DOI:
10.1039/D5TA04267D
(Review Article)
J. Mater. Chem. A, 2025, Advance Article
Exploring metal halide perovskites as active architectures in energy storage systems
Received
27th May 2025
, Accepted 14th September 2025
First published on 16th September 2025
Abstract
Metal halide perovskites (MHPs) have emerged as versatile, cutting-edge materials in the field of energy conversion and storage, expanding their influence well beyond photovoltaics to transform technologies such as lithium-ion batteries (LIBs), supercapacitors (SCs), and photo-induced energy storage systems. Initially renowned for their remarkable performance in solar cells, MHPs are now attracting significant attention in energy storage applications due to their outstanding properties, including high ionic conductivity (10−3 to 10−4 S cm−1), long charge-carrier diffusion lengths, tunable band gaps, large surface areas, and structurally flexible lattices. Both lead-based and lead-free variants have demonstrated considerable promise, particularly as electrode materials and in the fabrication of stable artificial solid electrolyte interphases (ASEIs). In addition, the strong light absorption capabilities of halide perovskites have opened pathways toward photo-rechargeable devices, where perovskite solar cells (PSCs) are integrated directly with energy storage systems to enable sustainable and efficient photo-charging. This review provides a concise overview of recent progress in the synthesis and compositional engineering of MHPs, examining how structural and chemical tuning governs their optoelectronic and physicochemical properties. It further explores the emerging applications of perovskite materials in diverse energy storage devices, emphasizing the role of composition in optimizing electrochemical performance. Special attention is given to the integration of PSCs with storage systems as a promising avenue for next-generation multifunctional energy technologies. Finally, the review outlines future opportunities and the key challenges that must be addressed to fully realize the potential of MHPs in high-performance, durable, and scalable energy storage solutions.
1. Introduction
In recent years, the demand for renewable energy sources has increased due to rising pollution from fossil fuels and the ongoing energy crisis. Solar energy is recognized as the most abundant clean alternative to conventional energy generation methods. However, due to their variable availability, the fourth technological revolution calls for innovative energy storage solutions like batteries and supercapacitors (SCs) to enhance the use of electricity generated from these renewable sources.1,2 Currently, lithium-ion batteries (LIBs) represent the most advanced and preferred technology for energy storage solutions in the market. The development of LIB technology began in the late 1980s and early 1990s, and its recognition culminated in the award of the Nobel Prize in 2019. Since then, it has established itself as the benchmark due to its superior energy density (100–265 Wh kg−1) compared to any other battery technology currently available.3 Given their significantly high energy and power density, LIBs have attracted a great deal of interest for their use in portable electronics, but also in battery electric vehicles (BEVs).4 Nonetheless, LIB technology faces certain drawbacks, for instance, capacity deterioration, safety concerns, and inadequate cycle stability.5,6 Given the current limitations, existing LIB technology is unable to achieve the ultimate goal of decarbonizing the world and necessitates additional research. In contrast, SCs offer superior cycle life and high power capability by storing energy through the reversible adsorption of ionic species on highly porous electrodes. However, their widespread adoption is constrained by a comparatively lower energy density, which limits their use in applications requiring long-duration energy storage.7 To address the limitations associated with LIBs and SCs, the development of next-generation LIB technologies offering enhanced longevity, faster charging, improved safety, and cost-effectiveness critically depends on advancements in key battery components, including electrodes, electrolytes, and interfaces. Similarly, achieving higher energy density in SCs requires significant innovation in material design and device architecture. Therefore, researchers are constantly exploring next-generation materials to replace traditional electrode materials. One of the emerging groups of active materials in this field is inorganic and organic–inorganic metal halide perovskites (MHPs), with the general formula ABX3, where A represents a small organic or inorganic cation, B is a divalent metal ion, and X is a halogen anion. MHPs are a specific family of crystalline materials featuring a soft crystal lattice with excellent structural and compositional tunability.8
Over the past decade, MHP materials and their wide-ranging applications have experienced remarkable growth. Their exceptional optoelectronic properties, combined with the ability to control photogenerated ions and electronic charges simultaneously, make them an outstanding material for numerous technological uses. MHPs first gained prominence as the foundation for a new era in photovoltaics, with perovskite solar cells (PSCs) rapidly transforming the renewable energy landscape.9–13 These cutting-edge devices offer simpler fabrication methods, lower production costs, and the potential to surpass silicon-based efficiencies, thanks to their high light absorption coefficients across a broad spectrum. As a result, PSCs not only have the capacity to complement the existing photovoltaic (PV) market but also to challenge the long-standing dominance of crystalline silicon cells, paving the way for new manufacturing paradigms.14 Among the diverse family of ABX3 perovskites, methylammonium lead iodide (MAPbI3) and formamidinium lead iodide (FAPbI3), along with their mixed cation or mixed halide forms, are prototype mono-cationic materials for the absorber layer in PSCs.15–19 The real boom started after the application of these MHPs as active material in solar cells, starting at a power conversion efficiency (PCE) ∼3.8% in 2009 (MAPbI3 in dye-sensitized architecture).20 Since then, intensive material innovation and device engineering have pushed efficiencies beyond 26%, particularly through A-site and X-site compositional engineering strategies. For example, the integration of multiple cations in FAPbI3-based systems like triple cation Csx(MA0.17FA0.83)(1−x)Pb(I0.83Br0.17)3 perovskite in meso-architecture has led to PCE values of ∼21.1% in 2016,19 and more recently, the efficiencies exceeded 