Fact-checked by Grok 2 weeks ago

Perovskite solar cell

A perovskite solar cell (PSC) is a type of thin-film photovoltaic device that employs metal halide perovskites, typically with the ABX₃ crystal structure where A is a monovalent cation (such as methylammonium or cesium), B is a divalent metal (often lead or tin), and X is a halide (like iodide or bromide), as the primary light-absorbing and charge-generating layer. These materials, exemplified by methylammonium lead iodide (MAPbI₃), exhibit strong visible-light absorption, long charge-carrier diffusion lengths, and defect tolerance, allowing the cell to convert sunlight into electricity by absorbing photons to create electron-hole pairs, which are then separated and collected via adjacent electron- and hole-transport layers flanked by electrodes. Unlike traditional crystalline silicon solar cells, PSCs feature an active layer only a few hundred nanometers thick, enabling fabrication through solution-based methods like printing or coating on flexible substrates. The development of PSCs began in 2009 with the first demonstration of a dye-sensitized solar cell incorporating a solid-state perovskite absorber, achieving an initial power conversion efficiency (PCE) of 3.8%. Rapid advancements followed, driven by improvements in material composition, interface engineering, and device architecture, propelling the certified PCE of single-junction PSCs from under 10% in 2012 to over 26% by 2025. In tandem configurations, where perovskites are stacked with silicon or other absorbers to capture a broader spectrum, efficiencies have approached 34%, surpassing single-junction silicon limits and highlighting their potential for next-generation photovoltaics. Key milestones include the introduction of solid-state architectures in 2012, which boosted stability and efficiency, and ongoing innovations like inverted p-i-n structures that reached a certified PCE of 26.96% in late 2025. PSCs offer several advantages over conventional silicon-based solar cells, including lower production costs due to the use of earth-abundant materials and scalable solution-processing techniques that require less energy and capital investment than high-temperature silicon manufacturing. Their structural versatility allows for bandgap tuning to optimize light absorption, and their mechanical flexibility supports applications in wearable electronics, building-integrated photovoltaics, and lightweight modules. However, challenges persist, particularly in long-term stability against moisture, heat, and light-induced degradation, as well as scalability for large-area production and the environmental concerns of lead content, which have spurred research into lead-free alternatives and encapsulation strategies. As of 2025, PSCs represent one of the fastest-progressing photovoltaic technologies, with global research focusing on commercialization through hybrid tandems and improved durability to meet Department of Energy targets by 2026. Recent breakthroughs, such as perovskite-perovskite-silicon triple-junction cells exceeding 27% PCE, underscore their role in achieving terawatt-scale solar deployment while addressing climate goals. Despite hurdles, ongoing efforts in materials innovation and manufacturing optimization position PSCs as a transformative option for sustainable energy.

Fundamentals

Crystal Structure and Composition

Perovskite materials used in solar cells adopt the general formula ABX_3, where A is a monovalent cation, B is a divalent metal cation, and X is a monovalent anion, typically a halide. The A-site cation can be organic, such as methylammonium (MA, CH_3NH_3^+) or formamidinium (FA, HC(NH_2)_2^+), or inorganic, like cesium (Cs^+); the B-site is commonly lead (Pb^{2+}) or tin (Sn^{2+}); and the X-site consists of halides such as iodide (I^-), bromide (Br^-), or chloride (Cl^-). This composition enables tunable optoelectronic properties while maintaining the characteristic perovskite lattice. The crystal structure of these halide perovskites features a framework of corner-sharing BX_6 octahedra, forming a three-dimensional network where the A cation resides in the 12-fold coordinated cuboctahedral voids between the octahedra. At high temperatures, many compositions exhibit a cubic structure (space group Pm\bar{3}m), which transitions to lower-symmetry phases upon cooling: tetragonal (I4/mcm) at intermediate temperatures, and orthorhombic (Pnma) at low temperatures. These phase transitions arise from octahedral tilting and distortions driven by the size mismatch between ions, quantified by the Goldschmidt tolerance factor t = \frac{r_A + r_X}{\sqrt{2}(r_B + r_X)}, where r denotes ionic radii, with $0.8 < t < 1 favoring the perovskite phase. The cubic form provides ideal symmetry for charge transport, while the tetragonal phase, prevalent at room temperature for common solar cell materials, introduces slight distortions that influence bandgap and stability. Hybrid organic-inorganic perovskites, such as MAPbI_3, incorporate organic A cations that introduce dynamic disorder and softer lattice vibrations compared to all-inorganic variants like CsPbI_3, which use smaller inorganic cations for enhanced thermal stability. In MAPbI_3, the organic MA cation rotates freely in the cubic phase but orients along the c-axis in the tetragonal phase, contributing to ferroelectric-like properties. All-inorganic CsPbI_3 stabilizes in the cubic \alpha-phase under specific conditions, avoiding the non-perovskite yellow \delta-phase, though it requires careful processing to maintain the photoactive black phase at room temperature. These structural differences impact moisture resistance and phase purity in device applications. Beyond three-dimensional bulk structures, lower-dimensional variants such as two-dimensional (2D) layered perovskites derive from Ruddlesden-Popper phases with the formula (A')_2(A)_{n-1}B_nX_{3n+1}, where A' is a larger organic spacer cation (e.g., butylammonium) that separates the structure into quantum-well-like sheets of n corner-sharing BX_6 octahedra. These 2D structures exhibit enhanced stability against environmental degradation due to hydrophobic organic barriers, while retaining optoelectronic functionality through inorganic layers, with n=1 yielding purely 2D phases and higher n approaching 3D behavior.

