Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Multifunctional polymers in perovskite photovoltaics: bridging efficiency, stability, and manufacturing

Farzad Zahedia, Mohsen Ameri*b, Mohammad Hossein Rajabi Manshadic, Mohammad Jafari Pouyac, Marziyeh Mohebbid, Maryam Alidaeie, 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

Received 12th September 2025 , Accepted 21st October 2025

First published on 21st November 2025


Abstract

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.


1 Introduction

Perovskite materials, known for their high absorption coefficients, long carrier diffusion lengths, high carrier mobilities, and bipolar transport characteristics, have significantly enhanced the efficiency of third-generation solar cells. They effectively absorb sunlight over a wide wavelength range, capturing more solar energy. These materials allow charge carriers to travel long distances without recombining, which is crucial for efficient charge extraction and high device performance. The bandgap of perovskite materials can be easily tuned by adjusting the composition of metal halides, and optimizing light absorption and energy conversion efficiency.1–3 The simplicity of perovskite thin film fabrication, compositional engineering, and accessible solution-based fabrication protocols have driven research and quickly boosted efficiency to over 27.5%,4 rivalling traditional silicon-based solar cells.

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.


image file: d5ta07481a-f1.tif
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).


image file: d5ta07481a-f2.tif
Fig. 2 Hierarchical chart summarizing the key themes of this paper, including polymer-assisted defect passivation strategies, the roles of semiconducting polymers, and the challenges and perspectives in large-scale production of hybrid perovskite solar cells.

2 Perovskite defects and passivation

The primary cause of defects in perovskite materials is their tendency for rapid crystallization. Ideally, a crystal structure would have each atom in a fixed location without imperfections or gaps. However, during the formation of films, numerous defects naturally occur within the crystal structure.13 Perovskite films are often composed of polycrystals with inherent grain boundaries, which contribute to defect creation and energy band misalignment. Any deviation in the atomic arrangement of a crystalline solid is considered a defect.14 In PSCs, defects can include traditional solid-state imperfections as well as various types classified as planar (e.g., grain boundaries and stacking faults), bulk (e.g., pores and precipitates), point defects (e.g., vacancies, interstitials, substitutions, antisites, Frenkel, Schottky, and impurity types), and line defects (e.g., dislocations, disclinations, and cracks).15 See Fig. 3 for a summary of the defects.
image file: d5ta07481a-f3.tif
Fig. 3 Defect types in the perovskite lattice.

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


image file: d5ta07481a-f4.tif
Fig. 4 Schematic of defect types in PSCs. (A) The consequences of imperfection effects in the PSCs (B).

2.1 Adverse effects of perovskite defects

Defects and trap states facilitate carrier recombination, especially at interfaces, lowering performance despite improvements in bulk film quality (Fig. 4B). High-quality perovskite films reduce recombination losses, but interfaces remain critical hotspots for charge accumulation and performance loss. Defects and grain boundaries trigger ion movement, causing instability, hysteresis, and phase segregation; passivation is required for long-term stability. Carrier trapping at interfaces causes scan-dependent JV curves and reduced efficiency. Structural strain (e.g., from lattice mismatch or the Goldschmidt tolerance factor) worsens instability, necessitating strain control.24 Intrinsic defects accelerate recombination, ion migration, and environmental sensitivity (moisture, heat, and light). These reduce the open-circuit voltage (VOC), mainly through non-radiative recombination. The VOC is limited by the bandgap (Eg) of the semiconductor in the absorber layer and is further decreased by the recombination of electrons and holes, both through radiative and non-radiative processes.25 Eqn (1) frequently defines the VOC:
 
VOC = VradOC − ΔVnon-radOC (1)
where Vnon-radOC represents the loss in VOC due to non-radiative recombination, and VradOC is the VOC at the radiative limit, where all recombination is radiative. Currently, the individual contributions of radiative and non-radiative activities to the voltage deficit (EgqVOC) remain unknown. However, it is observed that the VOC decreases with increasing non-radiative losses, primarily due to surface defects.26

2.2 Passivation strategies for perovskite defects

Passivation in PSCs is driven by chemical and physical interactions (e.g., charge balance, Lewis acid–base, and steric effects) to neutralize defects. Effective strategies, targeting either bulk or surface/interface traps, are essential for improving device performance and stability. Various strategies are introduced to mitigate the effects of perovskite defects, considering appropriate chemical and physical effects.27–29 Although the mechanisms of many passivation techniques are still not fully understood, an overview of passivation routes is provided in Fig. 5.
image file: d5ta07481a-f5.tif
Fig. 5 General methods for passivation of perovskite photovoltaics. Perovskite passivation strategies are diverse, employing various chemical and structural modifications to neutralize defects that hinder efficiency and stability. 2D passivation involves using large organic cations to create a two-dimensional perovskite layer on the surface of the three-dimensional film. This layer acts as a barrier, passivating surface defects and preventing moisture infiltration. Dual and triple passivation involves the simultaneous use of two or three different passivating agents, respectively, to target a wider range of defect types. These agents often have distinct properties, such as one being a Lewis base with lone-pair electrons that coordinate with undercoordinated Pb2+ ions, while another might be a Lewis acid that accepts electrons. Many effective passivators utilize multiple anchoring groups to create stronger, more stable bonds with the perovskite lattice. Additionally, some molecules are designed with a polar electron density distribution, which helps them to interact more effectively with charged defects and minimize non-radiative recombination.

3 Defect passivation by using polymers

Various passivation agents have been used in the literature including organic molecules, inorganic materials,30,31 zwitterions, ionic salts, and polymers. Polymers are promising passivation agents due to their durability under operation conditions, possible multi-functionality, presence of hydrophobic long chains, the possibility of in situ polymerization inside the layer, good solubility and interference in the film-formation process, ion immobilization ability, etc.32–34

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.

3.1 Bulk passivation by using polymers

As illustrated in Fig. 6A, polymers can be integrated into perovskite materials using either an ex situ or in situ approach.
image file: d5ta07481a-f6.tif
Fig. 6 A summary chart of polymer incorporation methods in perovskites, ex situ (A) and in situ (B).
3.1.1 Ex situ polymer introduction. In the ex situ method, a specific quantity of a polar polymer is directly introduced into the polar solvents that dissolve the perovskite precursors.35 This results in a polymer-precursor solution that enhances coordination and promotes long-range molecular arrangement along the polymer backbone, and even side chains. The underlying mechanism involves Lewis acid–base interactions between the polymer chains and perovskite precursor molecules, facilitating a more structured molecular organization. Furthermore, the presence of polymer chains increases the energy barrier required for nucleation, thereby reducing the likelihood of spontaneous nucleation events. Consequently, fewer nuclei are formed, and their subsequent growth occurs in a more continuous manner without collisions. Upon thermal annealing, the perovskite precursors that are initially bound within the polymeric matrix crystallize after overcoming the increased activation energy barrier for nucleation. The perovskite crystals that form near the extended polymer chains become stabilized by the surrounding polymer structure, promoting the interconnection of large perovskite grains through polymer linkages.36

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

3.1.2 In situ polymer introduction. According to Fig. 6B, incorporating polymers in situ (directly during perovskite formation) or using cross-linkable monomers is a powerful strategy to enhance PSC performance and stability.41 This approach allows polymers to integrate intimately with the perovskite structure, improving film quality, defect passivation, and mechanical robustness. Monomers are introduced into the perovskite ink and polymerized in situ via heat/UV light.42,43 In this method, the monomers or oligomers remain well-dispersed in the perovskite matrix during the solution-processing stage. Upon film formation and post-deposition treatment, these monomers polymerize or cross-link, forming a polymeric network within the perovskite layer. The cross-linking process locks the polymer chains in place, enhancing mechanical stability, and interfacial compatibility. The in situ cross-linking process affects the nucleation and growth of perovskite grains, leading to larger and more uniform grain sizes with fewer defects. The polymer network reduces internal stresses within the perovskite film, preventing microcracks and enhancing mechanical stability. This approach also suppresses ion migration, which, as mentioned, is one of the major degradation pathways in perovskite solar cells.44,45

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.


image file: d5ta07481a-f7.tif
Fig. 7 The most widely used insulating polymers in perovskite structures.

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[double bond, length as m-dash]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.


image file: d5ta07481a-f8.tif
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[double bond, length as m-dash]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[thin space (1/6-em)]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.


image file: d5ta07481a-f9.tif
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.


image file: d5ta07481a-f10.tif
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.


image file: d5ta07481a-f11.tif
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.

Table 1 Photovoltaic performance parameters of some PSCs integrating conventional polymeric components within the photoactive 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−xyPbBrzI3−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−yz: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−xyPbBrzI3−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−xyPbBrzI3−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−xyPbBrzI3−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.


image file: d5ta07481a-f12.tif
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.

3.2 Interface passivation by using polymers

High defect density on the surface of layers causes energetic reactivity, leading to moisture absorption and increased ion migration. Common solution-based coatings for polycrystalline perovskite limit the formation of high-quality layers with large grains.93 Often small grains result in widespread grain boundaries, which are concentrated regions of voids and defects. Therefore, a crucial aspect of perovskite passivation focuses on morphology control and surface treatment. When the perovskite layer interfaces with CTLs, defect issues worsen, and recombination rates increase. Interface engineering is essential for providing an ideal substrate for subsequent layers, especially for perovskite crystallization. Additionally, lattice mismatch strain between layers occurs at the contacts. Perovskite films made through rapid low-temperature methods often have many interfacial defects. Compared to bulk defects, interfacial defects can be more harmful to device performance due to their much higher densities, leading to significant non-radiative recombination losses. These interfacial defects accelerate the degradation of PSCs and must be addressed to achieve industrial-scale production. Passivation approaches are categorized into top and buried interfacial methods based on their spatial positioning.94,95
3.2.1 Top interface approach. The top interface passivation approach is a strategy used to reduce defects at the interface between the perovskite layer and the overlying CTLs in perovskite solar cells. These defects can be caused by the release of volatile organic components during the thermal annealing process, which can lead to deficiencies in the coordination of lead or halide vacancies on the perovskite surface. To address this issue, various methods have been developed, including using halide salts to induce recrystallization or employing low-dimensional perovskite passivation layers to modify the perovskite surface.96

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[triple bond, length as m-dash]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.

3.2.2 Buried interface approach. Defects that accumulate at the buried interface cause non-radiative recombination, a process where charge carriers are lost as heat instead of being converted into electricity. This leads to a significant loss in the device's VOC and overall PCE. The abundance of trap sites on the surfaces of CTLs can negatively affect perovskite crystal growth, interfacial charge transport, and overall device stability. Defects arise from several issues. Mismatches in the thermal expansion coefficients between the perovskite and the underlying layer can create stress and defects. Mismatched electrical properties can lead to poor contact, causing charge carriers to accumulate and recombine instead of being efficiently extracted. The rapid and sometimes non-uniform crystallization of the perovskite precursor can lead to a high density of defects at the interface. To optimize the buried interface, volatile organic additives are pre-buried in the charge transport layer, and multifunctional hybrid interfacial layers have been introduced. These methods effectively reduce trap densities at the buried interfaces, optimize interfacial charge transport, and significantly enhance device stability.98 However, the buried interface is difficult to modify because it is formed during the deposition of the perovskite film, making it inaccessible for post-treatment.

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).


image file: d5ta07481a-f13.tif
Fig. 13 The illustration of the buried interface fastening strategy. (A) A schematic of the strategy, where a polyurethane layer formed in situ fastens the SnO2 electron transport layer to the perovskite film above it via hydrogen bonds. (B) Details of the chemical reaction mechanism between the two additives, hexamethylene diisocyanate (HDI) and ethylene glycol (EG), which form this polyurethane bridge. (C) The FTIR spectroscopy data as proof, confirming the reaction occurred by showing the new N–H and C[double bond, length as m-dash]O bonds in the EG:HDI mixture. (D) A schematic showing how the HDI additive regulates the perovskite crystallization process, inhibiting complex intermediate phases and promoting oriented crystal growth for a higher-quality film.

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[double bond, length as m-dash]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

Table 2 A comparison summary of the top and buried interface approach
  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 IV 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.


image file: d5ta07481a-f14.tif
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[thin space (1/6-em)]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[double bond, length as m-dash]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.


image file: d5ta07481a-f15.tif
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.


image file: d5ta07481a-f16.tif
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.

Table 3 Photovoltaic performance parameters of some polymer-assisted electron transportation in perovskite–polymer hybrid solar cells
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


Table 4 Photovoltaic performance parameters of some polymer-assisted hole transportation in perovskite–polymer hybrid solar cells
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−xz/PVA/spiro-OMeTAD/Au 1.149 25.24 80.0 23.20 128
11 PEG ITO/SnO2/FAyMA1−yPbClxBrzI3−xz/PEG/spiro-OMeTAD/Au 1.126 24.94 79.6 22.35 128
12 PVK ITO/SnO2/FAyMA1−yPbClxBrzI3−xz/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−xyPbBrzI3−z/p-NP-E/Au 1.130 24.5 78.3 21.7 135
21 PBDB-O ITO/SnO2/CsxFAyMA1−xyPbBrzI3−z/PBDB-O/MoO3/Ag 1.072 16.98 56.24 10.23 136
22 PBDB-T ITO/SnO2/CsxFAyMA1−xyPbBrzI3−z/PBDB-T/MoO3/Ag 1.122 22.75 74.70 19.07 136
23 PBDB-Cz ITO/SnO2/CsxFAyMA1−xyPbBrzI3−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



image file: d5ta07481a-f17.tif
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.

4 Semiconducting polymers (SCPs)

SCPs or conjugated polymers are organic materials with π-electron delocalization along their backbone, enabling electrical conductivity while maintaining processability and flexibility. They bridge the gap between plastics and inorganic semiconductors.