26% in a small-area unit cell.18,21 The efficiency progress of PSCs and standard device architectures22 is shown in Fig. 1. A particularly promising direction in this field is the development of flexible PSCs, which have achieved PCEs exceeding 25% (certified at 24.90%), surpassing other flexible solar cell technologies.23 Thanks to their high power-to-weight ratio, flexible PSCs are well suited for use in mobile and space energy systems, as well as in portable functional devices.24 Despite these impressive performance gains, the commercialization of perovskite-based devices is still hindered by insufficient long-term operational stability.25–27 Device instability arises from degradation of both the active and buffer layers, which is linked to intrinsic properties of MHPs, such as lattice disorder, ion migration, and trap-state formation. These degradation processes can be triggered or accelerated by environmental factors (moisture, oxygen) and external stressors, including light exposure, heat, and electric fields. Over the past decade, significant progress has been made in improving PSC stability through various strategies.26,28 However, stability remains a key challenge requiring continued investigation. Furthermore, scaling-up PSCs to large-area devices presents additional difficulties, particularly in achieving high-quality, uniform perovskite films at industrially relevant scales, meaning that the performance of large-area PSC modules still falls short of commercialization requirements. The tunable optical and electronic properties of MHPs, combined with rapid progress in their compositional engineering, have positioned them as promising candidates for a wide range of optoelectronic applications. As a result, their potential has quickly expanded beyond PVs, with high-performance light-emitting diodes (LEDs) leading the way. Perovskite-based LEDs have already achieved electroluminescence external quantum efficiencies exceeding 25% for red and green emissions, while maintaining exceptional color purity.12,29,30 Beyond LEDs, MHPs show strong promise for other advanced devices, including lasers, field-effect transistors, photodetectors, and photocatalysts.31–33 More recently, their unique properties have been explored in next-generation memory devices34–36 and even artificial synapses for neuromorphic computing.37
 |
| Fig. 1 (a) Representation of the ABX3 perovskite building principle, (b) typical architectures of PSC devices, (c) the efficiency gain of PSCs22 Copyright 2024, Nat. Photonics. | |
Not surprisingly, MHPs have also paved their way for a new class of energy storage devices where solar PV systems have been integrated with batteries such as LIBs,38 Zn metal batteries,39 Na-metal batteries,40 LIBs,41 and also SCs.42 This can be attributed to the same properties that make them ideal for PV devices. The initial use of hybrid MHPs (MAPbBr3 and MAPbI3), synthesized through hydrothermal methods as anodes in LIBs, was documented by Xia et al.43 Since that time, various studies have been carried out, including the creation of new MHPs for their use as anodes in LIBs44–46 and their adoption as artificial solid electrolyte interphases (ASEI).47,48 Moreover, their integration in SCs has also gained considerable attention because of exciting properties such as electronic confinement, structural flexibility, attractive performance, and stability.49 For example, recently Riaz and coworkers showed that the CsSnBr3/PANI-based electrode exhibits good cyclic stability (91.6%). They further confirmed the high energy density and power density (37.5 Wh kg−1, 1275.4 W kg−1) of these electrodes at 3.4 A g−1.50
It is no coincidence that new review articles appear regularly in these booming research areas.8,9,12,25,51,52 We aim to provide an original and valuable expert perspective that delivers a concise yet comprehensive overview of recent advances in the preparation and compositional engineering of MHP materials in different forms, with particular emphasis on the factors influencing their optoelectronic and physicochemical properties. It offers a focused perspective on the promising yet still underexplored potential of these materials in energy storage systems. The discussion covers both all-inorganic and hybrid organic–inorganic MHPs, emphasizing their electrochemical performance across different device configurations. Key topics include the use of perovskites as anode materials for LIBs and SCs, the design of ASEIs on lithium metal, and a detailed examination of lithium storage mechanisms in perovskite-based anodes. Recent developments in photo-induced LIBs and SCs are also reviewed. In addition, the article highlights the critical physicochemical properties that underpin the versatility and high performance of MHPs, linking these characteristics to their functional behaviour in energy storage applications. The review concludes with an outlook on future research directions and the remaining challenges that must be addressed to unlock the full potential of MHPs in next-generation energy storage technologies.
2. Fundamental properties of MHP materials
MHPs offer distinct advantages over conventional inorganic semiconductors, particularly their exceptional chemical and structural tunability combined with favorable carrier dynamics and transport properties. By tailoring parameters such as stoichiometry, dimensionality, and nanostructure, key material characteristics, including bandgap energy, crystallinity, and chemical stability, can be precisely engineered. A defining feature of perovskites is ion-migration: although often regarded as a drawback that contributes to performance losses and instability in optoelectronic devices, it can also be strategically exploited to enable novel functionalities in halide-perovskite-based electronics.53 Consequently, detailed investigations of their crystal structures and elemental compositions are essential to fully understand and optimize the unique electrochemical behavior of different halide-perovskites and to identify the most suitable candidates for use as active materials in diverse devices, including LIBs and SCs. This section focuses on the fundamental attributes of MHPs that govern their chemical and optoelectronic properties, incorporating recent progress toward addressing stability challenges that impact device performance. Additionally, various synthesis strategies for MHPs will be discussed, given their decisive influence on the physicochemical characteristics of the resulting materials. Finally, special attention will be given to the interactions between MHPs and Li-ions, which play a critical role in their application within lithium-based energy storage systems.