Key Materials and Variants

The absorber layer in perovskite solar cells primarily consists of hybrid organic-inorganic materials such as formamidinium lead triiodide (FAPbI₃), which offers a suitable bandgap for efficient light absorption and demonstrates superior initial photovoltaic performance along with enhanced dark storage stability compared to methylammonium-based counterparts. Inorganic variants like cesium lead bromide (CsPbBr₃) provide improved thermal and phase stability, making them suitable for applications requiring robustness under operational stresses, though they often exhibit wider bandgaps that limit absorption in the near-infrared region. To address phase instability issues in pure FAPbI₃ or CsPbI₃, mixed-cation perovskites—incorporating combinations such as formamidinium-cesium or methylammonium-cesium-lead halides—have been developed, enhancing the cubic phase retention at room temperature and improving overall device longevity without significantly compromising efficiency. Electron transport layers (ETLs) facilitate efficient extraction of photogenerated electrons from the absorber; titanium dioxide (TiO₂) is a widely adopted ETL due to its chemical stability, high electron mobility, and ability to form a compact blocking layer that minimizes recombination, often processed via low-temperature methods for flexible substrates. Tin dioxide (SnO₂) serves as an alternative ETL, offering higher electron mobility and better energy level alignment with the perovskite absorber, which reduces hysteresis and boosts open-circuit voltage in planar architectures, particularly when doped or surface-modified for defect passivation. Hole transport layers (HTLs) symmetrically manage hole extraction; 2,2',7,7'-tetrakis[N,N-di(4-methoxyphenyl)amino]-9,9'-spirobifluorene (Spiro-OMeTAD) is a benchmark organic HTL prized for its excellent hole mobility and matched valence band energy, though it requires doping for optimal conductivity in high-efficiency devices. Poly[bis(4-phenyl)(2,4,6-trimethylphenyl)amine] (PTAA) emerges as a robust HTL alternative, providing hydrophobic protection against moisture ingress, thinner film requirements for comparable hole transport, and improved long-term stability over Spiro-OMeTAD in inverted configurations. Transparent conductive oxides like indium tin oxide (ITO) or fluorine-doped tin oxide (FTO) form the front electrodes, enabling high optical transmittance (>80%) while supporting ohmic contact for charge collection in rigid or flexible substrates. Back electrodes typically employ metals such as gold (Au) or silver (Ag), which provide low work functions for efficient hole collection and reflection of unabsorbed light to enhance photocurrent, though Ag requires encapsulation to prevent halide migration-induced degradation. Variants of perovskite materials extend functionality beyond 3D structures; two-dimensional (2D) hybrid perovskites with the general formula (RNH₃)₂PbI₄—where R is an organic spacer cation like butylammonium—feature layered quantum-well architectures that confer exceptional mechanical flexibility and resistance to environmental degradation, ideal for bendable solar cells. These 2D variants leverage van der Waals bonding between inorganic sheets for enhanced tensile strength and reduced cracking under strain, outperforming 3D perovskites in flexible device lifetimes. Additionally, incorporating chiral ligands into perovskite formulations induces spin-selective charge transport, where the molecular handedness filters spin-polarized carriers to suppress recombination and enable applications in spintronic-enhanced photovoltaics.

Operating Principles

Perovskite solar cells operate through the photovoltaic effect, where light absorption in the perovskite absorber layer generates charge carriers that are separated and collected to produce electrical power. The perovskite materials exhibit a direct bandgap tunable from approximately 1.5 to 2.5 eV by varying the halide composition and cation substitutions, allowing optimization for different portions of the solar spectrum. This direct bandgap, combined with a high absorption coefficient exceeding $10^5 cm^{-1}, enables efficient capture of photons in thin films typically 300–500 nm thick, converting a significant fraction of incident sunlight into excitons. Upon photoexcitation, absorbed photons promote electrons from the valence band to the conduction band, forming electron-hole pairs or excitons. The low exciton binding energy in three-dimensional perovskites, typically less than 50 meV, is below the thermal energy at room temperature (about 25 meV), promoting rapid thermal dissociation into free charge carriers without requiring additional energy input. This efficient dissociation contrasts with materials like organic semiconductors, where higher binding energies hinder free carrier generation, and supports high quantum yields for photocurrent.30170-3) The free electrons and holes exhibit advantageous charge carrier dynamics, including ambipolar diffusion lengths exceeding 1 μm and mobilities ranging from 10 to 100 cm²/V·s, which minimize recombination losses and enable carriers to traverse the absorber layer effectively. In a typical device, photoexcitation occurs within the perovskite layer, followed by ultrafast charge separation at the heterojunction interfaces: electrons transfer to the electron transport layer (ETL) due to favorable energy alignment, while holes move to the hole transport layer (HTL). These carriers are then transported through the respective layers and collected at the electrodes, driving current flow under an applied bias or load.30170-3) The efficiency of this process is encapsulated in the power conversion efficiency (PCE), given by \eta = \frac{\mathrm{FF} \times V_\mathrm{oc} \times J_\mathrm{sc}}{P_\mathrm{in}}, where FF is the fill factor (dimensionless, reflecting voltage-current curve shape), V_\mathrm{oc} is the open-circuit voltage, J_\mathrm{sc} is the short-circuit current density, and P_\mathrm{in} is the incident light power density (standardized at 100 mW/cm² under AM1.5G illumination). This formula arises from the maximum extractable electrical power (\mathrm{FF} \times V_\mathrm{oc} \times J_\mathrm{sc}) normalized to the input optical power, providing a direct measure of device performance.

Advantages and Performance Limits

Efficiency Achievements and Theoretical Limits

The theoretical efficiency limit for single-junction solar cells, known as the Shockley-Queisser (SQ) limit, arises from the detailed balance principle, which accounts for the fundamental losses due to radiative recombination, thermalization of high-energy photons, and transmission of sub-bandgap photons. For a perovskite absorber with a bandgap of approximately 1.5 eV, typical for compositions like methylammonium lead iodide (MAPbI₃), the SQ limit is around 33% under standard AM1.5 solar illumination. This limit is derived from the maximum power output, given by the equation for power conversion efficiency η = (J_{sc} × V_{oc} × FF) / P_{in}, where J_{sc} is the short-circuit current density, V_{oc} is the open-circuit voltage, FF is the fill factor, and P_{in} is the incident power density (100 mW/cm²); the SQ model assumes only radiative recombination, setting an upper bound on V_{oc} to about 0.9–1.0 V for a 1.5 eV bandgap due to the blackbody radiation balance. Perovskite solar cells have rapidly approached this SQ limit, with certified single-junction efficiencies reaching 27.0% as of 2025, verified by the National Renewable Energy Laboratory (NREL). This record, achieved in laboratory-scale devices using optimized halide perovskite compositions, represents over 80% of the SQ limit for the material's bandgap, highlighting the technology's potential to rival or exceed silicon photovoltaics in single-junction configurations. Earlier milestones included efficiencies surpassing 25% by 2020, driven by compositional tuning to minimize bandgap-related losses while maintaining strong visible-light absorption. Key to these high efficiencies is the inherent defect tolerance of halide perovskites, where shallow defects near the band edges act as benign traps rather than deep recombination centers, allowing high carrier mobilities and diffusion lengths exceeding 1 μm even in solution-processed films. Additionally, low non-radiative recombination rates—often below 10¹⁰ s⁻¹—stem from suppressed Shockley-Read-Hall processes, enabling near-radiative V_{oc} values close to the SQ ideal. These properties contrast with many inorganic semiconductors, where defects necessitate complex purification, and have enabled perovskites to achieve V_{oc} around 1.2 V, J_{sc} exceeding 24 mA/cm², and FF above 85% in top-performing cells under standard conditions.