4.1 Types of SCPs

SCPs are usually classified on two bases.
4.1.1 By structure.
4.1.1.1 Linear conjugated polymers. They are a class of SCPs with a continuous, straight-chain backbone of alternating single and double bonds (π-conjugation). This conjugation allows delocalized π-electrons, enabling charge transport along the polymer chain. Some linear conjugated polymers are listed in Table 5.
Table 5 Some linear conjugated polymers152–154
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) 2.7–3.2 104–10−1 (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



4.1.1.2 Hyperbranched/dendritic conjugated polymers. They are highly branched, 3D macromolecules with unique structural and functional properties compared to linear polymers. Table 6 shows some important polymers in this category.
Table 6 Some main hyperbranched/dendritic conjugated polymers155,156
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



4.1.1.3 Donor–acceptor copolymers. They are a class of conjugated polymers engineered by alternating electron-rich (donor, D) and electron-deficient (acceptor, A) units along their backbone. This design enhances intramolecular charge transfer, enabling tunable electronic and optical properties for advanced organic electronics. Donor units (D) are typically aromatic or heterocyclic (e.g., thiophene and carbazole). Acceptor units (A) are electron-withdrawing groups (e.g., benzothiadiazole and diketopyrrolopyrrole (DPP)). Alternating the D–A–D–A sequence creates a “push–pull” effect, lowering the bandgap. The HOMO (donor) and LUMO (acceptor) levels can be adjusted by modifying D/A units, which enables absorption/emission across visible to near-infrared (NIR) spectra. Some key donor–acceptor (D–A) copolymers are listed in Table 7.
Table 7 Key donor–acceptor (D–A) copolymers157,158
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


4.1.2 By function.
4.1.2.1 p-Type (hole transport). They are organic semiconductors that primarily transport holes when doped or in a device. They are characterized by high-lying HOMO levels (typically −4.5 to −5.5 eV), facilitating hole injection. Table 8 shows some p-type conjugated polymers and their HOMO energies, and the chart in Fig. 18 illustrated some p-type conjugated polymers and their properties.
Table 8 Some p-type conjugated polymers and their HOMO levels
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



image file: d5ta07481a-f18.tif
Fig. 18 Some important p-type conjugated polymers used in solar cells and their properties.159–165

4.1.2.2 n-Type (electron transport). They are organic semiconductors that primarily transport electrons. They are characterized by low-lying LUMO levels (typically −3.5 to −4.5 eV), facilitating electron injection, electron-deficient (acceptor) units in their backbone, and high electron affinity for stable charge transport. The chart in Fig. 19 illustrates some n-type conjugated polymers and their properties, and Table 9 shows some n-type conjugated polymers and their LUMO energies.
image file: d5ta07481a-f19.tif
Fig. 19 Some important n-type conjugated polymers used in solar cells and their properties.166–170
Table 9 Some n-type conjugated polymers and their LUMO levels
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



4.1.2.3 Ambipolar conjugated polymers. They are organic semiconductors that can transport both electrons and holes with comparable efficiency. These materials bridge the gap between p-type and n-type polymers, enabling complementary logic circuits and more efficient optoelectronic devices. Some of their main features include balanced charge transport so that they have similar hole (µh) and electron (µe) mobilities (typically 10−3–100 cm2 V−1 s−1) and LUMO levels and HOMO levels are placed around −3.5 to −4.2 eV and around −5.0 to −5.8 eV, respectively. Table 10 lists some ambipolar conjugated polymers and their hole and electron mobilities.
Table 10 Some ambipolar conjugated polymers and their hole and electron mobilities171–173
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


4.2 SCP role in perovskite hybrid solar cells

SCPs, with the potential to set energy bands, have unique electrical and optical features.174 Therefore, they can passivate defects and also increase the carrier extraction, transportation, and light harvesting ability. The feasibility of extraction and transportation of carrier charges besides the suitable energy band levels and regulation of grain boundaries and size via highly delocalized conjugated polymers reduces the energy losses significantly. In other words, SCPs can reduce energy losses in PSCs via four mechanisms. Improved charge transport reduces energy losses associated with charge recombination. Morphological control to enhance the crystallinity of perovskite films and reduce defect density subsequently lowers the recombination rate. Interfacial engineering to passivate the interfaces subsequently reduces trap-assisted recombination and improves charge extraction. And photon management can improve light absorption and minimize thermalization losses.175 On the other hand, conducting polymers mostly are soluble in non-polar solvents and these solvents play the role of anti-solvents of a perovskite precursor. So, two mechanisms of rapid nucleation via anti-solvent pipetting and the slowing down of crystal growth via the presence of polymers in the precursor solution occur at the same time.

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.


image file: d5ta07481a-f20.tif
Fig. 20 Advantages of HTM SCPs relative to traditional HTMs.

image file: d5ta07481a-f21.tif
Fig. 21 Advantages of ETM SCPs relative to traditional ETMs.
4.2.1 The backbone structures of SCPs. The backbone structures of SCPs used in perovskite solar cells can be categorized based on their chemical composition and the arrangement of their molecular units. Conjugated polythiophenes such as P3HT (poly(3-hexylthiophene)) have a backbone composed of thiophene rings. They are known for their good charge transport properties and are widely used in organic electronics. Polyphenylene derivatives include polymers like PPP (poly(p-phenylene)) and its derivatives, which have a backbone of benzene rings. They offer high thermal stability and good electronic properties.181 Diketopyrrolopyrrole (DPP)-based polymers contain DPP units in their backbone and are known for their high charge carrier mobility and stability.182 Naphthalene diimide (NDI)-based polymers have a backbone that includes naphthalene diimide units, providing high electron mobility and good environmental stability.183 Helicene-based polymers as a recent addition to this category exhibit remarkable film-forming and hole transport properties, as well as exceptional thermal durability.184,185 Each of these categories has unique properties that make them suitable for different applications within PSCs, such as active layers, or CTLs. The choice of polymer is crucial for the overall performance and stability of the solar cell.
4.2.2 The side chains of SCPs. Side chains in SCPs play a multifaceted role in PSCs.186,187 They can influence the polymer's solubility, film-forming ability, and the alignment of energy levels with the perovskite layer. Side chains can also impact the morphology of the active layer, charge carrier mobility, and overall device performance. For instance, side chains such as vinyloxy, allyloxy, and propargyloxy have been shown to improve hole mobility and match energy levels with the perovskite layer, which is essential for efficient device operation. Longer chains improve solubility but may reduce crystallinity. Branched chains prevent excessive aggregation. Additionally, side chains can be engineered to enhance the interfacial adhesion between layers, reduce charge recombination, and improve the environmental stability of the solar cells.188 SCP side chains used in PSCs can be categorized based on their chemical structure and function. Alkyl side chains are the most common side chains used to improve the solubility and processability of polymers; examples include P3HT.189 Aromatic side chains can enhance the electronic properties of polymers. For instance, phenyl-carbazole groups can be used for this purpose.190 Redox-active side chains can participate in charge transfer processes. They can improve the reproducibility and stability of perovskite solar cells.191–193 Functional side chains are designed to interact with the perovskite layer or to facilitate charge transport. For example, side chains terminated with vinyloxy, allyloxy, and propargyloxy have been shown to improve hole mobility.194 Conformational side chains, and twisted or planar side chains can influence the packing and the charge transport properties of the polymer. Twisted triphenylamine (TPA) groups, for example, can favour interfacial charge transfer.195 Each type of side chain can have a significant impact on the performance and stability of PSCs by influencing the polymer's physical and electronic interactions with the perovskite layer. Researchers continue to explore and optimize these side chains to enhance the efficiency and longevity of solar cells. Tables 11–16 categorize some important types of side chains of SCPs applicable in solar cells.
Table 11 Conjugated polymers include alkyl side chains196–199
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


Table 12 Conjugated polymers include ether/alkoxy side chains
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)


Table 13 Conjugated polymers include fluorinated side chains200,201
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


Table 14 Conjugated polymers include siloxane side chains202
Side chain type Example polymers Impact on performance
Polydimethylsiloxane (PDMS) PDMS-grafted DPP polymers Improves mechanical flexibility


Table 15 Conjugated polymers include ionic side chains203–205
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


Table 16 Conjugated polymers include hybrid side chains206,207
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.


image file: d5ta07481a-f22.tif
Fig. 22 The structures of some SCPs have been researched further.
4.2.3 The radius of gyration and Kuhn length of SCPs. The radius of gyration (Rg) and Kuhn length (Lk) are important parameters that describe the physical properties of SCPs, which can influence the performance of PSCs.208,209 The Rg measures the distribution of the constituent atoms around the center of mass of a polymer molecule. In PSCs, a larger Rg typically indicates a more extended polymer chain, which can lead to better overlap between the electronic orbitals of adjacent chains. This improved orbital overlap can enhance charge transport properties, which is beneficial for the efficiency of solar cells.210 The Lk is a measure of the stiffness of a polymer chain, representing the length of a hypothetical segment of the chain that behaves as a rigid rod. An Lk implies a stiffer chain, which can affect the polymer's ability to form ordered structures. In the context of PSCs, polymers with longer Kuhn lengths may form more crystalline domains, leading to improved charge mobility and, consequently, better device performance.211 The rigidity of SCPs in PSCs is crucial for maintaining structural integrity and facilitating efficient charge transport. A rigid polymer backbone can help to minimize the reorganization energy, which is beneficial for charge transfer processes.212 For example, as mentioned a helicene-based SCP with a high Tg demonstrated exceptional mechanical properties, contributing to the thermal durability and operational stability of PSCs. In summary, the radius of gyration and Kuhn length of SCPs can significantly impact the morphology, charge transport, and overall efficiency of PSCs.
4.2.4 The crystallinity of SCPs. The amorphous and crystalline regions of SCPs play crucial roles in the performance of PSCs. The crystalline regions are typically associated with higher charge mobility due to their ordered structure, which can enhance the efficiency of charge transport within the solar cell. On the other hand, the amorphous regions can contribute to the flexibility and processability of the polymer, which is beneficial for the fabrication of solar cells. In the context of PSCs, the presence of both amorphous and crystalline regions can affect the optical and electronic properties of the active layer.213 For instance, a study focusing on P3HT embedded with nitrogen/sulfur-doped graphene quantum dots (NS-GQDs) demonstrated that the degree of crystallinity influences the enhancement in electronic charge transfer for both the amorphous and crystalline regions.214 This suggests that controlling the ratio and distribution of these regions within the SCP can be key to optimizing the performance of PSCs. Moreover, the interaction between the polymer and the perovskite material can be affected by the degree of crystallinity, which in turn can influence the overall PCE of the solar cell. Therefore, understanding and manipulating the effects of amorphous and crystalline regions in SCPs is essential for the development of high-performance perovskite solar cells.215
4.2.5 H- and J-aggregation. Intra- and inter-chain interactions of SCP nanowires (J- and H-aggregate, respectively) play a significant role in the performance of PSCs (Fig. 23). Intra-chain interactions are the interactions within a single polymer chain. They affect the polymer's ability to transport charge along its length. Efficient intra-chain interactions facilitate the movement of charge carriers, such as electrons and holes, which is crucial for the high electrical conductivity required in PSCs. Inter-chain interactions occur between adjacent polymer chains. They are vital for creating pathways that allow charge carriers to move between chains. Good inter-chain stacking can lead to enhanced charge mobility across the polymer network, contributing to the overall efficiency of the solar cell.189 Both intra- and inter-chain interactions are influenced by the molecular structure of the polymer, such as the rigidity of the backbone and the nature of the side chains.216,217 For instance, a study on naphthalene-diimide-based copolymers showed that a synergistic effect between intra-chain and inter-chain interactions gives rise to high electron mobilities,218 which is desirable for PSCs. This indicates that neither intra-chain nor inter-chain interactions alone are sufficient; rather, their combination forms a robust network that supports efficient charge transport. Furthermore, the morphology of SCP nanowires, which is dictated by these interactions, can impact the light absorption, charge separation, and charge collection processes within PSCs.189 Nanowires offer a direct pathway for charge carriers, which can reduce the distance that electrons and/or holes need to travel, thus minimizing recombination losses. This direct path can lead to improved charge mobility and higher PCEs.
image file: d5ta07481a-f23.tif
Fig. 23 Schematic of the J-type and H-type interactions in the SCP chains.

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

4.2.6 Hot injection carriers and the role of SCPs. Hot electron/hole injection in solar cells refers to the process where high-energy charge carriers are generated by absorbing photons with energy greater than the bandgap of the solar cell material. Instead of losing this excess energy as heat, these hot carriers are quickly extracted before they can cool down, which can potentially increase the efficiency of the solar cell beyond the conventional limits (Fig. 24). The concept is a part of the efforts to surpass the Shockley–Queisser limit, which is the theoretical maximum efficiency of a single p–n junction solar cell under standard illumination. Hot carrier solar cells aim to utilize the excess kinetic energy of the charge carriers to achieve higher power conversion efficiencies.220–222 However, there are challenges in realizing hot carrier solar cells, such as the need for materials that can effectively extract hot carriers and the requirement to minimize energy loss through carrier relaxation. Research in this area is ongoing, with various approaches being explored to optimize the extraction of hot carriers and integrate them into practical solar cell designs. SCPs can address the challenge of hot carriers in perovskite solar cells by providing pathways for efficient charge transfer. A study has shown that hot electrons can be injected into the low-lying LUMOs of certain SCPs.223 This process can potentially lead to a loss pathway for hot electrons, which is significant because the final photocurrents of perovskite solar cells are affected by the suppression of this loss pathway. These findings suggest that by carefully selecting and engineering SCPs, it's possible to optimize the charge transport layers in perovskite solar cells, thereby improving their efficiency and stability while addressing the hot carrier challenge.
image file: d5ta07481a-f24.tif
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
4.2.7 Hybrid polymer–perovskite bulk heterojunctions. The interpenetrating network (IPN) structure of Bulk Hetero Junction (BHJ) donor/acceptor SCPs and also intertwining into the perovskite layer offers several benefits for perovskite solar cells.224 By employing a mixture of electron donor and electron acceptor materials the perovskite/BHJ structure can be formed.225,226 The interconnected pathways within an IPN facilitate efficient charge transport, which is crucial for the high performance of PSCs. The IPN structure can enhance the separation of electron–hole pairs, reducing the probability of recombination before they are extracted, which is a common source of energy loss. The large interface area provided by an IPN can lead to better contact between the polymer and the perovskite layer, enhancing charge extraction and reducing recombination losses. The large interface area in an IPN can improve the contact between the active layer and charge transport layers, which minimizes charge recombination at these interfaces, a critical factor in energy loss. IPNs can help stabilize the perovskite layer against environmental factors such as moisture and oxygen, which can degrade the material and lead to increased non-radiative recombination and energy loss.

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.


image file: d5ta07481a-f25.tif
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.


image file: d5ta07481a-f26.tif
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.

4.2.8 SCPs as CTLs. SCPs play a crucial role as CTLs in PSCs, where they primarily function to extract and transport charges while minimizing recombination losses. Their role can be categorized as extracting electrons from the perovskite, blocking opposite charge carriers to reduce charge recombination, improving interfacial contact with the cathode or anode, passivating perovskite surface defects, improving energy-level alignment between perovskite and the electrode, and enhancing stability by protecting the perovskite from moisture/oxygen.230,231

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.


image file: d5ta07481a-f27.tif
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[double bond, length as m-dash]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.


image file: d5ta07481a-f28.tif
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
4.2.9 Free dopant SCPs. The difference between dopant-free and doped SCPs in perovskite solar cells lies in their electrical properties and impact on device stability.66 Doped SCPs require the addition of dopants to enhance their electrical conductivity and PCE. Dopants can introduce instability as they may attract moisture, which can degrade the perovskite layer. Dopant-free SCPs are inherently conductive without the need for additional dopants.234,235 Dopant-free polymers can still achieve impressive efficiencies and are considered more stable, making them attractive for long-term applications. Overall, dopant-free materials are gaining attention for their potential to provide a balance between efficiency and stability in PSCs. Some researches focus on dopant-free polymeric HTMs due to high carrier mobilities, matching energy levels with perovskite, homogeneous film formation, and excellent hydrophobicity which are attractive options to replace conventional materials.

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.


image file: d5ta07481a-f29.tif
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.

5 Large scale production of hybrid PSCs

Large-scale production of hybrid PSCs involves several methods developed for scalability, efficiency, and stability.

5.1 Spin coating

Initially used as a laboratory technique, it can be adapted for larger production using automated systems to ensure uniform film formation.