2.1. Compositional engineering and phase transitions of MHPs
As mentioned previously, the typical three-dimensional (3D) MHPs share the basic chemical formula ABX3, where ‘A’ is a monovalent cation (such as Cs+, CH3NH3+ or MA, HC(NH2)2+ or FA etc.), ‘B’ is a divalent transition metal cation (generally Sn+, Pb+, etc.), and ‘X’ is a monovalent halide anion (Cl−, Br−, I−), forming corner-sharing [BX6]4− octahedra cavities containing the ‘A’ cations as shown in Fig. 2a.29 This family of materials can be further divided into two groups according to the nature of the A-site cation: all-inorganic halide perovskites incorporating inorganic cations and organic–inorganic halide perovskites (or so-called hybrid halide perovskites) incorporating charge-compensating amine cations. For achieving structural stability in the 3D network, the ionic radii of the cations and anions should follow the empirically derived Goldschmidt tolerance factor (t) and octahedral factor (μ).54 Both tolerance and octahedral factor predict the perovskite crystal stability with variations in metal cations and halide anions. The following equation can describe the tolerance factor: |
 | (1) |
where rA and rX are the ionic radii of the A-site cation and X-site halide anion, and rB is the Shannon ionic radius of the B-site cation, respectively, in the ABX3 perovskite halide compounds. The octahedral factor can be expressed by the following equation:55 |
 | (2) |
 |
| Fig. 2 (a) Schematic illustration of the typical ABX3 structure of MHPs. (b) Formability of 3D MHPs as a function of size of A-site cation. (c) ABX3 structure evolution with A-site cation alloying29 Copyright 2020, Adv. Funct. Mater. | |
A stable 3D perovskite structure at room temperature and ambient pressure is generally achieved when the t and μ fall within the ranges 0.813 < t < 1.107 and 0.442 < μ < 0.895, respectively,56 and a highly symmetric cubic structure characteristic of MHPs is typically stabilized only within a narrower tolerance factor range of 0.9 < t < 1.57 Deviation from this range results in distortion of the cubic lattice, leading to lower symmetry structures. The size of the A-site cation plays a crucial role in dictating the structural stability of perovskites. Since the A-site cations reside within the [BX6]4− framework, a cation that is lower than the ideal size causes tilting of the BX6 octahedra. This tilting induces structural distortion, giving rise to lower symmetry phases such as tetragonal or orthorhombic forms. Moreover, if the size of the A-site cation is larger than the ideal size, two-dimensional (2D) structures like Ruddlesden–Popper (RP) phase and Dion–Jacobson (DJ) phase will be formed (Fig. 2b).58 Moreover, when A-site cations are alloyed by small or oversized cations, various crystal structures will be formed, including one-dimensional (1D) structures (Fig. 2c).59,60
2.2. Factors guiding chemical and physical properties of MHPs
The physicochemical properties of MHPs are critical determinants of their functionality in energy storage devices. For example, in traditional LIBs, MHPs with intermediate band gaps (such as CsPbBr3) function effectively as stable electrodes owing to their favorable ion-electron conductivity, while their ability to tolerate defects reduces capacity loss.8 For photo-induced LIBs, MHPs with narrow band gaps (like MAPbI3) perform exceptionally as materials for light absorption, where their high carrier mobility facilitates effective conversion of solar energy to charge, and ambipolar transport allows for concurrent Li-ion storage.61 These advantageous properties stem from the intrinsic electronic structure of MHPs, which grants them high optical absorption, long carrier diffusion lengths, and tunable energy levels. The electronic and optoelectronic properties of MHPs can be influenced by a combination of intrinsic material factors and external stimuli, which in turn determine their performance in these energy storage devices. The key factors are specified more precisely below.
2.2.1 Compositional factors. The physicochemical properties of MHP materials originate from their unique inorganic lattice and the intricate interactions between the inorganic framework and organic components.62 Consequently, structural modification of MHPs achieved by introducing specific cations or anions at various sites within the parent ABX3 perovskite structure has become a widely adopted strategy to tailor their fundamental physical and chemical characteristics. Initially, the prevailing belief was that the role of the A-site cation was to maintain the overall charge neutrality. However, more recent studies have demonstrated that the versatility of the A-site cation indirectly affects the structure–property–performance.63 In contrast, the electronic properties of MHPs are more directly governed by the composition of the B- and X-site ions.64,65 For example, in MPbX3, the deeper region of valence band minima (VBM) consists mainly of p orbitals of the ‘X’ ions and a small contribution from the overlapping of s orbitals of the ‘B’ ions. The edge of VBM consists of anti-bonding states of the s orbitals of ‘B’ ions and p orbitals of ‘X’ ions. Therefore, when different halides are used with different electronegativity, the position of the VBM edge changes, hence the bandgap changes. On the other hand, the conduction band minima (CBM) are mainly determined by the antibonding overlap between the p orbitals of ‘B’ site ions and a small contribution of the p orbitals of ‘X’ site ions.66 In the case of tin-based perovskites (ASnX3), the interactions between the A-site cation and the [SnX3]− sublattice are more complex. Predicting the properties of these materials requires consideration of additional factors such as redox stability and defect chemistry.67,68 The featured electronic structure of MHPs is different from that of conventional semiconductor materials, i.e., gallium arsenide (GaAs), with their band gaps forming in between bonding and antibonding orbitals.69 A detailed discussion on the influence of A-, B-, and X-site doping on the physicochemical properties of MHPs is provided in the subsequent sections.
2.2.1.1 ‘A’ cation. The A-site cation in MHPs plays an indirect but essential role in influencing their electronic properties. By altering the lattice volume and introducing structural distortions to the ideal cubic ABX3 framework, the A-site cation leads to the formation of various non-cubic phases. Although these structural changes have only a limited impact on the electronic band structure, they are critical for understanding and optimizing the optoelectronic performance of MHPs.63,70 A highly symmetric cubic structure tends to exhibit reduced band gaps due to improved packing symmetry. However, as the A-site cation size increases beyond this optimal range (t > 1), the resulting lattice expansion and structural deformation lead to bandgap widening.71 This structural deformation primarily affects ionization energy (IE), while volumetric/lattice expansion has a more pronounced influence on electron affinity (EA). The ionic nature of MHPs makes them susceptible to ion migration, particularly under external electric fields or illumination. Such migration can create crystal defects that degrade device performance through current–voltage hysteresis, phase segregation, and chemical corrosion. A-site engineering has emerged as an effective strategy to mitigate these challenges. For instance, partial substitution of A-site cations with oversized organic ions like guanidinium (Gua+) introduces stabilizing hydrogen bonds with the [PbX6]4− lattice,72 suppressing lattice dynamics and enhancing device stability and performance.73 Additionally, A-site doping can stabilize desirable perovskite phases and address long-standing issues in long-term operational stability, a key barrier to commercialization.74The size of the A-site cation also constrains the formation and stability of the [BX6]4− framework. To date, five monovalent cations are known to form 3D lead-halide perovskites: MA,75 FA75 and Cs+ cations76 and the more recently reported aziridinium (Az)77 and methylhydrazinium (MHy).78 Of these, α-MAPbX3, α-AzPbX3, α-FAPbX3, and α-CsPbX3 can form perovskite structures across all halide compositions. However, α-FAPbI3 and α-CsPbI3 are metastable and tend to transform into the non-perovskite yellow δ-phases under ambient conditions.79 MHyPbCl3 and MHyPbBr3 form B-site distorted perovskite structures,78 while MHyPbI3 crystallizes in a perovskitoid phase similar to δ-CsPbI3.80 To improve stability and tailor the material properties, these A-site cations can be mixed to form alloyed perovskites, as mentioned previously. Moreover, certain large organic cations can be partially incorporated into the 3D perovskite lattice to enhance their optoelectronic performance. Examples include imidazolium (IM+; C3H5N2+),81 dimethylammonium (DMA+; CH3NHCH3+)82 ethylammonium (CH3CH2NH3+; EA),83 guanidinium (C(NH2)3+; Gua),73 and acetamidinium (CH3C(NH2)2+; Aca).84 All mentioned materials exhibit temperature-dependent phase transitions; therefore, this needs to be taken into account during device design.
