (yellower) intensity. This indicates a non-uniform light emission across the surface of the cell. (B) Polymeric poly-DBPP as the HTL. In stark contrast to the DBPP-based cell, this map is much more uniform in color and exhibits a significantly higher overall EL intensity (indicated by the brighter, more yellow appearance). Reprinted from Advanced Materials, Wiley, Feifei Wang, et al., 2025, under CC BY license.286 | ||
The transition from a precursor ink to a solid, crystalline perovskite film is a delicate process that can easily introduce defects. Polymers can help manage this process. Specifically designed polymers, like poly-DBPP, can interact with lead cations in the precursor ink through their functional groups (biphosphonic acid). This interaction slows down the perovskite crystallization speed, which allows for the formation of a film with higher crystallinity and fewer defects.287 By interacting with the perovskite precursors in the wet-film state, polymers like poly(p-xylylene) (PPX) can restrain the formation of undesirable intermediate phases and promote oriented crystallization.289 Polymers can be applied as ultra-thin passivation layers on top of the perovskite film. For example, a ∼1 nm layer of PPX deposited via CVD effectively passivates surface defects and trap states. This is demonstrated by a significant increase in PL intensity and a longer charge carrier lifetime, indicating a reduction in non-radiative recombination. Novel polymeric HTLs like poly-DBPP are designed with biphosphonic acid groups that can simultaneously bind to the underlying conductive substrate (ITO) and interact with the perovskite layer above. This dual interaction effectively passivates the crucial buried interface, reduces trap density, and suppresses charge recombination. In some advanced structures, a cross-linked polymer CLP can be used as an interlayer to inhibit the diffusion of mobile ions between different perovskite layers, which is a key mechanism for degradation and instability.287 The hydrophobic nature of polymers like PPX allows them to act as an effective barrier against moisture, a primary cause of perovskite degradation. Even a 1 nm layer of PPX can protect the perovskite film from degrading into lead iodide (PbI2) when exposed to ambient air.289 Conventional small-molecule SAM HTLs can suffer from poor UV stability, where UV radiation weakens their bond to the substrate and decomposes their molecular structure. In contrast, polymeric HTLs like poly-2PACz exhibit exceptional UV resistance, attributed to the delocalization of electrons along the conjugated polymer backbone and a stronger, more stable adhesion to the substrate. This allows devices to retain 80% of their initial efficiency even after nearly 500 hours of high-intensity UV illumination.288
The characteristics of the polymer used, including molecular weight, polydispersity, and interchain interactions, greatly affect the ink's viscosity and the performance of solar cells by influencing the absorbers' shape and mechanical properties.290 Many high-performance polymers show temperature-dependent aggregation, meaning their properties change with temperature during processing.291–293 Despite advancements in processing parameters, ensuring the scalability and consistency of polymers across different batches remains a significant challenge.
For example, PEO can adjust shear-thinning behavior for slot-die coating.294 PEG and its family (PEGAD and PEGDAM) reduce the coffee-ring effect in printable perovskite solar cells.295 PVP stabilizes colloidal perovskite precursors.296,297 Fluorinated polymers such as PTFE reduce surface tension for uniform wetting on hydrophobic substrates.102,298 PMMA passivates undercoordinated Pb2+ via carbonyl groups and PMMA can passivate defects in the perovskite film, reducing trap states that hinder charge transport. PMMA chains on PTAA create an interface dipole at the PTAA/perovskite junction due to electrostatic interactions between the ester groups in PMMA and the electron-deficient N+ free radicals in PTAA. This physical adsorption enhances the VOC by 60 mV and enables a PCE exceeding 20%, while also ensuring excellent stability.299,300 PVA forms hydrogen bonds with perovskites, suppressing ion migration.301 PDMS enables self-healing of cracks.302
Polymers are the linchpin of perovskite commercialization, bridging the gap between lab-scale innovation and GW-scale manufacturing. Future advancements will focus on AI-designed smart polymers that autonomously optimize device performance under real-world conditions.