5.2 Slot-die coating

Slot-die coating238–240 is a roll-to-roll (R2R) compatible deposition technique used to fabricate thin, uniform films for solar cells. It involves precisely dispensing a liquid ink (e.g., perovskite precursor solution) through a slit-shaped die onto a moving substrate, enabling high-throughput, large-area production with minimal material waste. Slot-die head contains a microfluidic channel to control ink flow. The pump system delivers ink at a controlled flow rate (µL min−1 to mL min−1). The moving substrate is typically glass, plastic, or metal foil (for flexible PSCs). The drying system includes inline heaters or IR lamps for solvent evaporation. Table 17 shows some key challenges and solutions of this method.
Table 17 Challenges and solutions of slot-die coating method
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


5.3 Vapor-assisted deposition

Vapour-assisted deposition (VAD)241,242 is a two-step, solvent-free method to fabricate high-quality perovskite films. This method is compatible with R2R and vacuum processes. Unlike solution processing, it involves depositing a solid metal halide (e.g., PbI2 or PbBr2) layer via spin-coating, evaporation, or sputtering. The layer exposed to organic halide vapor (e.g., MAI, FAI, or CsI) in a controlled environment diffuses into the solid film and reacts to form the perovskite crystal. This method is prized for precise stoichiometry control, reduced defects, and scalability. VAD bridges the gap between lab-scale precision and industrial scalability, offering a path to stable, high-efficiency PSCs. While slower than solution methods, its defect-suppressing capabilities make it ideal for premium applications (e.g., space PV and tandems). Hybrid approaches, and combining VAD with inkjet printing for patterned deposition are some of the future directions. Lead-free perovskites adapt VAD for Sn2+ or Bi3+-based films. Table 18 shows some important challenges and solutions of this method.
Table 18 Challenges and solutions of the vapour-assisted deposition method
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


5.4 Spray coating

Spray coating243,244 is a solution-based deposition technique where a perovskite precursor ink is atomized into fine droplets and sprayed onto a substrate using pressurized gas (e.g., air and nitrogen). The droplets dry on contact, forming a thin perovskite film. This method is low-cost, scalable, and compatible with large-area or flexible substrates, making it attractive for industrial production. Perovskite precursors (e.g., PbI2 + MAI/FAI) dissolved in solvents (DMF, DMSO, or mixtures). Additives (e.g., MACl, polymer binders) may be included to improve film quality. The ink is forced through a nozzle, breaking into fine droplets (∼1–50 µm diameter). Ultrasonic, pneumatic, or electrospray methods can be used. Droplets land on a heated substrate (50–120 °C), where the solvent evaporates. Multiple passes may be needed for uniform thickness. Annealing (100–150 °C) was used to crystallize the perovskite. Optional anti-solvent quenching was used to enhance grain growth. Several future research directions include high-speed robotic automated spray systems for R2R production and hybrid methods combining blade coating or inkjet printing. Some key challenges and solutions of this method are presented in Table 19.
Table 19 Challenges and solutions of the spray coating method
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


5.5 Inkjet printing

Inkjet printing245,246 is an additive, non-contact manufacturing technique that deposits functional inks (including perovskite precursors) in precise patterns by ejecting tiny droplets (10–100 picolitres) through microscopic nozzles. This method enables digital patterning (mask-less deposition), high material utilization (>95% efficiency), and low-temperature processing (compatible with flexible substrates). The print head contains 50–1000 nozzles (typically piezoelectric). The ink reservoir holds perovskite precursor solution. A high-precision XY-motion stage is used for accurate substrate alignment. The drying system utilizes IR lamps or a heated stage to remove solvents.

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.

Table 20 Challenges and solutions of the inkjet printing method
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


5.6 Blade coating

Blade coating (also called doctor blading)247,248 is a solution-based deposition technique where a sharp blade spreads a liquid perovskite precursor ink across a substrate to form a thin, uniform film. This method bridges the gap between lab-scale spin coating and industrial R2R production, offering high material utilization (>90% efficiency), large-area compatibility (unlimited substrate width), and precise thickness control (50–500 nm range). The process begins with ink deposition of a perovskite precursor (e.g., PbI2 + MAI in DMF/DMSO, often with additives like MACl or 2D seeds), followed by blade-induced shear flow that forms a liquid meniscus, which dries into a film across three zones—upstream (wet ink reservoir), meniscus (evaporation/crystallization), and downstream (dried film). Post-treatment includes solvent quenching (e.g., chlorobenzene) and thermal annealing (100–150 °C) to finalize the layer. Some critical parameters of this method are listed in Table 21. In Table 22, several significant challenges and the solutions related to this method are outlined.
Table 21 Critical parameters of the blade coating method
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


Table 22 Challenges and solutions of the blade coating method
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


5.7 Flexographic printing

Flexographic printing or flexo-printing249,250 is a R2R compatible technique that uses a flexible relief plate to transfer ink onto substrates. Key features include direct patterning via raised image areas on a photopolymer plate, anilox roller controls ink volume with microscopic cells, and fast deposition with µm-level resolution. Common substrates include flexible PET, PEN, ITO-coated glass, or metal foil. The substrate undergoes plasma treatment, UV-ozone exposure, or solvent cleaning to enhance adhesion and remove contaminants. Surface treatments (e.g., self-assembled monolayers) may be applied to improve wettability and charge transport. Perovskite ink includes organic and inorganic salts dissolved in solvents like DMF/DMSO. Additives (e.g., polymers and surfactants) are used for stability. Viscosity, surface tension, and drying behavior are optimized to ensure uniform ink transfer. An anilox roller (engraved with micro-cells) picks up a controlled amount of ink from an ink reservoir. A doctor blade removes excess ink, ensuring precise volume transfer. A flexographic plate (made of an elastomer or photopolymer) transfers the ink onto the moving substrate. Pressure, speed, and temperature are controlled for uniform deposition. Controlled heating (infrared or hot air) ensures proper drying without phase separation. Annealing is typically done at 100–150 °C to improve perovskite crystallization and polymer integration. Fig. 30 shows a schematic of the flexo-printing setup. Table 23 lists the current challenges and solutions of this method.
image file: d5ta07481a-f30.tif
Fig. 30 Schematic picture of the flexographic setup. A flexographic printing setup for PSCs is a roll-to-roll (R2R) system designed for high-speed, patterned deposition of liquid inks onto flexible substrates. Its core structure revolves around the precise interaction of several rotating cylinders. The process begins with an ink fountain that supplies the perovskite precursor ink. A rotating anilox roller, which is a hard cylinder laser-engraved with a microscopic cell pattern, picks up a controlled and metered volume of ink from the fountain. This anilox roller then transfers the ink to the raised features of a soft, patterned printing plate, which is mounted onto the plate cylinder. As the flexible substrate (such as PET or PEN coated with transparent electrodes) passes between this plate cylinder and a central impression cylinder, the ink is pressed from the printing plate onto the substrate, forming the desired patterned thin film. This entire “inking—metering—transfer—printing” sequence happens in a continuous, rapid motion, allowing for the fast and scalable fabrication of perovskite solar cell layers.
Table 23 Challenges and solutions of the flexo printing method
Challenge Innovation
Crystallization control Hybrid thermal/vapor annealing stations
Edge definition Laser-etched plates (5 µm precision)
Layer registration CCD camera alignment systems


5.8 Hybrid chemical vapor deposition (HCVD)

Hybrid Chemical Vapor Deposition (HCVD)251,252 is a thin-film deposition method used to fabricate high-quality perovskite layers with excellent crystallinity, uniformity, and scalability. It combines aspects of solution processing and vapor-phase deposition, making it an attractive method for large-area PSC fabrication. A precursor solution containing lead halide salts is spin-coated or blade-coated onto the substrate. This partially dried precursor layer acts as a reactive medium for the subsequent vapor-phase step. The substrate is placed inside a heated vacuum chamber. Organic halides (e.g., MAI and FAI) are vaporized at controlled temperatures (∼150–200 °C) and diffuse into the PbI2 film. The reaction forms a high-purity perovskite layer. The film undergoes post-annealing (∼100–150 °C) to enhance crystallinity and grain growth. This step also helps in removing residual solvents and improving film stability. HCVD combines solution processing (for metal halides) and vapor-phase deposition (for organic components) to grow high-quality perovskite films. The main challenges and solutions are listed in Table 24. The schematic of the HCVD printing setup is shown in Fig. 31.
Table 24 Challenges and solutions of the HCVD method
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)



image file: d5ta07481a-f31.tif
Fig. 31 Schematic picture of the HCVD setup. This machine is an integrated system that merges vapor-phase and solution-phase processing to create high-quality, uniform perovskite films, especially over large areas. Its structure typically consists of several key zones arranged in a linear or in-line sequence. The process begins in the precursor evaporation zone. Here, solid precursor powders (e.g., lead iodide, PbI2, and organic halides like methylammonium iodide, MAI) are loaded into separate, temperature-controlled crucibles or boats. These crucibles are heated to precise temperatures to sublimate the solids, converting them directly into a vapor without passing through a liquid phase. These precursor vapors are then transported by an inert carrier gas (such as nitrogen or argon) into the main deposition chamber. This chamber contains the substrate, which is typically pre-coated with other solar cell layers (like the electron transport layer) and is often heated on a temperature-controlled stage. It is within this chamber that the “hybrid” aspect often takes place: while the PbI2 is delivered via vapor, a separate solution spray nozzle can simultaneously mist a solution of the organic halide dissolved in a solvent. This allows for precise control over the reaction stoichiometry. As the vapor-phase PbI2 and the solution-phase organic halide meet at the heated substrate surface, they react to form the perovskite film (e.g., MAPbI3). The chamber is often equipped with vacuum pumps to maintain a controlled, low-pressure environment, which enhances vapor diffusion and leads to a more uniform film with large crystal grains and fewer defects. Finally, the substrate may pass through a post-annealing zone to crystallize the film fully before exiting the system. This combined vapor-solution approach allows for superior control over film morphology and coverage compared to purely solution-based methods.

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.

Table 25 Key differences between VAD and HCVD methods
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.

Table 26 Perovskite deposition method comparison table253–255
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


Table 27 Industrialization metrics of deposition methods253–255
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


5.9 Benefits of printing methods

The large-scale printing of solar cells presents numerous advantages essential for the progress and adoption of solar energy technology. By printing solar cells on a larger scale, the cost per unit can be significantly reduced, making solar energy more affordable and competitive with conventional energy sources.240,256 The high flexibility of printed solar cells allows them to be applied to various surfaces and objects, including buildings, vehicles, and clothing, broadening the possibilities for solar energy generation. Large-scale printing processes can optimize material use, reduce waste, and improve the environmental footprint of solar cell production. This inherently scalable printing process facilitates a swift transition from laboratory-scale prototypes to industrial-scale production, meeting the growing demand for renewable energy sources.257 The lightweight and adaptable nature of printed solar cells supports innovative deployment methods, such as integration into portable devices or temporary installations, which traditional rigid panels cannot achieve. These benefits underscore the transformative potential of large-scale printing technology in the solar industry, paving the way for more sustainable and widespread solar energy use.

5.10 Challenges and defects

When printing solar cells on a large scale, several defects may arise that can impact the performance and longevity of the panels.258 Material wastage is a significant amount of solution that can be wasted during the printing process, making it unsuitable for large-scale production. Achieving precise patterning is challenging and can hinder module fabrication, particularly when moving from small-scale to large-scale production. Film defects are another important challenge. The thin active layer (100–600 nm) of large-area solar cells is prone to defects that can cause leakage current, reducing production yields.259 Grain size and grain boundary defects, along with ion migration in the lattice and the compactness of the interface buffer layer, significantly affect the stability of perovskite solar cells.260 Transitioning from spin-coating to printing as the production scale increases is challenging and essential for fabricating large-area perovskite solar cells. Addressing these defects is crucial for improving the yield and performance of large-scale printed solar cells. Ongoing research and development are focused on refining printing processes and materials to minimize these defects.
5.10.1 Defect types in film formation due to printing. The printing of PSCs can lead to various film defects, which can significantly impact the performance and stability of the cells.
5.10.1.1 Point defects. These are atomic-level defects, including vacancies, interstitials, and substitutions in the perovskite lattice.261,262 In large-scale printing, inhomogeneous precursor distribution or solvent variations can introduce point defects, leading to non-radiative recombination, which lowers efficiency. Vacancy defects or missing atoms are categorized as iodine vacancies, which are most common due to weak Pb–I bonds. Lead vacancies are less frequent but highly detrimental. Cation vacancies mean that MA+ or FA+ is missing from the lattice. Interstitial defects or extra atoms are created because of interstitial halides or extra I ions between lattice sites and interstitial Pb that can introduce deep trap states. Anti-site defects or wrong atoms in a site cause severe non-radiative recombination in the case of Pb at the I site. Also, the case of I at the Pb site is rare but disruptive. Substitutional defects or impurity atoms act as dopants; for example replacing Br with I can passivate defects but may distort the lattice.

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


5.10.1.2 Non-uniform film thickness. Large-area printing methods can suffer from poor solution spreading, resulting in uneven thickness. Thicker regions may cause excess absorption and higher recombination, while thinner areas lead to low light absorption and shunting paths.265,266
5.10.1.3 Grain boundary (GB) defects. In printed large-area solar cells, GBs are particularly problematic because they introduce defects, recombination centres, and ion migration pathways, significantly impacting efficiency and stability.267–269 Fast solvent evaporation in printing techniques (e.g., inkjet and spray coating) leads to smaller grains and more GBs. Poor control over nucleation leads to random grain orientations, and likewise high defect density. In large-area printing, temperature gradients cause uneven crystallization. Some regions dry faster, leading to stress-induced cracks along GBs. Residual PbI2 or organic halides segregate at GBs, creating electrically inactive regions. If polymers do not uniformly distribute, they can accumulate at GBs, disrupting charge transport. These methods are used to tackle the destructive effects of GBs. Additive engineering (e.g., MACl and NH4Cl) promotes Ostwald ripening for larger grains. Hot casting slows crystallization, yielding micrometer-sized grains. Lewis base additives bind to undercoordinated Pb2+ at GBs. Halide compensation (excess MAI/FAI) fills iodine vacancies at GBs. In 2D/3D perovskite hybrids, 2D layers cover GBs, improving stability. Some polymers with high film formation ability form a protective mesh around grains, reducing ion migration. Solvent vapor annealing heals GBs by partial re-dissolution and recrystallization.270 Light soaking reduces trap density via photo-induced defect healing.264 Ink rheology control ensures uniform nucleation and pre-patterning substrates guide grain alignment.
5.10.1.4 Pinholes and voids. Pinholes (nanoscale pores) and voids (larger empty regions) can form due to rapid solvent evaporation, poor wetting, or insufficient precursor coverage.271 Unlike lab-scale spin-coated films, large-area printed PSCs are more prone to these defects due to process-related challenges. Fast-evaporating solvents (e.g., DMF and acetone) cause uneven drying, leading to film shrinkage cracks and voids. Slow solvents (e.g., DMSO and GBL) can induce coffee-ring effects, leaving pinholes. In the case of ink aggregation, perovskite precursor particles may clump together, creating gaps in the film. If the ink doesn't spread uniformly on the substrate, de-wetting occurs, leaving voids. Techniques like slot-die coating or inkjet printing dry unevenly, causing localized pinholes. Air bubbles trapped during ink dispensing burst, forming voids. If the perovskite doesn't bond well with the underlying layer (e.g., TiO2 and PEDOT:PSS), voids form. If crystallization occurs too fast (e.g., in ambient air printing), pinholes appear where grains don't merge properly. Solvent escaping too quickly can leave behind voids. Strategies to minimize pinholes and voids are as follows. Mixed solvents improve wetting and slow drying. Polymer binders fill gaps and improve film continuity. Surfactants (e.g., PEG and L-α-phosphatidylcholine) reduce surface tension for better spreading.239 Vacuum-assisted deposition removes trapped air bubbles before film formation. Plasma/UV-ozone cleaning removes organic contaminants for better ink adhesion. Pre-heating substrates helps in controlled crystallization. Exposing films to solvent vapors (e.g., DMF and NMP) re-dissolves and recrystallizes perovskite, sealing pinholes.
5.10.1.5 Surface roughness. Surface roughness refers to the uneven topography of perovskite films at the nanoscale to microscale level.272,273 In printed solar cells, excessive roughness creates interfacial problems that degrade performance and reproducibility, especially in large-area manufacturing. Performance impacts of roughness are presented in Table 28.
Table 28 The effect of roughness types on the fundamental properties of perovskite solar cells and modules
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