In parallel, the development of 2D perovskites using even larger organic cations as spacer ligands has gained considerable interest. These materials consist of inorganic [BX6]4− octahedral sheets separated by bulky organic layers, leading to a general composition of (L2BX4)(ABX3)n−1, where ‘L’ represents the organic spacer. Commonly used ligands include butylammonium (BA) and phenylethylammonium (PEA), which impart hydrophobicity and thus enhance moisture resistance. These 2D perovskites exhibit strong photoluminescence, attributed to their quantum well-like electronic structures. However, a significant limitation of these materials in optoelectronic applications is the structural heterogeneity resulting from the coexistence of domains with varying inorganic layer thicknesses. Additionally, their anisotropic charge transport properties further restrict their applicability. The structural diversity and design strategies for 2D perovskites will be discussed in detail in the following section of this review.
2.2.1.2 ‘B’ cation. In MHPs, the B-site is typically occupied by divalent metal cations such as Ge2+, Sn2+, and Pb2+, which possess fully or partially filled 4s, 5s, and 6s valence orbitals, respectively. These ns2 lone-pair electrons contribute significantly to the upper region of the VB, playing a crucial role in determining the VBM and, consequently, the electronic structure of MHPs. The stereochemical activity of these lone pairs strongly influences the electronic properties, and an increase in atomic radius (from Ge to Sn to Pb) has been shown to correspond with a widening of the bandgap. This trend is attributed to the changes in the energy level and activity of the lone-pair states. In 2D MHPs, reduced structural dimensionality allows for a broader range of off-center displacements of the B-site metal cation. As a result, the stereochemical activity of the ns2 lone-pair electrons is more pronounced in 2D Ge-, Sn-, and Pb-based perovskites.65 Goesten and Hoffmann conducted an in-depth analysis of the impact of substituting Pb2+ with Sn2+ or Ge2+, along with halide variation, on the bandgap of CsPbBr3.64 Complementarily, Nishat et al. proposed that a reduction in the B-site cation's atomic radius increases the nuclear electrostatic attraction on valence electrons, thereby raising the IE and EA.85 This stronger binding of valence electrons can lower both the bandgap and bonding energy. Thus, a decrease in atomic radius may result in narrower band gaps, in contrast to trends driven by lone-pair effects. Moreover, the stability of the B-site divalent oxidation state decreases across the series Ge < Sn < Pb, reflecting increasing electronegativity. For example, substituting Pb2+ with Sn2+ results in a significant reduction in IE and a moderate decrease in EA. More recently, Liang et al. demonstrated that B-site substitution with lanthanide or alkaline-earth metal ions enhances lattice cohesion and increases I− migration barriers, offering a more effective stabilization strategy than A-site or X-site doping.86 Beyond divalent substitution, MHP structures can also be tailored through the incorporation of trivalent B3+ cations (e.g., Bi3+, Sb3+) or a combination of monovalent (B+) and trivalent (B3+) cations. In the former case, the classic ABX3 structure is no longer maintained due to charge imbalance, resulting in new stoichiometries such as A3B2X9. These structures lack the corner-sharing BX6 octahedral connectivity, leading to significantly altered optoelectronic properties. In the latter case, double perovskites of the type Cs2B(I)B(III)X6 are formed, preserving the perovskite framework while enabling broader compositional versatility. The trivalent B-site in such double perovskites is often occupied by p-block elements like Bi3+,87,88 Tl3+,89 Sb3+,89 and In3+
90 though examples incorporating d- and f-block elements also exist.91–93 The monovalent B(I) cation can include Li+, Na+, Ag+, Tl+, Cu+, or Au+. Among these, Cu+, Tl+, and Au+ are rarely used due to their strong reducing capabilities, which can destabilize common B3+ species.94 Halide substitution in these systems frequently results in phase segregation. Cl-based perovskites exhibit high structural stability, whereas bromide analogues are typically metastable, and iodide-based compounds are often synthetically inaccessible.95,96 An alternative structural strategy involves vacancy-ordered perovskites, particularly for tetravalent B-site ions. Representative examples include Cs2TiX6
97 and Cs2SnX6,98 where charge balance is maintained by introducing ordered vacancies, offering unique optoelectronic properties while preserving structural stability.
2.2.1.3 ‘X’ anion. The composition of the halide (X-site) in MHPs plays a pivotal role in determining the feasibility of perovskite lattice formation. It enables fine-tuning of key physicochemical properties, including bandgap energy and photoluminescence characteristics (shown in Fig. 3).99 For instance, APbCl3 perovskites exhibit a wide bandgap of approximately 3.0 eV, resulting in white coloration and luminescence in the 400–450 nm range. In comparison, APbBr3 analogues possess a narrower bandgap (∼2.2 eV) with emission centered around 500–550 nm, whereas APbI3 counterparts exhibit the narrowest bandgap (∼1.5 eV), corresponding to emission in the near-infrared range of 800–850 nm. Mixed-halide compositions exhibit intermediate band gaps and emission profiles, which are approximately proportional to their halide content. This compositional flexibility enables the rational design of perovskite materials tailored for specific optoelectronic applications.100 Substitution at the halide site also significantly impacts the electronic structure, particularly the IE and EA of the material. This can be theoretically analysed by examining the trend in halide anion properties, such as increasing atomic radius and lone-pair electron energy levels along the series Cl− (3s2) → Br− (4s2) → I− (5s2). The CBM is primarily governed by the B-site cation p-orbital energy levels, which tend to shift downward as the atomic radius of the halide increases. This trend can be attributed to the quantum confinement effect: as the B–X bond length increases from Cl to I, the electron localization on the B-site is reduced, leading to a lowering of the CBM energy. Simultaneously, the VBM shifts upward across the same series, primarily due to the decreasing electronegativity of the halide ions (Cl > Br > I). These combined effects of halide substitution enable precise control over the electronic band structure of ABX3 perovskites, further enhancing their suitability for a wide range of optoelectronic applications.101