While techniques exist to enhance the operational stability of research solar cells, such as developing polymeric and oligomeric non-fullerene acceptors (NFAs) with high thermal stability, there has been no research on suitable materials for large-scale processing. These materials significantly impact the film morphology, especially polymeric NFAs, posing a challenge for large-scale printing. Processing a precursor ink with a polymeric donor and acceptor is particularly difficult due to the high likelihood of polymer self-aggregation, leading to severe phase segregation between the two polymers.303 This issue is especially pronounced when the meniscus-guided coating process involves slower solvent evaporation. Adding monoammonium zinc porphyrin as a surfactant to perovskite ink enhances the surface adherence of the precursor solution. It attaches to the surface of perovskite crystallites, preventing defect formation and cation escape. This leads to effective molecular encapsulation and surface passivation at grain boundaries and perovskite surfaces.304
The top CTL often serves as a passivator, reducing defects at the absorber surface and aiding in the extraction and transport of charge carriers. Conversely, the bottom CTL is crucial for templating film formation, significantly impacting the absorber's quality and overall device performance. Despite advancements in large-surface active layer printing methods, achieving uniform CTL thickness on flexible substrates with rough surfaces remains a challenge.305,306
Blade coating is commonly used for processing PTAA in inverted (p–i–n) PSCs. Fei and colleagues307 reported a 26.9 cm2 minimodule with a verified PCE of 21.8%, achieved by adding a small amount of bathocuproine (BCP), an ETL that can chelate with lead ions in the PTAA layer. This PTAA–BCP composite efficiently extracts and transports holes, improves the crystallinity of the perovskite layer near the HTL, and passivates the perovskite bottom surface, ensuring good contact between the perovskite layer and HTL.
At present, the P3HT polymer stands out as the only photoactive conjugated polymer that can be synthesized in a controllable, scalable (from 100 g to kg scale), and highly reproducible manner, providing a sufficient supply at a reasonable cost. However, this scale is still significantly behind the scalability of perovskite precursor materials (kg to hundreds-of-kg scale) required for gigawatt-scale solar cell manufacturing, although it could suffice for pilot-line production.308,309
Modern polymer donors such as PBDBT, PM6, and J71 are structurally optimized, necessitating complex chemical structures and lengthy synthetic pathways. As a result, their synthesis is only reproducible at the laboratory scale (milligram to tens-of-gram scale). The use of rare metal palladium catalysts, the high cost of expensive precursors, and the less controllable synthesis due to strong polymer aggregation impede larger-scale synthesis. These polymers tend to precipitate out of reaction mixtures, making homogeneous polymerization, which results in substantial polydispersity of the crude product, extremely challenging. This process requires more time and solvents.
Although certain polymers have been utilized in blade-coated small-area devices instead of modules, there are other polymers available with simpler chemical structures, reduced production costs, and similar performance.310 Furthermore, PTQ10 is notorious for its inconsistent batch-to-batch reproducibility, likely stemming from the non-regiospecific product formed during synthesis, which impacts its morphological and electrical properties.311 Inconsistent molecular weights in polymers like PTAA, and P3HT lead to unpredictable hole mobility.312 Machine learning-assisted polymerization monitoring was used for real-time control.313 Stille/Suzuki coupling—used for N2200 and DPP-DTT—requires toxic Pd catalysts, limiting kg-scale production.314
To increase the scalability of these essential elements, synthetic methods need to be delicately engineered to better regulate reactions. Shin et al.315 reported stepwise heating as a means of improving control over the polymerization reaction, producing the well-known PTB7 polymer with minimal batch-to-batch variation in a highly reproducible manner. Another approach is to create novel polymers whose characteristics are less dependent on their molecular weights. Recent studies suggest that terpolymers or even random polymers can more effectively address this issue, reducing the dependence of solar cell performance on the polymer batch.316 However, this technique also reduces polymer crystallinity, resulting in slightly lower device performance compared to those using more structurally well-defined polymers. This issue could be resolved by gaining a deeper understanding of the relationships between polymer properties and structures, polymer physics in the solution phase, and the fluid dynamics of the precursor ink. Additionally, understanding how polymer self-assembly in solution impacts the morphology of the solar cell absorber is crucial. Defect-responsive smart polymers can be a good roadmap, polymers that dynamically passivate defects under operational stress (e.g., thermally cross-linkable benzocyclobutene).317