5.10.1.6 Thermal instability. Some perovskite materials are sensitive to thermal stress during processing, which can cause defects or degradation. Thermal instability is a critical failure mode in printed PSCs, causing performance degradation and structural changes when exposed to temperature variations during manufacturing or operation.280–282 Unlike single-crystal silicon solar cells that withstand >200 °C, perovskite devices degrade at much lower temperatures (typically 90–150 °C), creating unique challenges for printed large-area modules. Degradation pathways and their impacts based on temperature range are presented in Table 29. Some mitigation strategies are as follows. Cs+/Rb+ alloying suppresses phase segregation. 2D perovskite armor at grain boundaries leads to a good improvement. Thermal-stable polymers such as PTAA, and P3HT instead of spiro-OMeTAD are effective. Chloride incorporation (5–10%) delays ion migration.283 Other items include UV-curable inks (<70 °C curing), and cold sintering techniques.
Table 29 Degradation pathways and their impacts
Temperature range Primary degradation mode Efficiency loss mechanism
50–70 °C Ion migration begins Increased JV 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



5.10.1.7 Coffee ring effect. Among these defects, the coffee ring effect is a common issue in the printing of PSCs.284,285 This effect occurs when solute particles in a liquid are transported to and deposited at the edges of a drying droplet, forming a ring-like pattern similar to a coffee stain. The mechanism of coffee ring formation is as follows. Capillary flow: evaporation at the droplet edge creates an outward fluid flow to replenish lost solvents. Contact line pinning: the droplet perimeter remains fixed while the solvent evaporates. Marangoni effect (Ma): surface tension gradients can partially counteract outward flow (but often fail in perovskite inks). This significantly affects the uniformity and quality of printed films, especially in droplet-based deposition methods like inkjet or spray-coating. In PSCs, this effect can lead to non-uniform film thickness, poor morphology of the perovskite structure, and ultimately lower device performance. It can result in reduced internal quantum efficiency and increased charge recombination rate. Overcoming this defect is crucial for enhancing the efficiency and stability of printed PSCs. Polymers can play a vital role in eliminating the coffee ring effect by increasing the ink's viscosity, which reduces capillary flow and leads to more uniform film formation. Additionally, polymers can adjust the solvent evaporation rate, helping to form a uniform solute distribution as the ink dries, and can also modify the surface tension of the ink, influencing its wetting properties and flow behavior on the substrate, thus reducing the likelihood of coffee ring formation.

5.11 What is the role of polymers in large-scale production?

Polymers play a pivotal role in addressing the defects that arise during the large-scale printing of PSCs. Polymers can improve the quality of the perovskite film, leading to fewer defects such as pinholes or grain boundary issues. Polymers can act as a template or scaffold, guiding the crystallization of the perovskite material. This controlled growth helps to prevent the formation of defects that can arise from uncontrolled, rapid crystallization. Polymers can be used to passivate the surface of the perovskite layer. This means they fill in any existing pinholes or voids and also neutralize charge traps at the grain boundaries and surfaces of the perovskite crystals. This reduces non-radiative recombination and improves charge extraction. Polymers can effectively mitigate defects in large-scale films of PSCs by addressing issues like the coffee-ring effect, improving crystal quality, and enhancing lattice stability. The introduction of an in situ polymer framework, such as poly(ethylene glycol) diacrylamide (PEGDAM), helps to modulate the Ma convection intensity within the precursor droplet.286 The polymer framework increases the precursor's viscosity, which in turn reduces the Ma number, a value that represents the ratio of surface tension to viscous forces. A lower Ma value leads to a decrease in the intensity of Ma convection and weakens the microfluidic flocculation motion inside the droplet during solvent evaporation. The polymer framework slows down thermal capillary flow and homogenizes the heat transfer process across the film's surface, which contributes to regulating the coffee-ring effect.286 The 3D network formed by the spontaneous cross-linking of the polymer at room temperature can anchor the flocculent movement of perovskite colloidal particles, counteracting the lateral capillary flow that causes the coffee-ring effect. The polymer framework can retard the crystallization rate of the perovskite precursor. It interacts with the perovskite components to form a complex, altering the nucleation energy barrier. This leads to the formation of a large number of nucleation sites in a short period, allowing sufficient time for subsequent crystal growth.286 Strong interactions between the polymer and the perovskite in the wet-film state can restrain the generation of multiple intermediate phases and promote oriented crystal growth. Polymers remaining at the grain boundaries can passivate deep trap states, which facilitates improvements in both the PCE and the long-term stability of the perovskite devices. Residual stress in perovskite films can lead to lattice distortions and negatively impact the photovoltaic performance. The in situ polymer framework helps to release tensile stress.286

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.


image file: d5ta07481a-f32.tif
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.

6 Challenges and perspectives

In summary, we present our perspectives on the future commercialization of printable and flexible hybrid PSCs. To expedite the development of printable thin-film photovoltaics for practical applications, several critical issues must be addressed. These include production, operation, and the challenges of scaling up the production and deployment of PSCs. Addressing these challenges requires collaboration among researchers, engineers, and consumers. To realize the envisioned terawatt-scale deployment of solar panels, it is imperative to establish a dependable and robust supply chain for essential materials, ensuring the scalable production of thin-film photovoltaics. While key precursors for PSCs, such as lead iodide and formamidinium salts, are readily accessible, other critical components, particularly synthetic organics like NFAs or PTAA, remain relatively costly.

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[thin space (1/6-em)]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


image file: d5ta07481a-f33.tif
Fig. 33 The competitive landscape between established silicon solar technology and emerging perovskite solar technology, highlighting current performance and future cost-reduction pathways.

7 Conclusion

Polymers, as a champion material, can significantly enhance the performance and stability of PSCs. The architecture of polymer passivation agents and their interaction with perovskite crystals are crucial factors. Features such as 3D structures and multiple functional groups with controllable configurations improve PSC performance. Selecting the most effective functional groups for the polymeric structure is essential. Polymers influence the kinetics of perovskite crystals during both nucleation and growth stages. Consequently, higher-quality hybrid PSCs exhibit more uniform film morphology, leading to lower series and transfer resistance, higher recombination resistance, and improved charge transfer and extraction. Additionally, denser perovskite films with larger grains enhance vertical alignment, facilitating carrier transport. Certain moieties, such as benzene rings, and carbonyl, ester, amide, amine, thiophene, and carboxyl groups, effectively coordinate Pb2+ and anchor I ions.

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.

Conflicts of interest

The authors declare no conflict of interest.

Data availability

This manuscript is a review article that compiles, interprets, and discusses findings from previously published, peer-reviewed studies in reputable scientific journals. All data, results, and figures presented here are drawn directly from these original sources. Permissions have been secured for the reproduction of figures and any other materials, and full citations and acknowledgments are provided to appropriately credit the original authors and publications. No new primary data were produced or analyzed for this work. Additional details regarding sources or permissions for reproduced content can be obtained from the corresponding author upon request.

Acknowledgements

The authors gratefully acknowledge Dr Masoud Alahbakhshi for his valuable guidance, insightful discussions, and constructive feedback throughout the course of this research. His expertise and critical input significantly contributed to the development and refinement of this work. The authors wish to acknowledge the support of the Henry Royce Institute for Advanced Materials through the Industrial Collaboration Programme (RICP-R4-100061) and MATcelerateZero (MATZ0), funded from a grant provided by the Engineering and Physical Sciences Research Council EP/X527257/1. The authors acknowledge the Department for Energy Security and Net Zero (Project ID: NEXTCCUS), University College London's Research, Innovation and Global Engagement, University of Sydney – University College London Partnership Collaboration Awards, UCL-Peking University Strategic Partner Funds, Cornell-UCL Global Strategic Collaboration Awards and UCL – Korea University Strategic Partner Funds for their financial support. The authors acknowledge the ACT program (Accelerating CCS Technologies, Horizon 2020 Project No. 691712) for the financial support of the NEXTCCUS project (project ID: 327327).