Comparison to Conventional Solar Cells

Perovskite solar cells offer significant cost advantages over conventional silicon solar cells primarily due to their low-temperature processing requirements, typically below 150°C, which enables solution-based fabrication methods that consume far less energy than the high-temperature steps (often exceeding 1000°C) involved in silicon crystal growth and wafering. This contrast reduces manufacturing energy inputs and material waste, with potential for lower production costs through techniques like printing or coating, in contrast to silicon's energy-intensive sawing and purification processes that account for substantial capital and operational expenses. In terms of efficiency, perovskite solar cells have achieved power conversion efficiencies exceeding 26%, closely approaching the 25-27% range of high-performance silicon cells, while utilizing much thinner active layers of 300-500 nm compared to silicon's 180 μm wafers. This thinner absorber layer enhances light absorption efficiency due to perovskites' direct bandgap and high absorption coefficients, allowing comparable performance with less material. Scalability represents another key differentiator, as perovskite solar cells are compatible with continuous roll-to-roll processing on flexible substrates, facilitating large-area production at high speeds, whereas silicon solar cells rely on batch-oriented wafer handling and assembly that limit throughput and increase complexity for module-scale manufacturing. The environmental footprint of perovskite solar cells is notably lower, with an energy payback time of approximately 3-4 months— the period required to generate energy equivalent to that used in production—compared to 1-2 years for silicon solar cells, driven by reduced material usage and processing energy demands. This shorter payback contributes to a reduced greenhouse gas emission profile over the device lifecycle.

Fabrication and Processing

Solution-Based Deposition Techniques

Solution-based deposition techniques represent a cornerstone of perovskite solar cell fabrication due to their compatibility with low-cost, scalable liquid-phase processing at ambient conditions. These methods involve dissolving perovskite precursors—such as lead halides (e.g., PbI₂) and organic cations (e.g., methylammonium iodide, MAI)—in polar solvents like dimethylformamide (DMF) or dimethyl sulfoxide (DMSO) to form a precursor solution, which is then applied to a substrate and crystallized through thermal annealing, typically at temperatures around 100°C. This approach enables rapid prototyping in laboratory settings and leverages solution chemistry to achieve high-quality perovskite films with controlled morphology, contrasting with more capital-intensive vacuum-based methods. Early demonstrations of these techniques highlighted their potential for efficiencies exceeding 15% shortly after the field's inception. One-step solution deposition, pioneered in the initial reports of solid-state perovskite solar cells, entails spin-coating a single precursor solution containing all ionic components onto a substrate, followed by annealing to induce perovskite crystallization. In this process, a stoichiometric mixture of PbI₂ and MAI in a DMF/DMSO solvent is deposited at high spin speeds (e.g., 4000 rpm) to form a thin wet film, which is then heated to evaporate the solvent and promote uniform phase formation. This method's simplicity facilitated the first certified efficiencies above 9% in 2012, as it allows for straightforward composition tuning to optimize bandgap and optoelectronic properties. However, challenges like incomplete solvent removal can lead to pinholes or irregular grain growth, necessitating solvent engineering for improved film uniformity. Two-step deposition addresses limitations of the one-step approach by sequentially applying inorganic and organic precursors, enabling better control over film morphology and stoichiometry. The process begins with spin-coating or dipping a lead iodide (PbI₂) solution onto the substrate, drying to form a compact PbI₂ layer, followed by exposure to an organic halide solution (e.g., MAI in isopropanol) that converts the PbI₂ to perovskite through an interfacial reaction. This sequential method, introduced in 2013, achieved a breakthrough power conversion efficiency of 15% by promoting large, oriented grains that enhance charge transport and reduce recombination. It is particularly effective for methylammonium-based perovskites, though it requires precise control of dipping time and concentration to avoid incomplete conversion or excess organic residues. Anti-solvent engineering enhances the quality of films produced via one- or two-step methods by inducing rapid, homogeneous nucleation during spin-coating. In this technique, a non-polar anti-solvent such as chlorobenzene or toluene is dripped onto the spinning wet precursor film, which extracts the polar solvent and triggers supersaturation, leading to compact, pinhole-free perovskite layers with micrometer-sized grains. First demonstrated in 2014, this approach propelled efficiencies beyond 19% by minimizing defects and improving surface coverage, as the fast crystallization suppresses unwanted phases like δ-FAPbI₃. Variations include blade-coating adaptations for larger areas, but the core benefit lies in its ability to tailor crystal growth without altering precursor chemistry. For transitioning beyond lab-scale spin-coating, solution-based techniques have been adapted to large-area formats using slot-die and blade coating, which deposit uniform precursor films over substrates up to square meters. Slot-die coating extrudes the solution through a narrow slot onto a moving substrate, enabling precise thickness control (e.g., 200-500 nm) and continuous processing, with reported efficiencies of 15-18% on 10 cm² modules using anti-solvent-free formulations. Blade coating, involving a doctor blade spreading the solution across the surface, offers similar scalability and has demonstrated 17% efficiency on rigid substrates by optimizing shear forces for grain alignment. These methods maintain the cost advantages of solution processing while approaching industrial viability, though they demand rheological tuning of inks for defect-free coverage.

Vapor and Other Deposition Methods

Vapor deposition methods for perovskite solar cells involve gas-phase techniques that enable the formation of high-quality thin films without solvents, offering precise control over material deposition. These approaches, conducted under vacuum conditions, typically include thermal evaporation and chemical vapor deposition (CVD) variants, which have been pivotal in achieving efficiencies exceeding 20% in laboratory devices. Co-evaporation is a prominent vacuum-based technique where multiple precursors, such as lead chloride (PbCl₂) and methylammonium iodide (MAI), are simultaneously evaporated from separate sources onto a heated substrate, allowing in situ formation of the perovskite layer. This method ensures uniform composition and compact films with minimal defects, as demonstrated in early work achieving 20% power conversion efficiency (PCE) in n-i-p configured devices using MAPbI₃. Sequential vapor transport, an alternative, deposits precursors in steps—first evaporating an inorganic halide like PbI₂, followed by exposure to an organic vapor such as MAI—facilitating controlled crystallization and high crystallinity. This sequential process has yielded n-i-p perovskite solar cells with PCEs over 15%, benefiting from improved phase purity and reduced grain boundaries. For inorganic perovskites, such as CsPbI₃, chemical vapor deposition (CVD) variants like close-space sublimation or dual-source vapor deposition are employed to grow stable, wide-bandgap films suitable for tandem solar cells. These methods involve sublimating metal halides and alkali sources under controlled pressure, resulting in pinhole-free layers with enhanced optoelectronic properties and PCEs up to 16% in single-junction devices. Hybrid methods, including vapor-assisted solution processing (VASP), combine elements of solution and vapor techniques by first spin-coating a lead halide film and then exposing it to organic vapors at low temperatures below 150°C, promoting uniform conversion to perovskite. VASP has produced large-grain films with PCEs ~16.8%, leveraging the kinetic control of vapor diffusion to minimize defects. Compared to solution-based methods, vapor and hybrid deposition offer superior stoichiometry control through precise precursor flux ratios and reduced pinhole formation due to solvent-free environments, leading to denser films with longer carrier lifetimes. These advantages make them particularly suitable for industrial-scale precision and compatibility with multi-junction architectures.