 |
| Fig. 3 Influence of the X-site halogen atom on optical properties of MHP: (a–c) luminescence and absorbance of CsPbX3 materials99 Copyright 2019, Nat Commun., (d) image and absorbance of MAPbX3 crystals, (e) visualization of AzPbX3 powders. | |
2.2.2 Phase transition and external factors. MHPs, owing to their inherently ionic lattice structures, exhibit significant ion migration, dynamic lattice behavior, and a pronounced sensitivity to external stimuli such as temperature, pressure, and redox conditions. These characteristics render both hybrid organic–inorganic and all-inorganic MHPs highly responsive to their environment and often undergo structural phase transitions. Polymorphism is one of the most notable consequences of this structural flexibility, which profoundly influences the electronic structure and, consequently, the optoelectronic properties of MHP.102–105 Structural features considered in the A-site cation engineering and temperature- and pressure-induced phase transitions in selected hybrid organic–inorganic MHPs are shown in Fig. 4. The diversity in crystal structures and associated phase transition temperatures for both lead-based and lead-free MHPs is summarized in Table 1. In general, 3D MHPs adopt a high-symmetry cubic structure at elevated temperatures and undergo sequential phase transitions from cubic to tetragonal to orthorhombic upon cooling. For instance, Hansen et al.106 Employed high-resolution powder neutron diffraction to study the temperature-dependent phase transitions in MAPbI3. The authors observed transitions from the γ- to the β-phase and from the β- to the α-phase at 165 K and 327 K, respectively. Interestingly, recent ab initio quantum dynamics simulations for pristine MAPbI3 demonstrated that the structural deformations induced by thermal fluctuations and phase transitions are on the same order as deformations induced by defects.107 Cs-based halide perovskites offer higher structural stability than their organic analogues; they are also susceptible to structural phase transitions under the influence of external factors. For example, CsPbI3 perovskite exhibits four polymorphs with temperature-induced transitions (Fig. 5a).108 Remarkably, the black phase α-CsPbI3 perovskite exhibits poor structural stability at room temperature and spontaneously transforms to the yellow, photoinactive, non-perovskite δ-phase. Moreover, as shown in Fig. 5b, these distinct phases exhibit markedly different optoelectronic characteristics, including variations in band gap, photoluminescence quantum yield, charge carrier mobility, and carrier lifetime.109 Moreover, the impact of high pressure has also been extensively utilized to deepen the understanding of structure–property relationships of MHPs.110 For example, Kong et al.111,112 investigated high-pressure-induced phase transitions in MAPbBr3. This material experiences a cubic–cubic phase transition from Pm
m (3C) to Im
(3R) at approximately 0.5 GPa, which was attributed to the distortion of the PbBr3 polyhedron (Fig. 5c). The authors also indicated that the narrowest band gap (Fig. 5d), along with the longest carrier lifetime (Fig. 5e), is also observed near the phase-transition pressure of approximately 0.5–0.6 GPa. Their results have highlighted the effectiveness of using pressure to modulate crystal structures, which results in a favorable enhancement of material properties. Additionally, the effect of phase changes on electronic and optical properties of three perovskite phases of CsSnI3, as well as their non-perovskite structure, was studied by D. D. Nematov et al.113 The authors showed that the absorption (Fig. 5f) and photoconductivity (Fig. 5g) of CsSnI3 are enhanced in the infrared and visible light ranges as it transitions from the low-temperature phase to the high-temperature phase. In contrast, the stable yellow phase of CsSnI3, known as the δ-phase, only absorbs short-wavelength light. The high levels of absorption and optical conductivity suggest that all CsSnI3 crystals, which have a perovskite structure, possess excellent spectral characteristics suitable for photovoltaic applications.
 |
| Fig. 4 (a) Structural features considered in A-site cation engineering for organic–inorganic perovskites, (b) temperature- and pressure-induced phase transitions in selected hybrid organic–inorganic MHPs (yellow, blue, and violet colors represent 1D preovskitoid, 3D perovskites, and 3D perovskitoid phases, respectively), (c) representation of the thermally- and pressure-induced structural transformations in δ-AcaPbI3.104 | |
Table 1 Different crystal systems (lead-based and lead-free) of commonly used MHPs in LIBs, their phase transition temperature, and space group
Compound |
Crystal system |
Space group (polytype) |
Temperature (K) |
References |
MASnI3 |
Tetragonal |
Pm m (3C/α) |
293 |
75 |
I4cm (3C/β) |
200 |
MASnCl3 |
Cubic |
Pm m (3C/α) |
478 |
114 |
Rhombohedral |
R3m |
350 |
Monoclinic |
P1c1 |
318 |
MASnBr3 |
Tetragonal |
P4mm (3C/β) |
|
115 |
MAPbBr3 |
Cubic |
Pm m (3C/α) |
298 |
116 |
Tetragonal |
I4/mcm (3C/β) |
220 |
117 |
P4/mmm (3C/β) |
150–155 |
118 |
Orthorhombic |
Pna21 |
|
118 |
MAPbCl3 |
Cubic |
Pm m (3C/α) |
200 |
116 |
Tetragonal |
P4/mmm (3C/β) |
173–179 |
118 |
Orthorhombic |
P2221 |
< 173 |
118 |
FAPbI3 |
Cubic |
Pm m (3C/α) |
420 |
75 |
Hexagonal |
P63 (2H/δ) |
270 |
75 |
Trigonal |
P3 (2H/δ) |
150 |
75 |
FAPbBr3 |
Cubic |
Pm m (3C/α) |
275 |
119 |
Tetragonal |
P4/mbm (3C/β) |
175 |
Orthorhombic |
Pnma (δ) |
100 |
FASnI3 |
Cubic |
Pm m (3C/α) |
275 |
119 |
Tetragonal |
P4/mbm (3C/β) |
175 |
Orthorhombic |
Pnma (3C/γ) |
100 |
FASnBr3 |
Cubic |
Pm m (3C/α) |
275 |
120 |
Tetragonal |
P4/mbm (3C/β) |
175 |
Orthorhombic |
Pnma (3C/γ) |
100 |
CsPbI3 |
Cubic |
Pm m (3C/α) |
593 |
75 |
Orthorhombic |
Pnma (δ) |
293 |
121 |
CsPbBr3 |
Cubic |
Pm m (3C/α) |
433 |
122 |
CsSnI3 |
Cubic |
Pm m (3C/α) |
500 |
123 |
Tetragonal |
P4/mbm (3C/β) |
380 |
Orthorhombic |
Pnma (3C/γ) |
300 |
CsSnBr3 |
Cubic |
Pm m (3C/α) |
298 |
124 |
CsSnCl3 |
Cubic |
Pm m (3C/α) |
413 |
114 |
CsGeI3 |
Cubic |
Pm m (3C/α) |
533 |
125 |
CsGeBr3 |
Cubic |
Pm m (3C/α) |