References

  1. Z. Huang, Z. Du, Z. Ma, Y. Zheng, Q. Zhang, W. You, J. Yang, T. Yu, S. Hou and G. Li, Chem. Eng. J., 2024, 499, 156684 CrossRef CAS.
  2. M. Knezevic, T.-H. Hoang, N. Rashid, M. Abdi-Jalebi, C. Colbeau-Justin and M. N. Ghazzal, Sci. China Mater., 2023, 66, 2545–2572 CrossRef CAS.
  3. T. Liu, R. A. Scheidt, X. Zheng, S. Joy, Q. Jiang, H. R. Atapattu, M. Chen, H. Pruett, K. Zhu and J. M. Luther, Cell Rep. Phys. Sci., 2023, 4, 1–17 Search PubMed.
  4. M. Li, H. Gao, L. Li, E. Wang, Z. Liu, I. T. Cheong, P. Wu, Y. Zhang, Y. Wang and X. Zheng, Nat. Photonics, 2025, 1–9 Search PubMed.
  5. Y. Guo, H. Tan and B. Xu, Energy Environ. Sci., 2025, 18, 8744–8755 RSC.
  6. Y. Jiang, Z. Zhang, Y. Wang, J. Jin, Y. Huang, W. Wang, D. Ma, H. Huang, C. Chen and S. Ren, Adv. Funct. Mater., 2025, 2504541 CrossRef CAS.
  7. M. Hu, Y. Zhu, Z. Zhou, M. Hao, C. Harnmanasvate, J. Waiyawat, Y. Wang, J. Lu, Q. An and X. Li, Adv. Energy Mater., 2023, 13, 2301888 CrossRef CAS.
  8. W. Cheng, R. Zhou, S. Peng, C. Wang and L. Chen, Mater. Today Commun., 2023, 107879 Search PubMed.
  9. S. Hou, Z. Ma, Y. Li, Z. Du, Y. Chen, J. Yang, W. You, Q. Yang, T. Yu and Z. Huang, Adv. Funct. Mater., 2024, 34, 2310133 CrossRef CAS.
  10. G. S. H. Thien, K.-Y. Chan, M. A. M. Sarjidan, W. H. Abd Majid and B. K. Yap, Appl. Surf. Sci. Adv., 2024, 19, 100538 CrossRef.
  11. M. Hu, Y. Zhu, Z. Zhou, M. Hao, C. Harnmanasvate, J. Waiyawat, Y. Wang, J. Lu, Q. An and X. Li, Adv. Energy Mater., 2023, 13, 2301888 CrossRef CAS.
  12. J. Yi, G. Zhang, H. Yu and H. Yan, Nat. Rev. Mater., 2024, 9, 46–62 CrossRef CAS.
  13. E. Akman, T. Ozturk, W. Xiang, F. Sadegh, D. Prochowicz, M. M. Tavakoli, P. Yadav, M. Yilmaz and S. Akin, Energy Environ. Sci., 2023, 16, 372–403 RSC.
  14. Z. Fang, J. Sun, S. F. Liu and L. Ding, J. Semicond., 2023, 44, 080201 CrossRef CAS.
  15. J. Ye, M. M. Byranvand, C. O. Martínez, R. L. Z. Hoye, M. Saliba and L. Polavarapu, Angew. Chem., 2021, 133, 21804–21828 CrossRef.
  16. A. Dey, P. Rathod and D. Kabra, Adv. Opt. Mater., 2018, 6, 1800109 CrossRef.
  17. Q. Cai, Q. Tan, J. He, S. Tang, Q. Sun, D. He, T. Cheng, G. Ma, J. Huang and G. Su, Joule, 2025, 101880 CrossRef CAS.
  18. Y. Shi, W. Chu, L. Zhang, B. Wang, W. A. Saidi, J. Zhao and O. V Prezhdo, Chem. Mater., 2025, 37, 655–664 CrossRef CAS.
  19. G. Liu, M. Ghasemi, Q. Wei, B. Jia, Y. Yang and X. Wen, Adv. Energy Mater., 2025, 2405239 CrossRef CAS.
  20. M. S. Hammer, H. Schlott, L. Lüer, C. J. Brabec, M. Sytnyk, J. Will, B. Meyer and W. Heiss, Nat. Rev. Mater., 2025, 1–15 CAS.
  21. X. Zhang, M. E. Turiansky, J.-X. Shen and C. G. Van de Walle, Phys. Rev. B, 2020, 101(14), 140101 CrossRef CAS.
  22. I. du Fossé, J. T. Mulder, G. Almeida, A. G. M. Spruit, I. Infante, F. C. Grozema and A. J. Houtepen, J. Am. Chem. Soc., 2022, 144, 11059–11063 CrossRef.
  23. Z. Guo, M. Yuan, G. Chen, F. Liu, R. Lu and W. Yin, Adv. Sci., 2024, 2305799 CrossRef CAS.
  24. Y. Lei, Y. Xu, M. Wang, G. Zhu and Z. Jin, Small, 2021, 17, 2005495 CrossRef CAS.
  25. G. J. W. Aalbers, T. P. A. van der Pol, K. Datta, W. H. M. Remmerswaal, M. M. Wienk and R. A. J. Janssen, Nat. Commun., 2024, 15, 1276 CrossRef CAS PubMed.
  26. J. Xia, C. Liang, H. Gu, S. Mei, S. Li, N. Zhang, S. Chen, Y. Cai and G. Xing, Energy Environ. Mater., 2023, 6, e12296 CrossRef CAS.
  27. J. Wang, L. Bi, Q. Fu and A. K. Jen, Adv. Energy Mater., 2024, 14, 2401414 CrossRef CAS.
  28. Y. Zhang, M. Abdi-Jalebi, B. W. Larson and F. Zhang, Adv. Mater., 2024, 36, 2404517 CrossRef CAS.
  29. Y. Chen, S. Tan, N. Li, B. Huang, X. Niu, L. Li, M. Sun, Y. Zhang, X. Zhang and C. Zhu, Joule, 2020, 4, 1961–1976 CrossRef CAS.
  30. H. Cheng, X. Zang, S. Wang and B. Cai, Sol. RRL, 2025, 9, 2400736 CrossRef CAS.
  31. W. Qi, X. Zhou, J. Li, J. Cheng, Y. Li, M. J. Ko, Y. Zhao and X. Zhang, Sci. Bull., 2020, 65, 2022–2032 CrossRef CAS PubMed.
  32. H. Guo, G. W. Yoon, Z. J. Li, Y. Yun, S. Lee, Y. Seo, N. J. Jeon, G. S. Han and H. S. Jung, Adv. Energy Mater., 2024, 14, 2302743 CrossRef CAS.
  33. J. Qin, Y. Chen, X. Guo, H. Yang, Y. Qin, C. Yao, B. Lv, Y. Bai and Y. Zhan, Adv. Opt. Mater., 2023, 11, 2202809 CrossRef CAS.
  34. P. Tang, M. Ameri, Q. Liang, P. I. Udenze, Z. Ouyang, M. Brown, L. Smith, I. Fedin and D. Li, ACS Appl. Mater. Interfaces, 2025, 17, 31614–31626 CrossRef CAS.
  35. K. Chandratre, P. Sharma, S. Thomas, E. Parvazian, R. Patidar, R. Garcia-Rodriguez, E. Pean, R. Suthar, T. O. Dunlop and S. K. Pathak, ACS Appl. Energy Mater., 2024, 7, 5595–5607 CrossRef CAS.
  36. J. Xi, G. Zhao, X. Sun, S. Zhang, X. Du, J. Li, H. Yang, R. Wang, Z. Lu and G. Zou, ChemNanoMat, 2025, 11, e202400478 CrossRef CAS.
  37. Y. Gou, S. Tang, C. Yuan, P. Zhao, J. Chen and H. Yu, Mater. Horiz., 2024, 11, 3465–3481 RSC.
  38. Z. Wu, S. Sang, J. Zheng, Q. Gao, B. Huang, F. Li, K. Sun and S. Chen, Angew. Chem., Int. Ed., 2024, 63, e202319170 CrossRef CAS PubMed.
  39. F. Liu, J. Dou, W. Zhang, Y. Li, Y. Mei, Q. Guo, Q. Chen, Y. Zhao, Y. Wang and X. Zhang, ACS Energy Lett., 2025, 10, 4091–4101 CrossRef CAS.
  40. L. Zuo, H. Guo, D. W. deQuilettes, S. Jariwala, N. De Marco, S. Dong, R. DeBlock, D. S. Ginger, B. Dunn and M. Wang, Sci. Adv., 2017, 3, e1700106 CrossRef.
  41. J. Ma, S. Fan, C. Shao, L. Wang, Y. Dong, G. Niu, Z. Nie, S. Yang, J. Wang and H. Yang, Angew. Chem., 2025, 137, e202425578 CrossRef.
  42. M. Zhang, B. He, B. Yang, Z. Yang, H. Sui, H. Chen, J. Duan and Q. Tang, Chem. Eng. J., 2025, 165948 CrossRef CAS.
  43. T. Li, P. Zhang, S. Wei, Y. Jing, J. Shi, Y. Chen, H. Zhong and G. Yang, Adv. Opt. Mater., 2024, 12, 2400486 CrossRef CAS.
  44. X. Huang, B. Wu and N. Zheng, Acc. Mater. Res., 2024, 6, 40–51 CrossRef.
  45. S. Li, X. Xu, X. Wang, N. Huang, J. Fang, D. Lin, Y. Shao, J. Zhou, A. K. K. Kyaw and S. He, Angew. Chem., Int. Ed., 2025, 64, e202421174 CrossRef CAS PubMed.
  46. Y. Yang, N. Yang, C. Zhang, S. Li, Y. Hao, Q. Sun, Z. Chen, S. Liu, F. Guo and J. Lu, Adv. Sci., 2025, e08253 CrossRef CAS.
  47. Z. Zhu, K. Mao, K. Zhang, W. Peng, J. Zhang, H. Meng, S. Cheng, T. Li, H. Lin and Q. Chen, Joule, 2022, 6, 2849–2868 CrossRef CAS.
  48. M. Li, R. Sun, J. Chang, J. Dong, Q. Tian, H. Wang, Z. Li, P. Yang, H. Shi and C. Yang, Nat. Commun., 2023, 14, 573 CrossRef CAS.
  49. Y. Wu, G. Xu, J. Xi, Y. Shen, X. Wu, X. Tang, J. Ding, H. Yang, Q. Cheng and Z. Chen, Joule, 2023, 7, 398–415 CrossRef CAS.
  50. C. Zhao, H. Zhang, M. Almalki, J. Xu, A. Krishna, F. T. Eickemeyer, J. Gao, Y. M. Wu, S. M. Zakeeruddin and J. Chu, Adv. Mater., 2023, 35, 2211619 CrossRef CAS PubMed.
  51. Z. Huang, X. Hu, C. Liu, X. Meng, Z. Huang, J. Yang, X. Duan, J. Long, Z. Zhao and L. Tan, Adv. Funct. Mater., 2019, 29, 1902629 CrossRef.
  52. D. Wang, L. Zhang, K. Deng, W. Zhang, J. Song, J. Wu and Z. Lan, Energy Technol., 2018, 6, 2380–2386 CrossRef CAS.
  53. K. Ma, H. R. Atapattu, Q. Zhao, Y. Gao, B. P. Finkenauer, K. Wang, K. Chen, S. M. Park, A. H. Coffey and C. Zhu, Adv. Mater., 2021, 33, 2100791 CrossRef CAS PubMed.
  54. Y. Zou, W. Yu, Z. Tang, X. Li, H. Guo, G. Liu, Q. Zhang, Y. Zhang, Z. Zhang and C. Wu, Appl. Phys. Lett., 2021, 119(15), 151104 CrossRef CAS.
  55. K. Suwa, K. Oyaizu, H. Segawa and H. Nishide, ChemSusChem, 2019, 12, 5207–5212 CrossRef CAS PubMed.
  56. T. Wang, Y. Li, Q. Cao, J. Yang, B. Yang, X. Pu, Y. Zhang, J. Zhao, Y. Zhang and H. Chen, Energy Environ. Sci., 2022, 15, 4414–4424 RSC.
  57. M. Li, R. Sun, J. Chang, J. Dong, Q. Tian, H. Wang, Z. Li, P. Yang, H. Shi and C. Yang, Nat. Commun., 2023, 14, 573 CrossRef CAS PubMed.
  58. Y. Zhao, J. Wei, H. Li, Y. Yan, W. Zhou, D. Yu and Q. Zhao, Nat. Commun., 2016, 7, 10228 CrossRef CAS PubMed.
  59. C. Zhao, H. Zhang, M. Almalki, J. Xu, A. Krishna, F. T. Eickemeyer, J. Gao, Y. M. Wu, S. M. Zakeeruddin and J. Chu, Adv. Mater., 2023, 35, 2211619 CrossRef CAS PubMed.
  60. Y. Shen, G. Xu, J. Li, X. Lin, F. Yang, H. Yang, W. Chen, Y. Wu, X. Wu and Q. Cheng, Angew. Chem., 2023, 135, e202300690 CrossRef.
  61. S. Lin, S. Wu, D. Guo, H. Huang, X. Zhou, D. Zhang, K. Zhou, W. Zhang, Y. Hu and Y. Gao, Small Methods, 2023, 7, 2201663 CrossRef CAS.
  62. S. Valero, T. Soria, N. Marinova and J. L. Delgado, Polym. Chem., 2019, 10, 5726–5736 RSC.
  63. Q. Dong, Z. Wang, K. Zhang, H. Yu, P. Huang, X. Liu, Y. Zhou, N. Chen and B. Song, Nanoscale, 2016, 8, 5552–5558 RSC.
  64. C.-C. Zhang, M. Li, Z.-K. Wang, Y.-R. Jiang, H.-R. Liu, Y.-G. Yang, X.-Y. Gao and H. Ma, J. Mater. Chem. A, 2017, 5, 2572–2579 RSC.
  65. A. Fakharuddin, M. Seybold, A. Agresti, S. Pescetelli, F. Matteocci, M. I. Haider, S. T. Birkhold, H. Hu, R. Giridharagopal and M. Sultan, ACS Appl. Mater. Interfaces, 2018, 10, 42542–42551 CrossRef CAS.
  66. J. Jiang, Q. Wang, Z. Jin, X. Zhang, J. Lei, H. Bin, Z. Zhang, Y. Li and S. Liu, Adv. Energy Mater., 2018, 8, 1701757 CrossRef.
  67. T.-H. Han, J.-W. Lee, C. Choi, S. Tan, C. Lee, Y. Zhao, Z. Dai, N. De Marco, S.-J. Lee and S.-H. Bae, Nat. Commun., 2019, 10, 520 CrossRef CAS PubMed.
  68. H. Zheng, X. Xu, S. Xu, G. Liu, S. Chen, X. Zhang, T. Chen and X. Pan, J. Mater. Chem. C, 2019, 7, 4441–4448 RSC.
  69. S. Collavini, M. Saliba, W. R. Tress, P. J. Holzhey, S. F. Völker, K. Domanski, S. H. Turren-Cruz, A. Ummadisingu, S. M. Zakeeruddin and A. Hagfeldt, ChemSusChem, 2018, 11, 1032–1039 CrossRef CAS PubMed.
  70. B. Zhang, C. Chen, X. Wang, X. Du, D. Liu, X. Sun, Z. Li, L. Hao, C. Gao and Y. Li, Angew. Chem., Int. Ed., 2023, 62, e202213478 CrossRef CAS PubMed.
  71. B. Yu, L. Zhang, J. Wu, K. Liu, H. Wu, J. Shi, Y. Luo, D. Li, Z. Bo and Q. Meng, J. Mater. Chem. A, 2020, 8, 1417–1424 RSC.
  72. Q. Cao, T. Wang, J. Yang, Y. Zhang, Y. Li, X. Pu, J. Zhao, H. Chen, X. Li and I. Tojiboyev, Adv. Funct. Mater., 2022, 32, 2201036 CrossRef CAS.
  73. T. Zheng, Q. Zhou, T. Yang, Y. Zhao, B. Fan, J. Bo, L. Fan and R. Peng, Carbon, 2022, 196, 213–219 CrossRef CAS.
  74. C. S. Pathak, G. Paramasivam, F. Mathies, K. Hirselandt, V. Schröder, O. Maus, J. Dagar, C. Klimm, E. Unger and I. Visoly-Fisher, ACS Appl. Energy Mater., 2022, 5, 4085–4095 CrossRef CAS.
  75. W. Chen, Y. Wang, G. Pang, C. W. Koh, A. B. Djurišić, Y. Wu, B. Tu, F. Liu, R. Chen and H. Y. Woo, Adv. Funct. Mater., 2019, 29, 1808855 CrossRef.
  76. N. Chen, X. Yi, J. Zhuang, Y. Wei, Y. Zhang, F. Wang, S. Cao, C. Li and J. Wang, Nano-Micro Lett., 2020, 12, 1–13 CrossRef PubMed.
  77. R. Garai, R. K. Gupta, A. S. Tanwar, M. Hossain and P. K. Iyer, Chem. Mater., 2021, 33, 5709–5717 CrossRef CAS.
  78. F. Li, J. Yuan, X. Ling, Y. Zhang, Y. Yang, S. H. Cheung, C. H. Y. Ho, X. Gao and W. Ma, Adv. Funct. Mater., 2018, 28, 1706377 CrossRef.
  79. Y. Xu, G. Liu, J. Hu, G. Wang, M. Chen, Y. Chen, M. Li, H. Zhang and Y. Chen, J. Phys. Chem. Lett., 2022, 13, 3754–3762 CrossRef CAS PubMed.