Scalability and Large-Area Challenges

Scaling perovskite solar cells to large-area modules presents significant challenges, particularly in achieving uniform absorber layers through methods like slot-die coating. While lab-scale devices routinely exceed 25% power conversion efficiency (PCE), module-scale efficiencies typically range from 15-24% due to non-uniform film formation, including pinholes and cracks that arise during the drying process of wet films on extended substrates. These uniformity issues stem from variations in ink flow, substrate wettability, and evaporation rates, which disrupt the crystallization of the perovskite absorber and lead to inconsistent light absorption and charge generation across larger areas. In July 2025, a collaborative effort between NREL and CubicPV achieved a certified 24% PCE for a perovskite mini-module, highlighting ongoing progress. Charge transport layers and electrodes introduce additional scaling hurdles, as their deposition must maintain high conductivity and minimal series resistance over large surfaces. Inkjet printing has emerged as a promising technique for hole transport layers (HTL), enabling precise patterning and uniform coverage without solvent wastage, as demonstrated in all-inkjet-printed devices achieving over 17% PCE. For back contacts, thermal evaporation provides excellent uniformity and adhesion, facilitating large-area deposition of metals like gold or silver, though it requires vacuum conditions that can limit throughput compared to solution-based alternatives. Yield in large-area fabrication is further compromised by defect-related factors, such as grain boundary defects that act as recombination centers, reducing carrier collection efficiency and overall module performance. In printing processes, the coffee-ring effect exacerbates inhomogeneity by causing solute accumulation at droplet edges, resulting in pinholes, poor interfacial contact, and lower device reliability. Recent advancements have mitigated some of these issues; for instance, in March 2025, UtmoLight reported 18.1% PCE for a 0.72 m² perovskite module, certified by the National Renewable Energy Laboratory, highlighting progress in uniform coating and defect passivation for practical scaling.

Device Architectures

Standard n-i-p Configuration

The standard n-i-p configuration of perovskite solar cells employs a layered architecture that promotes efficient charge separation and extraction, with light incident from the substrate side. This setup consists of a transparent conductive oxide substrate, such as fluorine-doped tin oxide (FTO)-coated glass, serving as the bottom electrode; a compact n-type electron transport layer (ETL), typically titanium dioxide (TiO₂) deposited at around 500°C via spray pyrolysis to form a dense blocking layer approximately 50-100 nm thick; an intrinsic perovskite absorber layer, often methylammonium lead iodide (CH₃NH₃PbI₃) or formamidinium-based variants with thicknesses of 300-500 nm; a p-type hole transport layer (HTL), commonly 2,2',7,7'-tetrakis[(N,N-di-4-methoxyphenyl)amino]-9,9'-spirobifluorene (Spiro-OMeTAD) spin-coated to about 200 nm; and a metal back contact, such as gold (Au) evaporated to 80-100 nm. This stack facilitates electron extraction at the ETL-perovskite interface and hole collection at the HTL-perovskite interface, leveraging the favorable band alignment of TiO₂ (conduction band ~ -4.0 eV) and Spiro-OMeTAD (valence band ~ -5.2 eV) with the perovskite's bandgap of approximately 1.5-1.6 eV. Fabrication follows a bottom-up deposition sequence starting from the FTO substrate to ensure sequential layer integrity and minimize contamination. The compact TiO₂ ETL is first applied to prevent direct contact between the perovskite and FTO, reducing recombination; this is followed by perovskite formation via one-step spin-coating of a precursor solution with anti-solvent quenching or two-step sequential deposition of lead iodide and organic halide for uniform, pinhole-free films. The Spiro-OMeTAD HTL is then spin-coated from a chlorobenzene solution, often doped with lithium bis(trifluoromethanesulfonyl)imide and 4-tert-butylpyridine for enhanced conductivity, before thermal evaporation of the Au electrode in a vacuum chamber. This process, typically conducted under controlled humidity (<30%) to avoid perovskite degradation, has enabled certified power conversion efficiencies exceeding 20% in lab-scale devices. A key advantage of the n-i-p configuration is the direct optical path for incident light through the transparent ETL to the perovskite absorber, which minimizes parasitic absorption losses in the HTL and supports high short-circuit current densities (JSC) values often surpassing 22 mA/cm² under standard AM1.5G illumination. However, this architecture is susceptible to UV-induced degradation at the TiO₂-perovskite interface, where photo-generated holes in TiO₂ trigger photocatalytic activity, leading to oxygen vacancy formation and subsequent perovskite decomposition via iodide oxidation and volatile byproduct release, which can reduce device lifetime under prolonged outdoor exposure.

Inverted p-i-n Configuration

The inverted p-i-n configuration in perovskite solar cells features a hole-first architecture, where the device stack begins with a transparent conductive oxide substrate followed by a hole transport layer (HTL), the perovskite absorber, an electron transport layer (ETL), and a rear metal electrode. A typical layer stack is indium tin oxide (ITO)/HTL (such as poly(3,4-ethylenedioxythiophene):polystyrene sulfonate, or PEDOT:PSS)/perovskite (e.g., methylammonium lead iodide, MAPbI₃, or formamidinium lead iodide, FAPbI₃)/ETL (such as [6,6]-phenyl-C₆₁-butyric acid methyl ester, PCBM)/silver (Ag). This arrangement enables top-down deposition, starting from the HTL on the substrate, which contrasts with bottom-up approaches in other architectures. In this setup, photogenerated holes are extracted first at the HTL-perovskite interface, creating an efficient charge separation pathway that aligns with the device's built-in electric field. This hole-priority extraction minimizes the accumulation of mobile ions at the interfaces, thereby reducing ion migration effects that can lead to performance degradation over time. By facilitating rapid hole removal, the configuration suppresses hysteresis in current-voltage characteristics, ensuring more reliable power output measurements compared to configurations with pronounced ionic effects. The inverted p-i-n design offers key advantages for practical applications, including enhanced stability under operational stresses and compatibility with low-temperature processing. It supports ambient air fabrication, avoiding the need for inert atmospheres during HTL and perovskite deposition, which simplifies scalability. Furthermore, the top-down processing is particularly suited to flexible substrates like ITO-coated polyethylene terephthalate (PET), enabling lightweight, bendable devices without compromising efficiency. For instance, inverted cells on flexible ITO/PET have achieved power conversion efficiencies exceeding 20%, while retaining over 90% of initial performance after repeated bending cycles. These attributes position the configuration as a promising pathway for roll-to-roll manufacturing and integration into tandem structures.