538 |
125 |
 |
| Fig. 5 (a) Variable-temperature phase transitions among the CsPbI3 polytypes,108 (b) illustration of variation of bandgap with the phase transition of CsPbI3 crystal109 Copyright 2018, ACS Nano. (c) Pb–Br inorganic frameworks of MAPbBr3 for low-pressure Pm m and high pressure Im phases. The high-pressure phase in MAPbBr3 exhibits the characteristic elongation of the lead–halide octahedral, together with smaller lead–halide–lead bond angles. (d and e) Demonstration of the band-gap narrowing and carrier-lifetime prolongation in MAPbBr3 at mild pressures, respectively.111 Copyright 2016, Proc. Natl. Acad. Sci. U. S. A. (f) and (g) Calculated absorption coefficient and optical conductivity of α-, β-, γ-, and δ-phases of CsSnI3 as a function of photon energy in the X direction.113 Copyright 2025, J. Electron. Mater. | |
In general, most perovskite and perovskitoid materials undergo no more than five distinct phase transitions before decomposing at elevated temperatures due to the volatility of A-site constituents or amorphization under high pressure.103,126,127 Notably, δ-AcaPbI3 (Fig. 4) represents a unique case with the most extensive polymorphism reported among MHPs, exhibiting at least nine distinguishable polymorphs. Of these, four occur under varying temperature conditions at ambient pressure, while five emerge under high-pressure conditions at room temperature. All identified phases are variations of distorted 2H polytypes. Interestingly, δ-AcaPbI3 also undergoes partial amorphization under high pressure, followed by recrystallization due to Pb atom displacement along the c-axis.104 Beyond polymorphism, ABX3 perovskites can also exhibit polytypism, which is a specific form of polymorphism characterized by variations in the stacking sequence of otherwise identical layers.15 Polytypes combine crystal sub-units of α/3C and δ/2H phases and are commonly observed in systems incorporating oversized A-site cations, such as Aca,104 DMA,82 as well as in mixed FA perovskites.15 For example, hybrid organic–inorganic lead halides can form polytypes within various crystal systems, including hexagonal (2H, 4H, 6H or 8H), cubic (3C), and rhombohedral (3R, 9R).15,128 Gratia et al.128 noted that hexagonal polytypes emerge as transitional products on the surface of thin films of mixed-ion perovskites during the annealing process, unveiling a crystallization polytypic sequence of 2H–4H–6H–3C. The perovskite structure can be cut into layers by the addition of elongated organic BA or PEA cations, leading to the formation of 2D perovskite phases.129
2.2.3 Dimensionality. By leveraging the diverse structures and compositions of MHPs, the dimensionality of MHPs can be precisely adjusted, ranging from 3D to various low-dimensional forms such as quasi-2D, 2D, 1D, and zero-dimensional (0D) configurations. There are two types of low-dimensionalities, one is the “structure-level” and “material-level”.130 The “structure-level” low-dimensional aspect highlights the different morphologies and typically refers to nanostructures like nanosheets, nanowires, and nanocrystals (NCs) (Fig. 6a). In contrast, the “material-level” low-dimensional aspect of perovskite focuses on the fundamental structure where the [BX6]4− octahedra are interspersed with large dielectric spacer molecules, resulting in a bulk formation of atom-level 0D clusters, 1D quantum wires, or 2D quantum wells (QWs) (Fig. 6b). The crystalline perovskite nanostructures with reduced dimensionality show distinctive optoelectronic properties with quantum-confined effect compared with their bulk counterparts (Fig. 6c).131 Due to this phenomenon, the bandgap energy of low-dimensional perovskites is greater than that of their bulk forms. This is accompanied by a variation in the geometric size of perovskite nanostructures, which further alters the bandgap. As depicted in Fig. 6d, the photoluminescence emission wavelengths were systematically adjusted through the gradual decrease in the size of FAPbI3 NCs.132 Consequently, the calculated bandgap energy rose from 1.5 eV in bulk material to more than 1.7 eV for perovskite NCs with an edge length of approximately 8 nm (Fig. 6e). In addition to the hybrid perovskite NCs, inorganic perovskite NCs and quantum dots (QDs) also exhibit bandgap energies dependent on size or diameter, such as CsPbI3 and CsPbBr3.133,134 This size effect can also be anticipated in 1D perovskite nanowires (NWs), where the diameter determines the bandgap energy and emission wavelength independent of the length.134 Furthermore, in 2D perovskites, the octahedral layers are situated between the organic spacers. These organic compounds provide extra functionality, such as a tunable quantum well structure that can be modified by varying the length and type of the organic chain. Fig. 6f presents the photoluminescence spectra of 2D BA2(MA)n−1PbnI3n+1 perovskites, which depend on the number of layers, demonstrating a significant wavelength tuning range from approximately 520 nm to 650 nm (∼130 nm difference, corresponding to a shift of about ∼0.47 eV).135 A comparable layer-dependent pattern was observed in 2D perovskites composed of three cations that include both organic (PEA, MA) and inorganic (Cs) components (Fig. 6g), revealing a consistent decrease in bandgap energy as the number of layers increases, with an energy difference of up to ∼0.6 eV between single-unit-cell and five-unit-cell crystals.136
 |
| Fig. 6 (a) Schematic representation of “material-level” and (b) “structure-level” 3D bulk, 2D, 1D, and 0D perovskite materials130 Copyright 2020, InfoMat, and (c) is the corresponding density of states versus the energy Eg131 Copyright 2021, Small. (d) Photoluminescence (PL) spectra of FAPbI3 nano-crystals (NCs) with different sizes. (e) Bandgap energies versus the edge length of FAPbI3 NCs132 Copyright 2021, Small. (f) PL of exfoliated monolayers for 2D layered BA2(MA)n−1PbnI3n+1 Ruddlesden–Popper perovskites of n = 1 to 4 homologues135 Copyright 2018, Nat. Mater. (g) The bandgap of PEA2A1.5Pb2.5Br8.5 (A = MA and Cs) perovskites with different numbers of layers136 Copyright 2018, Nat. Commun. | |
2.3. Preparation of MHPs and MHP-based batteries and supercapacitors