  80. K. Kim, J. Han, S. Lee, S. Kim, J. Choi, J. Nam, D. Kim, I. Chung, T. Kim and S. Manzhos, Adv. Energy Mater., 2023, 13, 2203742 CrossRef CAS.
  81. X. Meng, Z. Xing, X. Hu, Z. Huang, T. Hu, L. Tan, F. Li and Y. Chen, Angew. Chem., Int. Ed., 2020, 59, 16602–16608 CrossRef CAS.
  82. T. Han, Y. Zhao, J. Yoon, J. Y. Woo, E. Cho, W. D. Kim, C. Lee, J. Lee, J. Choi and J. Han, Adv. Funct. Mater., 2022, 32, 2207142 CrossRef CAS.
  83. C. Zhang, Z. Wang, S. Yuan, R. Wang, M. Li, M. F. Jimoh, L. Liao and Y. Yang, Adv. Mater., 2019, 31, 1902222 CrossRef.
  84. N. Jiang, B. Xing, Y. Wang, H. Zhang, D. Yin, Y. Liu, Y. Bi, L. Zhang, J. Feng and H. Sun, Sci. Bull., 2022, 67, 794–802 CrossRef CAS.
  85. Y. Zhao, P. Zhu, M. Wang, S. Huang, Z. Zhao, S. Tan, T. Han, J. Lee, T. Huang and R. Wang, Adv. Mater., 2020, 32, 1907769 CrossRef CAS PubMed.
  86. G. Li, Z. Su, L. Canil, D. Hughes, M. H. Aldamasy, J. Dagar, S. Trofimov, L. Wang, W. Zuo and J. J. Jerónimo-Rendon, Science, 2023, 379, 399–403 CrossRef CAS.
  87. B. P. Finkenauer, Y. Gao, X. Wang, Y. Tian, Z. Wei, C. Zhu, D. J. Rokke, L. Jin, L. Meng and Y. Yang, Cell Rep. Phys. Sci., 2021, 2(2), 100320 CrossRef CAS.
  88. Q. Cao, J. Yang, T. Wang, Y. Li, X. Pu, J. Zhao, Y. Zhang, H. Zhou, X. Li and X. Li, Energy Environ. Sci., 2021, 14, 5406–5415 RSC.
  89. Y. Lan, Y. Wang and Y. Song, Flexible Printed Electron., 2020, 5, 014001 CrossRef CAS.
  90. L. Wang, G. Liu, X. Xi, G. Yang, L. Hu, B. Zhu, Y. He, Y. Liu, H. Qian and S. Zhang, Crystals, 2022, 12, 894 CrossRef CAS.
  91. Z. Huang, X. Hu, C. Liu, L. Tan and Y. Chen, Adv. Funct. Mater., 2017, 27, 1703061 CrossRef.
  92. X. Duan, X. Li, L. Tan, Z. Huang, J. Yang, G. Liu, Z. Lin and Y. Chen, Adv. Mater., 2020, 32, 2000617 CrossRef CAS PubMed.
  93. M. Hu, Y. Zhu, Z. Zhou, M. Hao, C. Harnmanasvate, J. Waiyawat, Y. Wang, J. Lu, Q. An and X. Li, Adv. Energy Mater., 2023, 13, 2301888 CrossRef CAS.
  94. C. Liu, K. Huang, B. Hu, Y. Li, L. Zhang, X. Zhou, Y. Liu, Z. Liu, Y. Sheng and S. Chen, Adv. Funct. Mater., 2023, 33, 2212698 CrossRef CAS.
  95. L. Shen, P. Song, L. Zheng, L. Wang, X. Zhang, K. Liu, Y. Liang, W. Tian, Y. Luo and J. Qiu, Adv. Mater., 2023, 35, 2301624 CrossRef CAS PubMed.
  96. H. Huang, X. Zhang, W. Zhou, Y. Huang, Z. Zheng, X. Chen, Y. Zhang and H. Yan, Mater. Chem. Front., 2024, 8, 3528–3557 RSC.
  97. D. Liu, X. Wang, X. Wang, B. Zhang, X. Sun, Z. Li, Z. Shao, S. Mao, L. Wang and G. Cui, Angew. Chem., Int. Ed., 2023, 62, e202301574 CrossRef CAS PubMed.
  98. N. Yan, Z. Fang, Z. Dai, J. Feng and S. Liu, Adv. Funct. Mater., 2024, 34, 2314039 CrossRef CAS.
  99. Z. Cai, Z. Yao, Z. Xing, R. Dai, Z. Huang, X. Meng, X. Hu and Y. Chen, Adv. Funct. Mater., 2025, 2505921 CrossRef CAS.
  100. X. Zhao, R. Luo, C. Yu, X. Xu, W. Zhu, Y. Min and N. Cai, Adv. Funct. Mater., 2024, 34, 2307568 CrossRef CAS.
  101. H. Wang, Y. Zheng, G. Zhang, P. Wang, X. Sui, H. Yuan, Y. Shi, G. Zhang, G. Ding and Y. Li, Adv. Mater., 2024, 36, 2307855 CrossRef CAS.
  102. G. D. Tabi, D.-T. Nguyen, W. Liang, W. Ji, T. Lu, T. Trần-Phú, O. L. C. Lem, A. O. Mayon, K. Huang and L.-C. Chang, Chem. Eng. J., 2024, 482, 149062 CrossRef CAS.
  103. B. Zhang, C. Chen, X. Wang, X. Du, D. Liu, X. Sun, Z. Li, L. Hao, C. Gao and Y. Li, Angew. Chem., Int. Ed., 2023, 62, e202213478 CrossRef CAS PubMed.
  104. X. Tang, C. Yang, Y. Xu, J. Xia, B. Li, M. Li, Y. Zhou, L. Jiang, H. Liu and K. Ma, Nat. Photonics, 2025, 1–8 Search PubMed.
  105. Z. Li, C. Jia, Z. Wan, J. Xue, J. Cao, M. Zhang, C. Li, J. Shen, C. Zhang and Z. Li, Nat. Commun., 2023, 14, 6451 CrossRef CAS PubMed.
  106. H. Chen, H. Lin, Z. Sun, H. Li, C. He and D. Mao, J. Mater. Chem. A, 2023, 11, 2443–2451 RSC.
  107. Z. Li, C. Jia, Z. Wan, J. Cao, J. Shi, J. Xue, X. Liu, H. Wu, C. Xiao and C. Li, Nat. Commun., 2025, 16, 1771 CrossRef CAS PubMed.
  108. S. You, H. Zeng, Z. Ku, X. Wang, Z. Wang, Y. Rong, Y. Zhao, X. Zheng, L. Luo and L. Li, Adv. Mater., 2020, 32, 2003990 CrossRef CAS.
  109. F. Qin, W. Meng, J. Fan, C. Ge, B. Luo, R. Ge, L. Hu, F. Jiang, T. Liu and Y. Jiang, ACS Appl. Mater. Interfaces, 2017, 9, 26045–26051 CrossRef CAS.
  110. J. Wei, F. Guo, X. Wang, K. Xu, M. Lei, Y. Liang, Y. Zhao and D. Xu, Adv. Mater., 2018, 30, 1805153 CrossRef.
  111. Z. Xiong, L. Lan, Y. Wang, C. Lu, S. Qin, S. Chen, L. Zhou, C. Zhu, S. Li and L. Meng, ACS Energy Lett., 2021, 6, 3824–3830 CrossRef CAS.
  112. J. Duan, Y. Liu, X. Chen, S. Huang, W. Ou-Yang, G. Zhu, S. Zhang and Z. Sun, Org. Electron., 2020, 82, 105737 CrossRef CAS.
  113. J. Peng, J. I. Khan, W. Liu, E. Ugur, T. Duong, Y. Wu, H. Shen, K. Wang, H. Dang and E. Aydin, Adv. Energy Mater., 2018, 8, 1801208 CrossRef.
  114. D. Li, Y. Huang, R. Ma, H. Liu, Q. Liang, Y. Han, Z. Ren, K. Liu, P. W. Fong and Z. Zhang, Adv. Energy Mater., 2023, 13, 2204247 CrossRef CAS.
  115. D. Yang, X. Zhang, K. Wang, C. Wu, R. Yang, Y. Hou, Y. Jiang, S. Liu and S. Priya, Nano Lett., 2019, 19, 3313–3320 CrossRef CAS PubMed.
  116. J. Peng, D. Walter, Y. Ren, M. Tebyetekerwa, Y. Wu, T. Duong, Q. Lin, J. Li, T. Lu and M. A. Mahmud, Science, 2021, 371, 390–395 CrossRef CAS PubMed.
  117. J. Tian, Q. Xue, X. Tang, Y. Chen, N. Li, Z. Hu, T. Shi, X. Wang, F. Huang and C. J. Brabec, Adv. Mater., 2019, 31, 1901152 CrossRef.
  118. Y. Yue, T. Umeyama, Y. Kohara, H. Kashio, M. Itoh, S. Ito, E. Sivaniah and H. Imahori, J. Phys. Chem. C, 2015, 119, 22847–22854 CrossRef CAS.
  119. M. Jiang, Q. Niu, X. Tang, H. Zhang, H. Xu, W. Huang, J. Yao, B. Yan and R. Xia, Polymers, 2019, 11, 147 CrossRef.
  120. J. Zhou, X. Tian, R. Chen, W. Chen, X. Meng, X. Guan, J. Wang, S. Liu, F. Ren and S. Zhang, Chem. Eng. J., 2023, 461, 141914 CrossRef CAS.
  121. Z. Zhu, Q. Xue, H. He, K. Jiang, Z. Hu, Y. Bai, T. Zhang, S. Xiao, K. Gundogdu and B. R. Gautam, Adv. Sci., 2015, 3, 1500353 CrossRef PubMed.
  122. F. Tan, H. Tan, M. I. Saidaminov, M. Wei, M. Liu, A. Mei, P. Li, B. Zhang, C. Tan and X. Gong, Adv. Mater., 2019, 31, 1807435 CrossRef.
  123. E. Akman and S. Akin, Adv. Mater., 2021, 33, 2006087 CrossRef CAS PubMed.
  124. A. N. Mikheeva, I. E. Kuznetsov, M. M. Tepliakova, A. Elakshar, M. V Gapanovich, Y. G. Gladush, E. O. Perepelitsina, M. E. Sideltsev, A. F. Akhkiamova and A. A. Piryazev, Molecules, 2022, 27, 8333 CrossRef CAS PubMed.
  125. Q. Fu, H. Liu, X. Tang, R. Wang, M. Chen and Y. Liu, ACS Energy Lett., 2022, 7, 1128–1136 CrossRef CAS.
  126. H. Hwang, S. Park, J. H. Heo, W. Kim, H. Ahn, T.-S. Kim, S. H. Im and H. J. Son, J. Power Sources, 2019, 418, 167–175 CrossRef CAS.
  127. J. Ali, J. Song, Y. Li, K. Qian, Q. Zhang, G. Zhou, M. Zhang, L. Zhu, C. Qiu and J. Xu, J. Mater. Chem. C, 2019, 7, 11932–11942 RSC.
  128. M. Wang, Y. Zhao, X. Jiang, Y. Yin, I. Yavuz, P. Zhu, A. Zhang, G. S. Han, H. S. Jung and Y. Zhou, Joule, 2022, 6, 1032–1048 CrossRef CAS.
  129. F. Zhang, J. Song, R. Hu, Y. Xiang, J. He, Y. Hao, J. Lian, B. Zhang, P. Zeng and J. Qu, Small, 2018, 14, 1704007 CrossRef.
  130. M. Li, X. Yan, Z. Kang, Y. Huan, Y. Li, R. Zhang and Y. Zhang, ACS Appl. Mater. Interfaces, 2018, 10, 18787–18795 CrossRef CAS PubMed.
  131. B. Li, Y. Xiang, K. D. G. I. Jayawardena, D. Luo, J. F. Watts, S. Hinder, H. Li, V. Ferguson, H. Luo and R. Zhu, Sol. RRL, 2020, 4, 2000060 CrossRef CAS.
  132. L. Meng, C. Sun, R. Wang, W. Huang, Z. Zhao, P. Sun, T. Huang, J. Xue, J.-W. Lee and C. Zhu, J. Am. Chem. Soc., 2018, 140, 17255–17262 CrossRef CAS PubMed.
  133. C. Geffroy, E. Grana, T. Bessho, S. Almosni, Z. Tang, A. Sharma, T. Kinoshita, F. Awai, E. Cloutet and T. Toupance, ACS Appl. Energy Mater., 2020, 3, 1393–1401 CrossRef CAS.
  134. M. Luo, X. Zong, M. Zhao, Z. Sun, Y. Chen, M. Liang, Y. Wu and S. Xue, Chem. Eng. J., 2022, 442, 136136 CrossRef CAS.
  135. Y. Zhang, Y. Ren, X. Xie, Y. Wei, L. He, L. Fang, J. Zhang, Y. Yuan and P. Wang, Adv. Funct. Mater., 2022, 32, 2108855 CrossRef CAS.
  136. G. You, L. Li, S. Wang, J. Cao, L. Yao, W. Cai, Z. Zhou, K. Li, Z. Lin and H. Zhen, Adv. Energy Mater., 2022, 12, 2102697 CrossRef CAS.
  137. D. H. Lee, S. Lim, C. Kim, H. U. Lee, D. Chung, Y. Choi, J. Choi, Y. Kim, S. B. Cho and H. Il Kim, ACS Energy Lett., 2023, 8, 1839–1847 CrossRef CAS.
  138. X. Wu, D. Gao, X. Sun, S. Zhang, Q. Wang, B. Li, Z. Li, M. Qin, X. Jiang and C. Zhang, Adv. Mater., 2023, 35, 2208431 CrossRef CAS.
  139. Y. Bai, Z. Zhou, Q. Xue, C. Liu, N. Li, H. Tang, J. Zhang, X. Xia, J. Zhang and X. Lu, Adv. Mater., 2022, 34, 2110587 CrossRef CAS PubMed.
  140. S. Y. Son, K. Choi, J. Lee, H. Il Kim, T. Park and M. Kim, Chem. Mater., 2022, 34, 4856–4864 CrossRef CAS.
  141. J. Wang, Z. Yu, D. D. Astridge, Z. Ni, L. Zhao, B. Chen, M. Wang, Y. Zhou, G. Yang and X. Dai, ACS Energy Lett., 2022, 7, 3353–3361 CrossRef CAS.
  142. H. Lin, T. Doba, W. Sato, Y. Matsuo, R. Shang and E. Nakamura, Angew. Chem., Int. Ed., 2022, 61, e202203949 CrossRef CAS.
  143. X. Meng, Z. Cai, Y. Zhang, X. Hu, Z. Xing, Z. Huang, Z. Huang, Y. Cui, T. Hu and M. Su, Nat. Commun., 2020, 11, 3016 CrossRef CAS PubMed.
  144. F. Wu, Q. Xiao, X. Sun, T. Wu, Y. Hua and L. Zhu, Chem. Eng. J., 2022, 445, 136760 CrossRef CAS.
  145. C. Y. Xu, W. Hu, G. Wang, L. Niu, A. M. Elseman, L. Liao, Y. Yao, G. Xu, L. Luo and D. Liu, ACS Nano, 2019, 14, 196–203 CrossRef.
  146. L. Wan, Y. Zhao, Y. Tan, L. Lou and Z.-S. Wang, Chem. Eng. J., 2023, 455, 140569 CrossRef CAS.
  147. K. Eom, U. Kwon, S. S. Kalanur, H. J. Park and H. Seo, J. Mater. Chem. A, 2017, 5, 2563–2571 RSC.
  148. R. Wang, L. Xie, T. Wu, C. Ge and Y. Hua, Chem. Sci., 2023, 14, 2877–2886 RSC.
  149. Z. Iqbal, F. Zu, A. Musiienko, E. Gutierrez-Partida, H. Köbler, T. W. Gries, G. V Sannino, L. Canil, N. Koch and M. Stolterfoht, ACS Energy Lett., 2023, 8, 4304–4314 CrossRef CAS.
  150. H. Li, G. Xie, X. Wang, S. Li, D. Lin, J. Fang, D. Wang, W. Huang and L. Qiu, Adv. Sci., 2023, 10, 2300586 CrossRef CAS PubMed.
  151. E. Ochoa-Martinez, M. Ochoa, R. D. Ortuso, P. Ferdowsi, R. Carron, A. N. Tiwari, U. Steiner and M. Saliba, ACS Energy Lett., 2021, 6, 2626–2634 CrossRef CAS.
  152. A. Facchetti, Chem. Mater., 2011, 23, 733–758 CrossRef CAS.
  153. A. Facchetti, Chem. Mater., 2011, 23, 733–758 CrossRef CAS.
  154. W. Barford, Electronic and Optical Properties of Conjugated Polymers, OUP Oxford, 2013, vol. 129 Search PubMed.
  155. Z. Zhou, N. Luo, X. Shao, H. Zhang and Z. Liu, ChemPlusChem, 2023, 88, e202300261 CrossRef CAS PubMed.
  156. J. Qiao, C. Yang, Q. He, F. Bai and Y. Li, J. Appl. Polym. Sci., 2004, 92, 1459–1466 CrossRef CAS.
  157. R. S. Kularatne, H. D. Magurudeniya, P. Sista, M. C. Biewer and M. C. Stefan, J. Polym. Sci., Part A: Polym. Chem., 2013, 51, 743–768 CrossRef CAS.
  158. C. Yang, B. Cheng, J. Xu, J. Yu and S. Cao, EnergyChem, 2024, 6, 100116 CrossRef CAS.
  159. H. Zheng, X. Xu, S. Xu, G. Liu, S. Chen, X. Zhang, T. Chen and X. Pan, J. Mater. Chem. C, 2019, 7, 4441–4448 RSC.