Interface and Layer Engineering

Interface and layer engineering in perovskite solar cells (PSCs) involves targeted modifications at the boundaries between layers and within the active material to reduce non-radiative recombination losses, enhance charge carrier extraction, and optimize energy level alignment for improved device performance. These strategies address inherent defects at interfaces, such as trap states that hinder charge transport, by introducing thin passivation layers or compositional gradients that promote efficient carrier collection without significantly altering the overall device architecture. Such engineering has enabled power conversion efficiencies (PCEs) exceeding 22% in various configurations, approaching theoretical limits while maintaining compatibility with scalable fabrication methods. Passivation techniques at the electron transport layer (ETL)/perovskite interface play a crucial role in mitigating defect-induced recombination. Self-assembled monolayers (SAMs), such as iodine-terminated variants, form ultrathin layers that enhance adhesion toughness by up to 50% (from 1.91 J·m⁻² to 2.83 J·m⁻²), reduce hydroxyl-related traps, and minimize voids, leading to improved open-circuit voltages (VOC) from 1.131 V to 1.185 V and PCEs from 20.2% to 21.4%. Alkylammonium halides, like 1,4-butanediammonium iodide (BDAI) or phenylethylammonium iodide (PEAI), deposit as interlayers that suppress phase segregation, lower trap densities (e.g., from 1.18 × 1016 cm−3 to 4.11 × 1015 cm−3), and decrease surface recombination velocities (e.g., from 637.7 cm/s to 149.4 cm/s), resulting in VOC gains to 1.21 V and PCEs up to 22.31% in inverted PSCs. These molecules coordinate with undercoordinated lead sites, passivating ionic defects while promoting uniform perovskite crystallization. Doping the hole transport layer (HTL) with alkali metals, such as cesium (Cs+), rubidium (Rb+), or sodium (Na+), enhances electrical conductivity by incorporating into the perovskite lattice or HTL matrix, thereby reducing defect densities and improving charge carrier mobility. This doping stabilizes the crystal structure, inhibits ion migration, and enlarges grain sizes, which collectively boosts hole extraction efficiency and overall photovoltaic performance, enabling PCEs that rival commercial silicon cells. For instance, alkali incorporation in HTLs like spiro-OMeTAD facilitates better band alignment and conductivity, minimizing energy barriers at the perovskite/HTL interface. Gradient compositions enable precise band alignment engineering, particularly through wide-bandgap buffer layers or compositional grading within the perovskite absorber, to facilitate smooth charge transfer and reduce voltage losses. In wide-bandgap PSCs (e.g., ~1.7 eV), treatments like 4-fluorinated-phenylethylammonium iodide (F-PEAI) create gradients that penetrate the film, suppressing halide segregation and defects while optimizing conduction/valence band offsets at the ETL interface, yielding PCEs of 20.6% and stable operation over 1000 hours under inert conditions. These approaches ensure favorable energy funnels for carriers, enhancing fill factors and VOC without compromising light absorption. Two-dimensional/three-dimensional (2D/3D) heterostructures provide effective defect passivation by forming Ruddlesden-Popper phases at the surface or interfaces, which encapsulate the 3D perovskite core and reduce non-radiative recombination. Gradient engineering with amphiphilic spacer cations diffuses 2D layers into the bulk, passivating both surface and volumetric defects, achieving PCEs of 22.54% with VOC of 1.186 V and fill factors of 0.841, alongside enhanced stability under moisture, light, and thermal stress for over 1000 hours. This heterojunction modulates the energy landscape, improving charge extraction while blocking moisture ingress at vulnerable boundaries.

Stability and Degradation

Intrinsic Degradation Mechanisms

Intrinsic degradation mechanisms in perovskite solar cells arise from the material's inherent structural and electronic properties, leading to performance loss over time even under controlled conditions. These processes fundamentally stem from the ionic nature of halide perovskites, such as methylammonium lead iodide (MAPbI3), which enable high defect tolerance and charge carrier mobility but also promote instability. Key mechanisms include ion migration, phase transitions and decomposition, photo- and thermal-induced trap formation, and hysteretic current-voltage (I-V) behavior, each contributing to reduced efficiency, fill factor, and open-circuit voltage. Ion migration is a primary intrinsic degradation pathway, driven by the high mobility of ionic species within the perovskite lattice under electric fields or illumination. In MAPbI3, methylammonium cations (MA+) and iodide anions (I-) diffuse preferentially along grain boundaries and defects, with activation energies as low as 0.2-0.6 eV for I- migration. This diffusion leads to ion accumulation at electrode interfaces, creating local electric fields that screen applied bias, impede charge extraction, and induce non-radiative recombination. Under forward bias, such migration exacerbates defect growth and hysteresis, while reverse bias can partially reverse the effect, highlighting the reversible yet detrimental nature of this process. Seminal studies have quantified this using impedance spectroscopy, showing ion migration rates increasing exponentially with temperature and bias, directly correlating to operational instability. Phase instability further compounds degradation through structural transitions and chemical decomposition inherent to the perovskite crystal. MAPbI3 exhibits a tetragonal-to-cubic phase transition around 55°C, which disrupts lattice order and increases defect densities, leading to carrier trapping and reduced photovoltaic output. More critically, exposure to trace moisture triggers irreversible decomposition via the reaction: \text{MA PbI}_3 + \text{H}_2\text{O} \rightarrow \text{PbI}_2 + \text{CH}_3\text{NH}_2 + \text{HI} This hydrolytic pathway forms insulating PbI2 precipitates that scatter light and block charge transport, with decomposition accelerating above 70% relative humidity even at room temperature. Atomic-scale simulations reveal that water molecules coordinate with MA+, weakening I-Pb bonds and initiating phase segregation, underscoring the material's sensitivity to its soft ionic lattice. Photo- and thermal degradation mechanisms involve light- and heat-induced alterations to the electronic structure, promoting trap states and reactive species formation. Under illumination, photoexcitation generates electron-hole pairs that facilitate iodide vacancy migration, leading to trap formation at grain boundaries and surfaces; these deep traps increase non-radiative recombination, with defect densities rising from ~1015 cm-3 to over 1016 cm-3 after prolonged exposure. In the presence of oxygen, superoxide (O2-) radicals form via electron transfer from photoexcited perovskites, oxidizing iodide to iodine and decomposing the lattice into PbI2 and volatile organics. Thermally, elevated temperatures (>85°C) induce similar trap emergence through phonon-assisted vacancy formation and partial phase melting, with MAPbI3 decomposing to PbI2 around 150-180°C; this process is exacerbated by thermal expansion mismatches at interfaces, creating microcracks that amplify subsequent degradation. Hysteretic I-V behavior manifests as scan-direction-dependent photocurrents, a direct consequence of the aforementioned mechanisms, particularly ion migration and capacitive charge buildup. During voltage sweeps, slow ion redistribution lags behind electronic transients, causing transient polarization and apparent fill factor variations; faster scan rates minimize hysteresis by limiting ion movement time. The severity is quantified by the hysteresis index (HI): \text{HI} = \frac{\text{PCE}_\text{RS} - \text{PCE}_\text{FS}}{\text{PCE}_\text{RS}} where PCERS and PCEFS are power conversion efficiencies from reverse and forward scans, respectively; HI values near 0 indicate negligible hysteresis, while >0.2 signal significant degradation impact. Capacitive effects from interfacial ion accumulation contribute ~20-50% of the hysteresis amplitude, with the remainder tied to trap filling/detrapping, underscoring how these intrinsic processes distort device metrics and hinder reliable performance assessment.