2.3.1 Preparation of MHPs. Preparation methods of MHPs, alongside the chemical composition, are one of the key factors influencing the physicochemical properties of the title materials. To date, numerous techniques for obtaining MHPs in the form of bulk materials,137 nanocrystalline,31,138,139 single-crystal materials,100 and thin film13,100 have been reported. These processes have been applied, and they can be categorized as occurring in liquid, gas, and solid phases. Selected methods for the preparation of perovskite materials are presented in Fig. 7. Solution-based synthesis of bulk or single-crystal MHPs remains one of the most widely employed methods for preparing perovskite materials (Fig. 7a, c and d); however, it is constrained by the limited availability of suitable solvents. In most cases, highly ionic perovskite precursors are dissolved in toxic solvents such as dimethylformamide (DMF), often in combination with an antisolvent to promote crystallization. Other solvents capable of dissolving perovskite precursors, including N,N-dimethylacetamide and N-methyl-2-pyrrolidone, are also recognized as toxic.140 Although the less hazardous dimethyl sulfoxide (DMSO) can effectively dissolve metal halides. However, due to its high boiling point, DMSO is difficult to dispose of and poses challenges for large-scale processing.141
 |
| Fig. 7 Selected methods of the preparation of various forms of MHP materials, including (a, c and d) solution-based synthesis, (b) solid-state mechanochemical synthesis, (e) one-step and (f) two-step deposition, (g) vacuum evaporation, and (h) hybrid evaporation-solution approach. | |
Furthermore, the solution-based approaches for the preparation of MHPs may be considered as relatively versatile, they are also limited by substrate solubility and long-term stability. In 2015, the Lewiński group made a groundbreaking discovery, describing a new pathway for the production of perovskite materials through a mechanochemical process (Fig. 7b), involving solvent-free grinding of substrate powders in a ball mill.142 This solvent-free synthetic approach is very attractive to curtail the chemical waste generation and simplify the preparation process. The solid-state mechanochemical preparation of MHPs enables the production of high-purity material in a short time and with low energy input, facilitating the use of precursors that are difficult to dissolve, such as CsCl143 or AgCl.95 Hence, the mechanochemical synthesis offers a highly flexible platform for doping engineering at different atomic sites.137 Subsequently, the MHPs derived from mechanochemical synthesis were successfully used for the fabrication of PSC, which contributed to an approximately 10% relative improvement in cell efficiency and operational characteristics compared to devices based on analogous solution-derived perovskite materials, and resolved the issue of the long-term storage of perovskite materials.142,144 Later on, the solution of PSC properties was attributed to the lower density of trap states within materials, due to the higher quality of MHP grains on the surface.145 Mechanochemical synthesis offers several advantages, including precise control over stoichiometry, improved reproducibility, enhanced stability, and higher phase purity in the resulting mechano-perovskite materials. It also enables more feasible large-scale production.146 Although this method is fast, efficient, and environmentally friendly, relying on energy generated through grinding, shearing, and compression, the underlying mechanisms remain relatively poorly understood.
In the process of preparing perovskite thin films for various devices, deposition,100,147,148 and patterning13 are often key steps. Regarding wet methods, a further complication arises from the solvent-mediated growth of polycrystalline perovskite thin films, which can be carried out through either a one-step or two-step deposition approach. The one-step coating usually involves pre-mixed perovskite solution (e.g., PbI2 and MAI) followed by anti-solvent dripping, and the two-step sequential procedure,149 one of the main methods used for depositing perovskite films for various applications,150 which involves deposition of an inorganic layer before converting the film to perovskite by reacting with organic solution (Fig. 7e and f, respectively). By altering synthesis conditions, different nucleation and growth behaviors occur, which in turn directly shape the microscopic morphology of perovskite films, then change their optical and electrical properties.151 These dominant deposition techniques typically use DMF-based precursor solutions, which remain the dominant deposition technique, yet they face the same environmental and health concerns associated with toxic solvents.152 DMSO, while a safer alternative, is highly hygroscopic, often resulting in poor film crystallinity and the formation of numerous point defects.141 Moreover, PSCs fabricated using DMSO-mediated synthesis generally exhibit lower efficiencies than those produced with DMF.141,153,154 Consequently, solvent engineering to achieve high-quality perovskite films remains a major challenge.155 To overcome the cost and scalability issues associated with the low-humidity requirements for device fabrication, studies are now investigating high-humidity conditions. The goal is to better understand perovskite crystal growth and to develop more robust, stable perovskite films.147
The solvent-mediated growth of polycrystalline MHP thin films suffers from critical limitations, including sensitivity to processing conditions, poor control over solvent-mediated crystallization, low reproducibility between laboratories, and restrictions on film size. To address these drawbacks, vacuum evaporation has emerged as a rapidly advancing solvent-free fabrication route (Fig. 7g).156,157 This method enables uniform deposition of precursors and precise control over film thickness. Nevertheless, the power conversion efficiencies of vacuum-evaporated PSCs still lag behind those of high-quality solution-processed devices.158,159 Recently, a hybrid evaporation-solution approach has been proposed as a promising alternative (Fig. 7h). This technique involves evaporating metal halide precursors onto a substrate, followed by spin-coating an organic halide precursor solution to complete film formation.157,160–162 This hybrid strategy combines the advantages of both techniques, offering improved film quality and greater processing flexibility, and may represent a viable route toward scalable, high-performance perovskite device fabrication.