In the context of 2025, the techno-economic requirements for perovskite solar modules (PSMs) have become more clearly defined. Recent analyses indicate that rigid perovskite modules with a PCE of around 23% would need to achieve operational lifetimes of approximately 24 years to compete with crystalline silicon photovoltaics, while flexible perovskite modules would require about 17 years of stability at similar efficiency levels.318
From an economic perspective, the National Renewable Energy Laboratory (NREL) has recently assessed the minimum sustainable price (MSP) of perovskite–silicon tandem architectures. Their study suggests that two-terminal tandems could reach MSPs of about 0.428 $ per WDC, while four-terminal designs are close at 0.423 $ per WDC, depending on efficiency and production capacity [2,3].319 In comparison, the current manufacturing cost of single-junction perovskite modules is estimated to be 0.57 $ per W, which remains higher than the benchmark for crystalline silicon. However, sensitivity analyses demonstrate that with further advances, specifically achieving higher than 25% PCE and operational lifetimes of more than 25 years, alongside reductions in material and equipment costs, perovskite modules could surpass silicon in terms of cost competitiveness.320
These updated projections highlight both the progress and the remaining challenges for perovskite solar technology. While current large-area PSMs still lag behind silicon in terms of cost and durability, the combination of improved stability, scalable manufacturing, and tandem designs points toward a viable economic pathway in the near future.
The PCE of a perovskite standard module is now reaching 18.2% at an area of 7200 cm2 while the 24.7%-Si standard module is realized at an area of 17
806 cm2. Besides, according to the calculation by the NREL, the minimum sustainable price (MSP, defined as the price that provides the minimum rate of return necessary in a given industry to support a sustainable business over a long term) with a 15%-gross margin of perovskite solar modulus manufactured at a small scale is calculated to be 0.38 $ per W, higher than that of c-Si (0.25–0.27 $ per W), but it is with possible reductions to 0.21 $ per W for a larger scale if PCE can be improved to 22% for a single-junction PSM without incurring additional costs. Further developing an all-perovskite two-junction tandem module with a PCE of 30% could lower the cost to 0.18 $ per W in the future (Fig. 33).320
![]() | ||
| Fig. 33 The competitive landscape between established silicon solar technology and emerging perovskite solar technology, highlighting current performance and future cost-reduction pathways. | ||
SCPs hold even greater potential in PSCs. Engineering aspects like side chains, backbone, radius of gyration, Kuhn length, amorphous and crystalline regions, and intra- and inter-chain interactions can optimize light absorption, charge separation, and charge collection, alongside morphological control in PSCs. Future studies should investigate the effects of various structures, components, and chain factors in non-conducting and SCPs within different parts of PSCs in a systematic and targeted manner.
The future commercialization of printable solar cells hinges on addressing several critical issues related to production, operation, and scaling up the deployment of organic materials and PSCs. Collaboration among researchers, engineers, and consumers is essential to overcome these challenges. A reliable supply chain for essential materials is crucial for scalable thin-film photovoltaic production. While key precursors for PSCs are readily available, other components, particularly synthetic organics, remain costly. Modern polymers require complex synthesis processes, making large-scale production challenging due to high costs and synthesis difficulties. To improve scalability, synthetic methods must be refined to better control reactions. Approaches like stepwise heating and developing novel polymers less dependent on molecular weights can help, though they may impact polymer crystallinity and device performance. A deeper understanding of polymer properties, structure, and self-assembly in solution is crucial for advancing the field.
| This journal is © The Royal Society of Chemistry 2025 |