Strategies for Enhanced Durability

One primary strategy to mitigate extrinsic degradation in perovskite solar cells involves robust encapsulation to shield the active layers from environmental factors such as moisture and oxygen ingress. Glass-glass encapsulation, often combined with ethylene vinyl acetate (EVA) interlayers, forms a multi-layer barrier that prevents water vapor transmission, enabling devices to withstand damp heat tests (85°C, 85% relative humidity) for over 1,000 hours with minimal efficiency loss. Polymer-based barriers, including ultrathin plasma-deposited adamantane or atomic layer deposition (ALD) of Al₂O₃, further enhance protection by creating conformal, pinhole-free coatings that retain over 90% of initial power conversion efficiency (PCE) under prolonged humidity exposure. These approaches directly address ion migration and phase segregation exacerbated by ambient conditions. Composition engineering of the perovskite absorber layer is another critical method to bolster intrinsic thermal and phase stability. Mixed-cation formulations, such as formamidinium (FA⁺), methylammonium (MA⁺), and cesium (Cs⁺) combinations in structures like Cs₀.₁FA₀.₉PbI₃ or FA₀.₈₅MA₀.₁₅Pb(I₀.₈₅Br₀.₁₅)₃, suppress the formation of non-photoactive δ-phase and improve tolerance to temperatures up to 150–200°C by enhancing lattice energy and reducing defect formation energies. Fluorinated cation variants, including partially substituted fluoromethylammonium, further stabilize the crystal structure against thermal-induced decomposition, achieving PCEs above 24% with 95% retention after 1,000 hours at 85°C. These compositions counteract degradation mechanisms like halide segregation under heat or light stress. Incorporation of additives during precursor solution preparation or post-deposition treatment effectively passivates defects and inhibits ion migration. Lewis base adducts, such as bidentate molecules like 2,2'-bipyridine or thiourea derivatives, coordinate with undercoordinated Pb²⁺ sites to form stable complexes that reduce non-radiative recombination and enhance charge carrier lifetimes, leading to improved operational stability under illumination. For instance, low-concentration Lewis acid-base adducts in cesium-based perovskites have demonstrated 92% PCE retention after 1,800 hours at maximum power point tracking. These additives also promote larger grain sizes and smoother films, minimizing pathways for environmental degradation. Recent advancements in 2025 have showcased the efficacy of these strategies in achieving unprecedented durability. Interfacial contact engineering with fluorinated barriers enabled perovskite solar cells to retain 100% of their initial efficiency after 1,200 hours of continuous 1-sun illumination at maximum power point. Additionally, defect-passivation techniques combined with optimized compositions have projected operational lifespans exceeding 30 years based on accelerated aging tests, aligning perovskite performance with commercial silicon benchmarks. These results underscore the potential for lab-scale strategies to translate into reliable, long-term photovoltaic devices.

Toxicity and Environmental Impact

Lead-Based Toxicity Issues

Perovskite solar cells typically incorporate lead in the form of lead halides within the absorber layer, resulting in a lead content of approximately 0.4 g/m² for a standard 300 nm thick layer. This lead is highly soluble in water, with lead iodide (PbI₂) exhibiting a solubility product constant ranging from 8.3 × 10⁻⁹ to 1.84 × 10⁻⁸, facilitating the release of Pb²⁺ ions under aqueous conditions. Such solubility raises concerns about unintended leakage during device operation or damage, potentially contaminating surrounding environments. The health risks associated with lead from perovskites stem primarily from its neurotoxicity, where even low-level exposure can impair neurological development, particularly in children who absorb lead at rates 4-5 times higher than adults. Blood lead levels exceeding 5 μg/dL may cause anemia, while concentrations above 40 μg/dL lead to overt poisoning symptoms, including reduced IQ. In ecosystems, lead bioaccumulates through uptake by plants, with studies on mint showing root concentrations up to 26.9 mg/kg from perovskite-derived sources at 5 mg/kg soil contamination—10 times more bioavailable than natural lead—potentially entering the food chain and amplifying toxicity. Throughout the lifecycle of perovskite solar cells, lead poses risks during fabrication, where waste from precursor materials like PbI₂ can expose workers to inhalation or ingestion hazards, and at end-of-life, where damaged modules under rain exposure may leach significant Pb²⁺ into soil and water. For instance, leaching from a single broken 1 m² panel can elevate soil lead levels to 70 mg/kg, heightening contamination in runoff and posing threats to aquatic organisms and human health via indirect pathways like hand-to-mouth activity in contaminated areas. Regulatory frameworks, such as the EU Restriction of Hazardous Substances (RoHS) Directive, impose strict limits on lead in electrical and electronic equipment, capping it at 0.1% (1,000 mg/kg) by weight in homogeneous materials. While fixed-location photovoltaic modules are exempt, portable or flexible perovskite devices often exceed this threshold—reaching 22,400 mg/kg on polymer substrates—potentially hindering commercialization without temporary exemptions justified by socioeconomic benefits.