  160. S. Valero, T. Soria, N. Marinova and J. L. Delgado, Polym. Chem., 2019, 10, 5726–5736 RSC.
  161. N. Yaghoobi Nia, M. Bonomo, M. Zendehdel, E. Lamanna, M. M. H. Desoky, B. Paci, F. Zurlo, A. Generosi, C. Barolo and G. Viscardi, ACS Sustain. Chem. Eng., 2021, 9, 5061–5073 CrossRef CAS.
  162. J. Jiang, Q. Wang, Z. Jin, X. Zhang, J. Lei, H. Bin, Z. Zhang, Y. Li and S. Liu, Adv. Energy Mater., 2018, 8, 1701757 CrossRef.
  163. T. Pratyusha, G. Sivakumar, D. Gupta and A. Yella, Mater. Today Proc., 2017, 4, 6820–6826 CrossRef.
  164. N. Yaghoobi Nia, M. Mendez, B. Paci, A. Generosi, A. Di Carlo and E. Palomares, ACS Appl. Energy Mater., 2020, 3, 6853–6859 CrossRef CAS.
  165. L. Shen, T. Zhu, X. Zhang, K. Gong, H. Wang and X. Gong, Nano Energy, 2022, 93, 106907 CrossRef CAS.
  166. F. Li, J. Yuan, X. Ling, Y. Zhang, Y. Yang, S. H. Cheung, C. H. Y. Ho, X. Gao and W. Ma, Adv. Funct. Mater., 2018, 28, 1706377 CrossRef.
  167. C. Ren, Y. He, S. Li, Q. Sun, Y. Liu, Y. Wu, Y. Cui, Z. Li, H. Wang and Y. Hao, Org. Electron., 2019, 70, 292–299 CrossRef CAS.
  168. A. F. Akbulatov, E. A. Khakina, N. A. Emelianov, O. A. Kraevaya, L. A. Frolova and P. A. Troshin, Sustain. Energy Fuels, 2024, 8, 4638–4645 RSC.
  169. T. Chakrabarti, S. Kaarthik and A. Perumal, ACS Appl. Energy Mater., 2025, 8(5), 2837–2844 CrossRef CAS.
  170. H. Choi, S. U. Ryu, D. H. Lee, H. Kim, S. Song, H. Il Kim and T. Park, ACS Energy Lett., 2024, 9, 3136–3168 CrossRef CAS.
  171. R. L. Carey, S. Giannini, S. Schott, V. Lemaur, M. Xiao, S. Prodhan, L. Wang, M. Bovoloni, C. Quarti and D. Beljonne, Nat. Commun., 2024, 15, 288 CrossRef CAS.
  172. T. Pratyusha, G. Sivakumar, D. Gupta and A. Yella, Mater. Today Proc., 2017, 4, 6820–6826 CrossRef.
  173. J. Grey, Nat. Mater., 2016, 15, 705–706 CrossRef CAS.
  174. L. Ding, Z.-D. Yu, X.-Y. Wang, Z.-F. Yao, Y. Lu, C.-Y. Yang, J.-Y. Wang and J. Pei, Chem. Rev., 2023, 123, 7421–7497 CrossRef CAS.
  175. Z. Ren, Z. Yang, W. Mu, T. Liu, X. Liu and Q. Wang, Adv. Mater., 2025, 37, 2416314 CrossRef CAS PubMed.
  176. F. He, T. Li, T. Shen, Y. Zhao, Z. Jin, Z. Zhang, Y. Pu, L. Deng, L. Qin and Y. Zhan, Adv. Funct. Mater., 2024, 2405611 CrossRef CAS.
  177. S. Kim, S. Sabury, C. A. R. Perini, T. Hossain, A. O. Yusuf, X. Xiao, R. Li, K. R. Graham, J. R. Reynolds and J.-P. Correa-Baena, ACS Energy Lett., 2024, 9, 4501–4508 CrossRef CAS PubMed.
  178. Z. Xie, J. Park, H. Kim, B. H. Cho, C. Lakshman, H.-Y. Park, T. Gokulnath, Y.-Y. Kim, J. Yoon and J.-S. Jee, ACS Appl. Mater. Interfaces, 2024, 16, 3359–3367 CrossRef CAS PubMed.
  179. Z. Azher, S. M. Said, A. Mainal, E. Y. Putri, H. F. B. Hizaddin, S. B. Ibrahim, N. Muhammad Sarih and M. U. Bin Mohd Junaidi, Int. J. Green Energy, 2024, 1–24 Search PubMed.
  180. L. Shen, X. Zhang, L. Liu, H. Sawwan, T. Zhu and X. Gong, ACS Appl. Electron. Mater., 2024, 6, 3925–3932 CrossRef CAS.
  181. I. R. Nikolényi and J. Tóth, J. Magn. Magn. Mater., 2021, 517, 167281 CrossRef.
  182. D. Molina, M. J. Álvaro-Martins and Á. Sastre-Santos, J. Mater. Chem. C, 2021, 9, 16078–16109 RSC.
  183. J. Wu, C. Fu, R. M. Pankow, Y. Chen, D. Zheng, Z. Lu, Y. Huang, T. J. Marks and A. Facchetti, Chem. Mater., 2023, 35, 10106–10118 CrossRef CAS.
  184. C. Gedeon, N. Del Rio, F. Furlan, A. Taddeucci, N. Vanthuyne, V. G. Gregoriou, M. J. Fuchter, G. Siligardi, N. Gasparini and J. Crassous, Adv. Mater., 2024, 2314337 CrossRef CAS.
  185. Z. Xing, S. Li, M. An and S. Yang, ChemSusChem, 2024, 17, e202301662 CrossRef CAS PubMed.
  186. M. Zhang, H. Liu, Y. Liu, M. Zeng, B. Zhao and S. Tan, Energy Fuels, 2023, 37, 6122–6128 CrossRef CAS.
  187. Z. Xie, J. Park, H. Kim, B. H. Cho, C. Lakshman, H.-Y. Park, T. Gokulnath, Y.-Y. Kim, J. Yoon and J.-S. Jee, ACS Appl. Mater. Interfaces, 2024, 16, 3359–3367 CrossRef CAS.
  188. C. Liu, L. Shao, S. Chen, Z. Hu, H. Cai and F. Huang, Prog. Polym. Sci., 2023, 101711 CrossRef CAS.
  189. F. Zahedi, S. Pourmahdian, M. Zendehdel, E. Rezasoltani and N. Y. Nia, Mater. Today Energy, 2025, 101938 CrossRef CAS.
  190. X. Liu and H. Jiang, Small, 2024, 20, 2404753 CrossRef CAS PubMed.
  191. B. Li, Q. Liu, J. Gong, S. Li, C. Zhang, D. Gao, Z. Chen, Z. Li, X. Wu and D. Zhao, Nat. Commun., 2024, 15, 2753 CrossRef CAS PubMed.
  192. R. Li, M. Liu, S. K. Matta, A. Almasri, J. Tian, H. Wang, H. P. Pasanen, S. P. Russo, P. Vivo and H. Zhang, Sol. RRL, 2023, 7, 2300367 CrossRef CAS.
  193. K. Hatakeyama-Sato and K. Oyaizu, Chem. Rev., 2023, 123, 11336–11391 CrossRef CAS.
  194. Z. Xie, Y. Do, S. J. Choi, H.-Y. Park, H. Kim, J. Kim, D. Song, T. Gokulnath, H.-B. Kim and I. W. Choi, J. Mater. Chem. A, 2023, 11, 9608–9615 RSC.
  195. R. Chetri and A. TN, Energy Fuels, 2025, 39, 9232–9261 CrossRef CAS.
  196. S. Mardi, M. Pea, A. Notargiacomo, N. Yaghoobi Nia, A. Di Carlo and A. Reale, Materials, 2020, 13, 1404 CrossRef CAS PubMed.
  197. S. W. Tong, J. Balapanuru, D. Fu and K. P. Loh, ACS Appl. Mater. Interfaces, 2016, 8, 29496–29503 CrossRef CAS.
  198. N. Yaghoobi Nia, E. Lamanna, M. Zendehdel, A. L. Palma, F. Zurlo, L. A. Castriotta and A. Di Carlo, Small, 2019, 15, 1904399 CrossRef CAS.
  199. Y. Wu, N. Ding, Y. Zhang, B. Liu, X. Zhuang, S. Liu, Z. Nie, X. Bai, B. Dong and L. Xu, Adv. Energy Mater., 2022, 12, 2200005 CrossRef CAS.
  200. M. Almalki, A. Dučinskas, L. C. Carbone, L. Pfeifer, L. Piveteau, W. Luo, E. Lim, P. A. Gaina, P. A. Schouwink and S. M. Zakeeruddin, Nanoscale, 2022, 14, 6771–6776 RSC.
  201. M. A. Ruíz-Sánchez, P. M. Moreno-Romero, P. G. Abrego-Martínez, D. M. Torres-Herrera, I. M. De Los Santos, L. Hechavarría-Difur, M. Courel, F. J. Sánchez-Rodríguez, H. Hu and H. J. Cortina-Marrero, Mater. Sci. Semicond. Process., 2023, 157, 107309 CrossRef.
  202. Z. Cao, Z. Li, M. Mooney, C. Do, K. Hong, S. Rondeau-Gagné, W. Xia and X. Gu, Macromolecules, 2024, 57, 10379–10388 CrossRef CAS PubMed.
  203. S. Subianto, R. Balu, L. de Campo, A. Sokolova, N. K. Dutta and N. R. Choudhury, ACS Appl. Mater. Interfaces, 2018, 10, 44116–44125 CrossRef CAS PubMed.
  204. C. Beaumont, S. Naqvi and M. Leclerc, Trends Chem., 2022, 4, 714–725 CrossRef CAS.
  205. D. Pan, J. S. Olsson and P. Jannasch, ACS Appl. Energy Mater., 2021, 5, 981–991 CrossRef.
  206. J. Wu, Z. Ren, F. Tang, W. Zhang, G. He, X. Yin, C. Gao and L. Wang, Chem. Eng. J., 2024, 490, 151650 CrossRef CAS.
  207. J. Wheeler and I. Zozoulenko, Adv. Electron. Mater., 2024, 2400662 CrossRef.
  208. K. Ji, J. Yuan, F. Li, Y. Shi, X. Ling, X. Zhang, Y. Zhang, H. Lu, J. Yuan and W. Ma, J. Mater. Chem. A, 2020, 8, 8104–8112 RSC.
  209. T. Zhu, L. Shen, H. Chen, Y. Yang, L. Zheng, R. Chen, J. Zheng, J. Wang and X. Gong, J. Mater. Chem. A, 2021, 9, 21910–21917 RSC.
  210. M. Xiao, R. L. Carey, H. Chen, X. Jiao, V. Lemaur, S. Schott, M. Nikolka, C. Jellett, A. Sadhanala and S. Rogers, Sci. Adv., 2021, 7, eabe5280 CrossRef CAS PubMed.
  211. S. Fratini, M. Nikolka, A. Salleo, G. Schweicher and H. Sirringhaus, Nat. Mater., 2020, 19, 491–502 CrossRef CAS.
  212. C.-C. Wu, E. Y. Li and P.-T. Chou, Chem. Sci., 2022, 13, 7181–7189 RSC.
  213. H. Zheng, Y. Guo, Z. Lin and J. Peng, Acc. Mater. Res., 2025, 6(9), 1158–1171 CrossRef CAS.
  214. M. Acik and S. B. Darling, J. Mater. Chem. A, 2016, 4, 6185–6235 RSC.
  215. G. Ganesh, A. Yasin, I. I. Misnon, A. Fakharuddin, L. Schmidt-Mende, M. H. Ab Rahim, S. Thomas and R. Jose, Sol. Energy, 2023, 257, 266–306 CrossRef CAS.
  216. C. Kim, H. Joung, H. J. Kim, K. Paeng, L. J. Kaufman and J. Yang, NPG Asia Mater., 2023, 15, 32 CrossRef CAS.
  217. W. Khelifi and C. K. Luscombe, Prog. Polym. Sci., 2024, 101804 CrossRef CAS.
  218. L. Zhao, W. Li, H. Qin, X. Yi, W. Zeng, Y. Zhao and H. Chen, Macromolecules, 2023, 56, 2990–3003 CrossRef CAS.
  219. D. Ko, H. Jung and M. Park, ACS Appl. Energy Mater., 2024, 7, 5986–5994 CrossRef CAS.
  220. T. Wu, P. Wang, L. Zheng, Y. Zhao and Y. Hua, Adv. Energy Mater., 2024, 14, 2400078 CrossRef CAS.
  221. J. Jiménez-López, B. M. D. Puscher, W. Cambarau, R. H. Fink, E. Palomares and D. M. Guldi, Nanoscale, 2019, 11, 23357–23365 RSC.
  222. I. Ahmed, L. Shi, H. Pasanen, P. Vivo, P. Maity, M. Hatamvand and Y. Zhan, Light Sci. Appl., 2021, 10, 174 CrossRef CAS PubMed.
  223. J. Jiménez-López, B. M. D. Puscher, W. Cambarau, R. H. Fink, E. Palomares and D. M. Guldi, Nanoscale, 2019, 11, 23357–23365 RSC.
  224. T. Wang, W. Deng, J. Cao and F. Yan, Adv. Energy Mater., 2023, 13, 2201436 CrossRef CAS.
  225. S. Wu, Z. Li, J. Zhang, X. Wu, X. Deng, Y. Liu, J. Zhou, C. Zhi, X. Yu and W. C. H. Choy, Adv. Mater., 2021, 33, 2105539 CrossRef CAS PubMed.
  226. Q. Lou, X. Xu, Z. Xu, T. Dai, H. Zhang, L. Zhao, F. Wang, E. Zhou, F. Gao and H. Zhou, Cell Rep Phys Sci., 2025, 6(4), 102515 CrossRef CAS.
  227. Y. Yang, Q. Xiong, J. Wu, Y. Tu, T. Sun, G. Li, X. Liu, X. Wang, Y. Du and C. Deng, Adv. Mater., 2024, 36, 2310800 CrossRef CAS PubMed.
  228. G. Welyab, M. Abebe, D. Mani, J. K. Paduvilan, L. Thottathi, A. Thankappan, S. Thomas, T. H. Wondimu and J. Y. Kim, Next Nanotechnol., 2024, 5, 100053 CrossRef.
  229. L. He, H. Zhang, D. Zhang, C. Gao, H. Su, D. Du, D. Ding, H. Liu and W. Shen, Adv. Funct. Mater., 2024, 34, 2403020 CrossRef CAS.
  230. R. F. Pineda, B. R. M. Lake, J. Troughton, I. Sanchez-Molina, O. Chepelin, S. Haque, T. Watson, M. P. Shaver and N. Robertson, Phys. Chem. Chem. Phys., 2018, 20, 25738–25745 RSC.
  231. D. A. M. de Alencar, G. Koch, F. De Rossi, A. Generosi, G. Ferraro, M. Bonomo, S. Noola, G. Pellis, P. Quagliotto and B. Paci, Adv. Sustain. Syst., 2025, 9, 2400674 CrossRef CAS.
  232. S. Yin, X. Luo, F. Tang, W. Zhong, W. Yang, Z. Xiong, Y. Lin, F. Peng and L. Ying, Adv. Energy Mater., 2025, 15, 2404575 CrossRef CAS.
  233. N. Rai, S. S. Reddy, A. D. Scully, K. L. J. Weerarathna, U. Bach and A. N. Simonov, Adv. Mater. Technol., 2024, 9, 2301539 CrossRef CAS.
  234. D. H. Lee, S. Lim, C. Kim, H. U. Lee, D. Chung, Y. Choi, J. Choi, Y. Kim, S. B. Cho and H. Il Kim, ACS Energy Lett., 2023, 8, 1839–1847 CrossRef CAS.
  235. Z. Zhang, J. Fu, Q. Chen, J. Zhang, Z. Huang, J. Cao, W. Ji, L. Zhang, A. Wang and Y. Zhou, Small, 2023, 19, 2206952 CrossRef CAS PubMed.
  236. Z. Xie, J. Park, H. Kim, B. H. Cho, C. Lakshman, H.-Y. Park, T. Gokulnath, Y.-Y. Kim, J. Yoon and J.-S. Jee, ACS Appl. Mater. Interfaces, 2024, 16, 3359–3367 CrossRef CAS PubMed.
  237. Z. Xie, J. Park, H. Kim, B. H. Cho, C. Lakshman, H.-Y. Park, T. Gokulnath, Y.-Y. Kim, J. Yoon and J.-S. Jee, ACS Appl. Mater. Interfaces, 2024, 16, 3359–3367 CrossRef CAS.
  238. T. R. Rana, M. Abbas, E. Schwartz, F. Jiang, M. Y. Yaman, Z. Xu, D. S. Ginger and D. MacKenzie, ACS Energy Lett., 2024, 9, 1888–1894 CrossRef CAS.
  239. J. T. Matondo, H. Hu, Y. Ding, M. Mateen, G. Cheng and J. Ding, Adv. Mater. Technol., 2024, 9, 2302082 CrossRef CAS.
  240. R. Keshavarzi, F. Hajisharifi, Z. Saki, M. Omrani, R. Sheibani, N. Afzali, M. Abdi-Jalebi, L. Vesce and A. Di Carlo, Nano Today, 2025, 61, 102600 CrossRef CAS.