Mitigation and Lead-Free Alternatives

Efforts to mitigate lead toxicity in perovskite solar cells have focused on developing lead-free alternatives that maintain viable photovoltaic performance while addressing environmental concerns. Tin-based perovskites, such as methylammonium tin triiodide (MASnI3), represent a prominent lead-free option due to their suitable bandgap of approximately 1.3 eV, which enables efficient absorption of the solar spectrum. However, these materials suffer from instability caused by the oxidation of Sn2+ to Sn4+, leading to p-type self-doping and reduced device efficiency. To counter this, strategies like incorporating antioxidants or hybridizing with germanium (Ge) have been explored; for instance, cesium tin-germanium triiodide (CsSn0.5Ge0.5I3) solid solutions exhibit improved stability and power conversion efficiencies exceeding 9% in all-inorganic configurations. Recent strategies, including interface engineering in tin-based perovskites, have achieved certified PCEs up to 17.71% as of October 2025, while materials like cesium tin-germanium triiodide (CsSn0.5Ge0.5I3) solid solutions exhibit improved stability and power conversion efficiencies exceeding 9% in all-inorganic configurations. As of 2025, tin-based lead-free PSCs have reached certified efficiencies of 16.65% (April 2025) and 17.71% (October 2025), demonstrating substantial improvements in performance and stability. Bismuth (Bi)-based hybrids, including germanium or silver combinations, further expand lead-free possibilities by offering wider bandgaps suitable for tandem applications, though their efficiencies remain lower, typically around 5-7%. Double perovskites provide another avenue for lead replacement, featuring ordered structures that enhance stability without relying on toxic elements. A notable example is Cs2AgBiBr6, which has a bandgap of about 2.0 eV and demonstrates good thermal and moisture resistance, achieving power conversion efficiencies of 3-5% in initial devices. Recent advancements, such as defect passivation and hydrogenation, have pushed efficiencies to over 6%, highlighting its potential for scalable, non-toxic photovoltaics. These materials prioritize environmental safety over matching the high efficiencies of lead-based counterparts, with ongoing research aimed at optimizing charge transport and reducing indirect bandgaps. For lead-containing cells, encapsulation and containment strategies effectively minimize leakage risks. Silica or polymer-based seals, such as sulfosuccinic acid-modified polyvinyl alcohol coatings, can prevent over 99% of lead leakage under simulated rain or immersion conditions by forming robust barriers that withstand mechanical damage. Carbon-based adsorbents integrated into electrodes or encapsulation layers further enhance containment; for example, multifunctional carbon electrodes with lead-chelating additives capture leaked ions, reducing environmental release by adsorbing up to 95% of solubilized lead in damaged modules. These approaches maintain device performance while ensuring compliance with toxicity standards. Reducing lead usage through material engineering also mitigates toxicity without fully eliminating the element. Thinner absorber layers, typically reduced to 200-300 nm from standard 500 nm thicknesses, significantly lower total lead content—by up to 50%—while preserving light absorption through optimized light management techniques like nanophotonic structures. Diluted precursor compositions during fabrication, achieved via co-solvent strategies, enable uniform thin films with minimal lead incorporation, supporting sustainable scaling for commercial modules. These methods balance efficiency, with cells retaining over 20% power conversion, and reduced environmental impact.

Tandem and Advanced Applications

Silicon-Perovskite Tandem Cells

Silicon-perovskite tandem solar cells integrate a wide-bandgap perovskite top cell with a mature silicon bottom cell to capture a broader spectrum of sunlight, thereby overcoming the efficiency limitations of single-junction silicon photovoltaics, which are capped near 29% under the Shockley-Queisser limit. This hybrid approach leverages the tunable bandgap of perovskites, typically around 1.67–1.75 eV for the top cell, to absorb higher-energy photons while allowing the silicon bottom cell (bandgap of 1.12 eV) to utilize lower-energy light transmitted through the top layer, optimizing overall power conversion efficiency. These tandems can be configured in two primary architectures: two-terminal (2T) monolithic designs, where the cells are electrically connected in series via a tunnel junction or recombination layer, or four-terminal (4T) mechanically stacked setups, which operate independently with separate contacts for greater flexibility but added complexity in optical coupling. In 2T configurations, precise current matching between the sub-cells is essential to avoid bottlenecks, often achieved through bandgap engineering of the perovskite and the use of transparent conductive oxides or silicon-based tunnel junctions to facilitate efficient charge recombination without significant voltage loss. 4T designs sidestep current matching issues but require anti-reflective coatings and index-matching layers to minimize parasitic absorption and reflection losses at the interface. Laboratory efficiencies for silicon-perovskite tandems have rapidly advanced, with the certified record reaching 34.85% for a 2T device on a 1 cm² area, achieved by LONGi Solar in April 2025 and verified by the National Renewable Energy Laboratory (NREL). For larger-area cells relevant to commercialization, TrinaSolar reported 31.1% efficiency on a ~22 cm² (210 mm × 105 mm) perovskite-silicon tandem in April 2025, demonstrating scalability toward module-level production. Recent progress includes a certified 33.6% efficient flexible perovskite-silicon tandem in November 2025, advancing applications in lightweight and bendable devices. Key challenges in 2T tandems include fabricating robust tunnel junctions—such as indium tin oxide or doped silicon layers—that maintain low resistance and optical transparency under operational stresses, while ensuring lattice matching and defect passivation at the perovskite-silicon interface to minimize recombination losses. These hurdles, though significant, position silicon-perovskite tandems as a commercially viable path to efficiencies exceeding 35% in the near term.

All-Perovskite and Multi-Junction Tandems

All-perovskite tandem solar cells stack two perovskite subcells with complementary bandgaps to capture a broader portion of the solar spectrum, surpassing the efficiency limits of single-junction devices. Typically, the top subcell employs a wide-bandgap perovskite with an energy gap of approximately 1.8 eV, while the bottom subcell uses a narrow-bandgap perovskite around 1.2 eV, enabling efficient current matching in a two-terminal configuration. Recent advancements have achieved certified power conversion efficiencies exceeding 31%, as demonstrated by SolaEon Technology's device at 31.38% in April 2025. Extending this approach to multi-junction architectures, all-perovskite triple-junction cells incorporate three perovskite layers with progressively narrower bandgaps, such as 1.9 eV, 1.5 eV, and 1.2 eV, to further enhance spectral utilization. Hybrid perovskite/perovskite/silicon triple-junction designs have also emerged, leveraging mature silicon bottom cells for added stability. An international team reported a record efficiency approaching 30% in October 2025 for a large-area triple-junction perovskite cell, with a 1 cm² champion device at 27.06% and retaining 95% efficiency after 400 hours of operation. Similarly, stabilized phase engineering in perovskite/perovskite/silicon triples yielded 28.7% efficiency on 1 cm² areas. These configurations offer key advantages, including solution-processable fabrication that enables low-cost, scalable production without high-vacuum equipment, and lightweight, flexible devices suitable for diverse applications like building-integrated photovoltaics. However, challenges persist in multi-junction stacking, particularly lattice mismatch at interfaces due to polycrystallinity and differing crystal structures among perovskite layers, which can introduce defects and reduce charge collection efficiency.

Recent Efficiency Records in Tandems

In 2025, significant advancements in perovskite tandem solar cell efficiencies were reported, particularly for silicon-perovskite and triple-junction configurations. Trinasolar announced a certified power conversion efficiency (PCE) of 31.1% for a two-terminal perovskite-silicon tandem solar cell in April, achieved through optimized perovskite layer deposition on textured silicon substrates. Later that year, LONGi reported a record 34.85% PCE for a similar two-terminal silicon-perovskite tandem, verified by the National Renewable Energy Laboratory (NREL), highlighting improvements in interface passivation and wide-bandgap perovskite absorbers. For multi-junction tandems, a University of Sydney-led team achieved a breakthrough in October 2025 with a perovskite-perovskite-silicon triple-junction cell, reaching 27.06% PCE on a 1 cm² area and 23.3% on a 16 cm² device, both under standard AM1.5G illumination. This design demonstrated exceptional durability, retaining 95% of initial efficiency after 407 hours of maximum power point tracking (MPPT) operation and passing 200 thermal cycles per IEC 61215 standards (-40°C to 85°C). The high stability stems from nanoscale interface engineering that minimizes non-radiative recombination, marking a key step toward scalable triple-junction tandems. These records were independently verified by authoritative labs, including NREL for the LONGi silicon-perovskite cell and Fraunhofer ISE CalLab for efficiencies up to 28.90% on 1 cm² and 25.2% on 60 cm² perovskite-silicon tandems in September 2025. Fraunhofer ISE's certification process involved precise current-voltage (J-V) measurements under calibrated conditions, confirming reproducibility for larger areas. Simulation tools play a crucial role in optimizing tandem performance, particularly for current matching between sub-cells to maximize short-circuit current density (Jsc). Optoelectronic software like SCAPS-1D enables detailed modeling of layered structures, accounting for optical filtering, carrier transport, and recombination losses. In these simulations, J-V curves are generated to align the Jsc of the top perovskite layer (typically ~1.6-1.7 eV bandgap) with the bottom silicon or perovskite sub-cell, often revealing optimal thicknesses where Jsc plateaus around 20-22 mA/cm² for balanced operation. For instance, SCAPS modeling of perovskite-silicon tandems shows open-circuit voltage (Voc) exceeding 1.8 V, with fill factors above 80%, guiding experimental iterations for current-voltage harmony. Theoretically, double-junction perovskite tandems, such as perovskite-silicon configurations, approach a detailed-balance limit exceeding 45% under one-sun illumination, surpassing the 29.4% Shockley-Queisser limit for single-junction silicon by better utilizing the solar spectrum across bandgaps of ~1.7 eV (top) and 1.1 eV (bottom). This potential, derived from radiative recombination assumptions, underscores the pathway to efficiencies beyond current records through refined bandgap tuning and minimal losses.
RecordConfigurationEfficiency (%)Area (cm²)CertificationDateSource
TrinasolarSi-perovskite (2T)31.1~22Independent labApr 2025Perovskite-Info
LONGiSi-perovskite (2T)34.851NRELApr 2025LONGi
University of SydneyPerovskite-perovskite-Si (3J)27.06 (champion) / 23.3 (large)1 / 16Third-partyOct 2025Nature Nano
Collaborators (e.g., KAUST/Fraunhofer)Si-perovskite25.260Fraunhofer ISESep 2025Nature Comm

History and Commercialization

Historical Development and Milestones

The development of perovskite solar cells (PSCs) began in 2009 when Akihiro Kojima and colleagues, working under Tsutomu Miyasaka at Toin University of Yokohama, first demonstrated the use of methylammonium lead halide perovskites (CH3NH3PbX3, where X = I or Br) as visible-light sensitizers in liquid-electrolyte dye-sensitized solar cells (DSSCs), achieving power conversion efficiencies (PCEs) of 3.81% for the iodide and 2.01% for the bromide variant. This pioneering work highlighted the perovskites' strong light absorption and suitable bandgaps but was limited by poor stability in liquid electrolytes. Michael Grätzel, a leading figure in DSSC research at École Polytechnique Fédérale de Lausanne (EPFL), later collaborated on advancing these materials toward more robust architectures. A major breakthrough occurred in 2012 with the transition to solid-state configurations, eliminating liquid electrolytes to improve stability. Nam-Gyu Park's group at Sungkyunkwan University reported the first solid-state PSC using CH3NH3PbI3 as the absorber and spiro-OMeTAD as the solid hole-transport material, attaining a certified PCE of 9.7%. Concurrently, Henry Snaith's team at the University of Oxford developed a similar solid-state device with a PCE of 10.9%, emphasizing the perovskite's role as a complete light harvester rather than just a sensitizer. These advancements by Park, Snaith, and Grätzel—often credited as foundational contributors—shifted focus to planar heterojunction architectures, enabling rapid efficiency gains. Subsequent milestones accelerated PSC progress. In 2013, efficiencies surpassed 15%, with Burschka et al. at EPFL achieving 15.4% through sequential deposition techniques that improved film quality and coverage. By late 2013, devices reached approximately 16% PCE, as reported by Liu et al. using mixed-halide perovskites and optimized electron-transport layers. The field continued to evolve, with PCEs climbing to over 25% for single-junction PSCs by 2019, exemplified by McMeekin's work on wide-bandgap perovskites that minimized non-radiative recombination. These gains were driven by innovations in composition, such as incorporating formamidinium and cesium to stabilize the alpha-phase perovskite structure. In 2025, scaling efforts marked further industrialization milestones, with UtmoLight in China launching GW-scale production of perovskite modules at its GW-scale facility, targeting 20% efficiency in mass production and achieving 18.1% for 0.72 m² modules through vacuum deposition and solution processing. This progress builds on the foundational R&D by early pioneers like Grätzel, Snaith, and Park, whose interdisciplinary approaches in materials synthesis and device engineering propelled PSCs from lab curiosities to viable photovoltaic contenders.

Current Commercial Efforts and Future Prospects

As of 2025, several companies are advancing perovskite solar cell commercialization through pilot production and strategic partnerships. Oxford PV has initiated commercial sales of perovskite-silicon tandem solar panels, marking the first market entry for this technology following the development of large-area modules with efficiencies of up to 26.9%. In April 2025, Oxford PV entered a patent licensing agreement with Trinasolar to accelerate the deployment of perovskite-silicon tandem technologies, enabling broader industrial scaling. In September 2025, Oxford PV announced the commercial sale of next-generation perovskite-silicon tandem panels achieving 25% module efficiency. Trinasolar achieved a record 31.1% efficiency for a 210 mm industrial-grade two-terminal perovskite-silicon tandem cell in early 2025, and subsequently developed an 800 W+ module prototype using these cells, certified for peak output. UtmoLight commenced GW-scale perovskite solar module production in February 2025 at its facility in China, targeting 20% efficiency in mass production and achieving 18.1% for 0.72 m² modules by March 2025. The market features multiple pilot lines with capacities surpassing 100 MW annually, including 150 MW lines operational in China for tandem perovskite modules. According to IDTechEx forecasts, annual perovskite photovoltaic revenue is projected to reach nearly US$12 billion by 2035, driven by tandem configurations and flexible applications. These efforts reflect a shift from lab-scale demonstrations to initial commercial viability, with over a dozen pilot facilities worldwide contributing to early deployments in regional projects. Key challenges hindering widespread adoption include achieving 20-year stability certification, as current perovskite modules often degrade within one year under operational conditions, far short of the 25-30 years typical for silicon cells. Supply chain limitations for high-purity precursors, such as lead halides and organic components, pose additional barriers due to inconsistent sourcing and elevated costs. Long-term field testing and adherence to IEC standards remain essential to validate durability for certification. Future prospects emphasize integration into building-integrated photovoltaics (BIPV), where semi-transparent and flexible perovskite cells enable energy-generating windows and facades with tunable aesthetics. Cost projections indicate perovskite-silicon tandem modules could reach below $0.30/Wp by 2030, approaching parity with conventional silicon through scaled production and efficiency gains. These advancements position perovskites to capture a significant share of the solar market, particularly in urban and flexible applications.