  241. T. Yokoyama, T.-B. Song, D. H. Cao, C. C. Stoumpos, S. Aramaki and M. G. Kanatzidis, ACS Energy Lett., 2017, 2, 22–28 CrossRef CAS.
  242. K. B. Lohmann, J. B. Patel, M. U. Rothmann, C. Q. Xia, R. D. J. Oliver, L. M. Herz, H. J. Snaith and M. B. Johnston, ACS Energy Lett., 2020, 5, 710–717 CrossRef CAS PubMed.
  243. J. P. Chen, T. W. Colburn, J. Risner-Jamtgaard, A. Vailionis, A. Barnum, M. Golding, A. Carbone, A. C. Flick and R. H. Dauskardt, Matter, 2025, 8(3), 101990 CrossRef CAS.
  244. B. Tyagi, N. Kumar, H. B. Lee, M. M. Ovhal, V. V. Satale, A. Mohamed, D. Kim and J. Kang, Small Methods, 2024, 8, 2300237 CrossRef CAS.
  245. N. A. N. Ouedraogo, Y. Ouyang, B. Guo, Z. Xiao, C. Zuo, K. Chen, Z. He, G. O. Odunmbaku, Z. Ma and W. Long, Adv. Energy Mater., 2024, 14, 2401463 CrossRef CAS.
  246. F. Mathies, H. Eggers, B. S. Richards, G. Hernandez-Sosa, U. Lemmer and U. W. Paetzold, ACS Appl. Energy Mater., 2018, 1, 1834–1839 CrossRef CAS.
  247. Y. Zhong, R. Munir, J. Li, M.-C. Tang, M. R. Niazi, D.-M. Smilgies, K. Zhao and A. Amassian, ACS Energy Lett., 2018, 3, 1078–1085 CrossRef CAS.
  248. H. Fang, W. Shen, H. Guan, G. Chen, G. Li, W. Ai, S. Fu, Z. Xu, W. Chen and P. Jia, Adv. Mater., 2025, 37, 2414790 CrossRef CAS PubMed.
  249. J. E. Huddy and W. J. Scheideler, in 2024 IEEE 52nd Photovoltaic Specialist Conference (PVSC), IEEE, 2024, pp. 483–485 Search PubMed.
  250. J. E. Huddy, Y. Ye and W. J. Scheideler, Adv. Mater. Technol., 2022, 7, 2101282 CrossRef.
  251. P. Luo, S. Zhou, W. Xia, J. Cheng, C. Xu and Y. Lu, Adv. Mater. Interfaces, 2017, 4, 1600970 CrossRef.
  252. T. Huang, Y. Zhang, J. Sun, Y. Gu, F. Wang, Y.-W. Lu and P. Luo, ACS Appl. Energy Mater., 2024, 7, 11225–11232 CrossRef CAS.
  253. X. Tian, S. D. Stranks, J. Huang, V. M. Fthenakis, Y. Yang and F. You, Energy Environ. Sci., 2025, 18, 194–213 RSC.
  254. L. de Queiroz Corrêa, B. H. de Souza Miranda, G. de Amorim Soares and D. Bagnis, in Sustainable Strategies in Organic Electronics, Elsevier, 2022, pp. 521–560 Search PubMed.
  255. E. Aydin, T. G. Allen, M. De Bastiani, A. Razzaq, L. Xu, E. Ugur, J. Liu and S. De Wolf, Science, 2024, 383, eadh3849 CrossRef CAS PubMed.
  256. A. Khorasani, F. Mohamadkhani, M. Marandi, H. Luo and M. Abdi-Jalebi, Adv. Energy Sustain. Res., 2024, 5, 2300275 CrossRef CAS.
  257. J. Liu, M. Zhang, X. Sun, L. Xiang, X. Yang, X. Hu, Z. Wang, T. Hou, J. Qin and Y. Huang, Nano-Micro Lett., 2025, 17, 318 CrossRef PubMed.
  258. B. Kang and F. Yan, Energy Environ. Sci., 2025, 18, 3917–3954 RSC.
  259. P. Wang, Y. Han, Z. Lv, L. Miao, M. Wang, Z. Liu, T. Feng, M. Gao, Y. Shi and R. Cai, J. Phys. Chem. Lett., 2025, 16, 2695–2704 CrossRef CAS PubMed.
  260. N. Camaioni, C. Carbonera, L. Ciammaruchi, G. Corso, J. Mwaura, R. Po and F. Tinti, Adv. Mater., 2023, 35, 2210146 CrossRef CAS PubMed.
  261. H. Zheng, Z. Tang, P. Liang, H. Shu, L. Wang, J. Huang, A. A. Levin, P. N. Brunkov and Z. Liu, Thin Solid Films, 2021, 732, 138781 CrossRef CAS.
  262. G. G. Njema, B. C. Mosonik, C. C. Ahia and J. K. Kibet, Chem.–Eur. J., 2024, 30, e202403192 CrossRef CAS PubMed.
  263. L. Lin, L. Yang, G. Du, X. Li, Y. Li, J. Deng, K. Wei and J. Zhang, ACS Appl. Energy Mater., 2023, 6, 10303–10318 CrossRef CAS.
  264. J. P. Velásquez, E. A. Ramírez, A. Flórez, J. F. Montoya, R. Betancur, D. Ramírez, M. da S. Carvalho and F. Jaramillo, Adv. Eng. Mater., 2023, 25, 2201561 CrossRef.
  265. X. Tian, J. Wang, J. Ji and G. Lu, Renew. Energy, 2025, 240, 122170 CrossRef.
  266. F. Li, C. Gong, B. Fan, Z. Xing, X. Meng, S. Zhang, X. Hu and Y. Chen, Adv. Funct. Mater., 2022, 32, 2206412 CrossRef CAS.
  267. R. Chen, Y. Yang, Z. Dai, L. Yuan, J. Du, P. Yang, Y. Yang, H. Shen, Z. Liu and H. Wang, Nano Energy, 2023, 115, 108759 CrossRef CAS.
  268. Y. Wu, D. Liu, W. Chu, B. Wang, A. S. Vasenko and O. V Prezhdo, Nanoscale, 2025, 17, 2224–2234 RSC.
  269. Q. Wei, D. Zheng, L. Liu, J. Liu, M. Du, L. Peng, K. Wang and S. Liu, Adv. Mater., 2024, 36, 2406295 CrossRef CAS PubMed.
  270. H. Hu, D. B. Ritzer, A. Diercks, Y. Li, R. Singh, P. Fassl, Q. Jin, F. Schackmar, U. W. Paetzold and B. A. Nejand, Joule, 2023, 7, 1574–1592 CrossRef CAS.
  271. C. Chen, J. Chen, H. Han, L. Chao, J. Hu, T. Niu, H. Dong, S. Yang, Y. Xia and Y. Chen, Nature, 2022, 612, 266–271 CrossRef CAS PubMed.
  272. B. Dou, J. B. Whitaker, K. Bruening, D. T. Moore, L. M. Wheeler, J. Ryter, N. J. Breslin, J. J. Berry, S. M. Garner and F. S. Barnes, ACS Energy Lett., 2018, 3, 2558–2565 CrossRef CAS.
  273. K. Chen, Y. Zeng, X. Gao, X. Liu, L. Zhu and F. Wu, ChemSusChem, 2024, 17, e202301349 CrossRef CAS PubMed.
  274. T. Xu, Z. Wan, H. Tang, C. Zhao, S. Lv, Y. Chen, L. Chen, Q. Qiao and W. Huang, J. Alloys Compd., 2021, 859, 157784 CrossRef CAS.
  275. Y. Wang, Y. Wang, L. Deng, X. Li, X. Zhang, H. Wang, C. Li, Z. Shi, T. Hu and K. Liu, Adv. Energy Mater., 2024, 14, 2402066 CrossRef CAS.
  276. J. T. Matondo, H. Hu, Y. Ding, M. Mateen, G. Cheng and J. Ding, Adv. Mater. Technol., 2024, 9, 2302082 CrossRef CAS.
  277. J. T. Matondo, H. Hu, Y. Ding, M. Mateen, G. Cheng and J. Ding, Adv. Mater. Technol., 2024, 9, 2302082 CrossRef CAS.
  278. M. Wang, Y. Zhao, X. Jiang, Y. Yin, I. Yavuz, P. Zhu, A. Zhang, G. S. Han, H. S. Jung and Y. Zhou, Joule, 2022, 6, 1032–1048 CrossRef CAS.
  279. J. Hu, C. Wang, S. Qiu, Y. Zhao, E. Gu, L. Zeng, Y. Yang, C. Li, X. Liu and K. Forberich, Adv. Energy Mater., 2020, 10, 2000173 CrossRef CAS.
  280. L. J. Sutherland, H. C. Weerasinghe and G. P. Simon, Adv. Energy Mater., 2021, 11, 2101383 CrossRef CAS.
  281. X. Tan, S. Wang, Q. Zhang, H. Wang, H. Liu, W. Li, L. Zhu, T. Song, Z. Cui and Y. Bai, J. Mater. Chem. A, 2023, 11, 1920–1926 RSC.
  282. J. Cho and P. V Kamat, Chem. Mater., 2020, 32, 6206–6212 CrossRef CAS.
  283. L. Li, Z. Huang, X. Meng, Z. Xing, B. Fan, J. Li and Y. Chen, Adv. Mater., 2024, 36, 2310752 CrossRef CAS PubMed.
  284. Z. Chu, B. Fan, W. Shi, Z. Xing, C. Gong, J. Li, L. Li, X. Meng, M. B. K. Niazi and X. Hu, Adv. Funct. Mater., 2025, 2424191 CrossRef CAS.
  285. L. Li, Z. Huang, X. Meng, Z. Xing, B. Fan, J. Li and Y. Chen, Adv. Mater., 2024, 36, 2310752 CrossRef CAS.
  286. F. Wang, T. Liu, Y. Liu, Y. Zhou, X. Dong, Y. Zhang, X. Shi, Y. Dou, Z. Ren and L. Wang, Adv. Mater., 2024, 36, 2412059 CrossRef CAS.
  287. T. Liu, Z. Ren, Y. Liu, Y. Zhang, J. Liang, F. Cheng, Y. Li, X. Shi, Y. Dou and X. Hu, Sci. Adv., 2025, 11, eadu3493 CrossRef CAS PubMed.
  288. M. M. Byranvand, F. Behboodi-Sadabad, A. A. Eliwi, V. Trouillet, A. Welle, S. Ternes, I. M. Hossain, M. R. Khan, J. A. Schwenzer and A. Farooq, J. Mater. Chem. A, 2020, 8, 20122–20132 RSC.
  289. S. Smeets, Q. Liu, J. Vanderspikken, T. J. Quill, S. Gielen, L. Lutsen, K. Vandewal and W. Maes, Chem. Mater., 2023, 35, 8158–8169 CrossRef CAS.
  290. Z. Wang, Z. Peng, N. Balar and H. Ade, J. Mater. Chem. A, 2024, 12, 113–120 RSC.
  291. C.-K. Pan, Y.-F. Ding, Z.-F. Yao, X.-Y. Wang, Y.-Y. Zhou, X.-Z. Wu, J.-Y. Wang and J. Pei, Macromolecules, 2024, 57, 4107–4117 CrossRef CAS.
  292. Z. Chen, B. Wu, B. Yin, Y. Li, K. Liu, S. Kim, X. Gao, Z. Huang, Z. Liu and C. Yang, Macromol. Chem. Phys., 2025, 2400464 CrossRef CAS.
  293. F. Ershad, W. Song and C. Yu, Nat. Photonics, 2023, 17, 928–930 CrossRef CAS.
  294. L. Li, Z. Huang, X. Meng, Z. Xing, B. Fan, J. Li and Y. Chen, Adv. Mater., 2024, 36, 2310752 CrossRef CAS PubMed.
  295. H. Hu, M. Chen, N. Yao, L. Wu, Q. Zhong, B. Song, M. Cao and Q. Zhang, ACS Appl. Mater. Interfaces, 2021, 13, 4017–4025 CrossRef CAS PubMed.
  296. A. Alasiri, K. Zubair, S. Rassel, D. Ban and O. D. Alshehri, Heliyon, 2024, 10(20), e39141 CrossRef CAS PubMed.
  297. N. A. N. Ouedraogo, M. Yang, C. He, Y. Chen, X. Zhang, H. Yan, C. B. Han and Y. Zhang, Sustain. Energy Fuels, 2020, 4, 4257–4263 RSC.
  298. J. Peng, J. I. Khan, W. Liu, E. Ugur, T. Duong, Y. Wu, H. Shen, K. Wang, H. Dang and E. Aydin, Adv. Energy Mater., 2018, 8, 1801208 CrossRef.
  299. L. Hu, L. Zhang, W. Ren, C. Zhang, Y. Wu, Y. Liu, Q. Sun, Z. Dai, Y. Cui and L. Cai, J. Mater. Chem. C, 2022, 10, 9714–9722 RSC.
  300. X. Meng, J. Lin, X. Liu, X. He, Y. Wang, T. Noda, T. Wu, X. Yang and L. Han, Adv. Mater., 2019, 31, 1903721 CrossRef CAS PubMed.
  301. J. Kang, R. Huang, S. Guo, G. Han, X. Sun, I. Ismail, C. Ding, F. Li, Q. Luo and Y. Li, Sol. Energy, 2021, 217, 105–112 CrossRef CAS.
  302. F. Qi, Y. Li, F. R. Lin and A. K. Jen, ChemSusChem, 2024, e202301559 CrossRef CAS.
  303. C. Li, J. Yin, R. Chen, X. Lv, X. Feng, Y. Wu and J. Cao, J. Am. Chem. Soc., 2019, 141, 6345–6351 CrossRef CAS PubMed.
  304. N. Y. Nia, M. Zendehdel, M. Abdi-Jalebi, L. A. Castriotta, F. U. Kosasih, E. Lamanna, M. M. Abolhasani, Z. Zheng, Z. Andaji-Garmaroudi and K. Asadi, Nano Energy, 2021, 82, 105685 CrossRef.
  305. M. Haghighi, N. Ghazyani, S. Mahmoodpour, R. Keshtmand, A. Ghaffari, H. Luo, R. Mohammadpour, N. Taghavinia and M. Abdi-Jalebi, Sol. RRL, 2023, 7, 2201080 CrossRef CAS.
  306. C. Fei, N. Li, M. Wang, X. Wang, H. Gu, B. Chen, Z. Zhang, Z. Ni, H. Jiao and W. Xu, Science, 2023, 380, 823–829 CrossRef CAS PubMed.
  307. C. Fei, N. Li, M. Wang, X. Wang, H. Gu, B. Chen, Z. Zhang, Z. Ni, H. Jiao and W. Xu, Science, 2023, 380, 823–829 CrossRef CAS PubMed.
  308. H. Kassem, A. Salehi and M. Kahrizi, Energy Technol., 2024, 12, 2301032 CrossRef CAS.
  309. C. Sun, F. Pan, H. Bin, J. Zhang, L. Xue, B. Qiu, Z. Wei, Z.-G. Zhang and Y. Li, Nat. Commun., 2018, 9, 743 CrossRef PubMed.
  310. X. Rodríguez-Martínez, E. Pascual-San-José and M. Campoy-Quiles, Energy Environ. Sci., 2021, 14, 3301–3322 RSC.
  311. N. Y. Nia, F. Matteocci, L. Cina and A. Di Carlo, ChemSusChem, 2017, 10, 3854–3860 CrossRef CAS PubMed.
  312. L. Gao, J. Lin, L. Wang and L. Du, Acc. Mater. Res., 2024, 5, 571–584 CrossRef CAS.
  313. L. Ding, Z.-D. Yu, X.-Y. Wang, Z.-F. Yao, Y. Lu, C.-Y. Yang, J.-Y. Wang and J. Pei, Chem. Rev., 2023, 123, 7421–7497 CrossRef CAS PubMed.
  314. S. M. Lee, K. H. Park, S. Jung, H. Park and C. Yang, Nat. Commun., 2018, 9, 1–8 CrossRef.
  315. W. Shin, W. Ko, S.-H. Jin, T. Earmme and Y.-J. Hwang, Chem. Eng. J., 2021, 412, 128572 CrossRef CAS.
  316. X. Wang, Z. Ding, X. Huang, X. Liu, Y. Wang, Y. Wang and W. Huang, J. Mater. Chem. C, 2024, 12, 351–387 RSC.
  317. P. Holzhey, M. Prettl, S. Collavini, N. L. Chang and M. Saliba, Joule, 2023, 7, 257–271 CrossRef CAS.
  318. J. J. Cordell, M. Woodhouse and E. L. Warren, Joule, 2025, 9(2), 101781 CrossRef CAS.
  319. P. Zhu, C. Chen, J. Dai, Y. Zhang, R. Mao, S. Chen, J. Huang and J. Zhu, Adv. Mater., 2024, 36, 2307357 CrossRef CAS PubMed.
  320. P. Zhu, C. Chen, J. Dai, Y. Zhang, R. Mao, S. Chen, J. Huang and J. Zhu, Adv. Mater., 2024, 36, 2307357 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.