2.3.2 Preparation of MHP-based batteries and supercapacitors. An in-depth understanding of the crystal structure and elemental composition of MHPs is crucial to explain the unique electrochemical behaviours observed in various MHP-based LIBs.43 This disparity underscores the importance of examining the influence of chemical composition on Li-ion storage properties. It is also important to note that both MHP-based batteries and SCs are fabricated by ISO standards. Typically, the electrode fabrication involves preparing a slurry composed of the active material (MHPs), a binder (polyvinylidene fluoride, PVDF), and a conductive additive (e.g., hard carbon) in a particular weight ratio (most commonly 80
:
10
:
10). This mixture is dissolved usually in N-methyl-2-pyrrolidone (NMP) and cast to form the anode. While this method is widely regarded as the “gold standard” in LIB manufacturing, initially developed for chemically robust materials such as oxides, it poses challenges when applied to MHPs. The soft crystal lattice of MHPs is prone to dissolution in polar organic solvents, which can compromise the phase integrity during processing. The MHP solubility is also intentionally used in solution-based methods for producing perovskite solar cells, where solvent mixtures such as DMF
:
DMSO enable the formation of uniform thin films. Although this property does not preclude the use of MHPs in LIBs, it necessitates careful control over processing conditions to preserve phase purity. As a result, alternative fabrication routes such as solid-state approaches, including mechanochemical synthesis, are being explored to address these challenges and enhance the stability of MHP-based electrodes during device fabrication.146,163One of the biggest challenges in current perovskite processing is the need for extremely dry and inert conditions. This is because MHPs are quite sensitive to moisture and oxygen. The fact that we can't carry out solid-state processing in regular environments really drives up both the initial and ongoing costs, especially when it comes to large-scale production of active materials. The use of gloveboxes, dry rooms, and specialized equipment for controlled atmospheres adds significantly to these expenses. This financial hurdle highlights the importance of future research not just in enhancing material stability, but also in creating moisture-tolerant MHP compositions and scalable fabrication methods that can work in normal conditions. At the moment, there aren't any thorough studies looking into the socio-economic factors of using MHPs in LIBs. This is mainly because the use of MHPs in LIBs is still in its early stages of development, which is making it tough to accurately assess material and processing costs. However, similar to the early forecasts for perovskite solar cells, there's hope that MHP-based LIBs could become cost-effective as technology advances. This optimism stems from the relatively low cost of precursors, like lead salts, and the possibility of low-temperature processing.164,165 While MHPs aren't quite ready to be used as commercial anode materials yet, their unique characteristics, such as mixed ionic–electronic conductivity, the ability to tune their composition, and high dielectric constants, point to a promising future in energy storage technologies.
2.4. Mechanism of lithium interaction with MHPs
The LIB utilizing MHPs (MAPbBr3 and MAPbI3) as an anode was first reported in 2015 by Xia et al.43 The authors described good electrochemical performance for MAPbBr3 with a first discharge capacity of 331.8 mAh g−1 at a current density of 200 mA g−1, which is six times of the maximum theoretical capacity (55.96 mAh g−1) if it is assumed that one Li-ion could intercalate per formula unit. This significant difference in electrochemical performance suggests a conversion reaction or another effect that is at play, since six Li-ions per formula unit would otherwise be required.166 Despite these impressive performance metrics, the key mechanism by which Li-ion is intercalated in MHPs remains unknown. Initially, two mechanisms were proposed for Li uptake in perovskites, which are indicated below:
(1) Intercalation mechanism: Vicente et al.167 suggested that there is topotactic Li+ insertion into perovskites without severe structural alterations, as shown in Fig. 8a. The XPS analysis indicated that no Li–Pb alloying reactions occur (Fig. 8b). However, it was unclear which species are reduced because of Li intercalation, given that the Pb and Br electronic structures remain mostly unchanged.
 |
| Fig. 8 (a) Crystal structure of organometal halide perovskite MAPbBr3 indicating multiple Li-ion interaction. (b) Pb 4f and Br 3d core-level XPS signal of MAPbBr3 anodes at different states: pristine, 0.10 V (discharge), and 1.80 V (subsequent charge)167 Copyright 2017, J. Phys. Chem. Lett. (c) Schematic illustration of conversion and intercalation mechanism occurring in organic–inorganic MHPs166 Copyright 2017, ACS Energy Lett. (d) The recorded discharge profile, where each stage is highlighted in a different color: first, the pure material (red), and lithiated phase (red); second, the conversion stage (orange), and, third, the alloying stage (blue)168 Copyright 2019, ChemElectroChem. (e) Schematic illustration of the structural evolution process of CPB/o-CNT composite upon cycling44 Copyright 2025, Energy Storage Mater. | |
(2) Conversion mechanism: in the same year, using electrochemical and powder X-ray diffraction techniques, combined with DFT, Dawson et al., proposed that both intercalation and conversion reaction occurs during the charge/discharge cycle. The authors indicated that for all three hybrid perovskites, MAPbX3 (X = Cl, Br, I), the energy for conversion reactions is more favorable than the intercalation reactions. Some main features of their results are indicated below.166
(a) There are two possible sites for Li-ion insertion in MAPbX3 (X = I/Br/Cl), octahedral and tetrahedral. Also, the site preference depends on both material and Li-ion concentration.
(b) At low Li-ion concentrations, insertion into MAPbI3 becomes more favorable because it is easier to insert Li into larger MAPbI3, causing less distortion in PbI6 octahedra.
(c) Upon full Li intercalation (x = 1.0), the PbX6 octahedra displays structural distortion in all three materials, with MAPbBr3 and MAPbCl3 exhibiting the greatest structural distortion. This may indicate the presence of conversion (to lithium halides (LiX) and Pb metal) or decomposition reactions upon full Li-ion interaction.
(d) The Li-ion conversion process is more energetically favorable than the Li-ion intercalation.
In 2019, Vicente et al.168 conducted another study to understand the Li uptake mechanism in MAPbBr3 using operando-XRD analysis in conjunction with galvanostatic lithiation (discharge) and delithiation (charge) steps (Fig. 8c and d). The authors identified three reaction stages (lithiated phase, conversion, and alloying) associated with varying Li-ion molar content. However, the mechanism of phase transition during the subsequent cycles was not established. For all-inorganic halide perovskites, ex situ XRD patterns of the CsPbCl3 electrode indicated that the preferred mechanism for Li-ion storage in the initial discharge cycle is through conversion reactions rather than Li-ion intercalation.45 As a result, the disputed and vague findings from the ex situ characterizations obscure the understanding of the structural changes of halide perovskites throughout the lithiation and delithiation process. Recently, X.-H. Wu et al.44 utilized in situ X-ray diffraction (XRD) and electrochemical impedance spectroscopy (EIS), along with ex situ characterizations, to explore the intricate mechanisms of lithium storage and release in CsPbBr3, including the phase transition that occurs during the first three cycles. The findings indicate that at full discharge, CsPbBr3 breaks down into CsBr, LiBr, Pb, and Li22Pb5 phases via intercalation–conversion–alloying reactions, followed by the regeneration of CsPbBr3 during the charging process (Fig. 8e).
However, several studies have reported different Li-ion transport mechanisms and SEI formation (shown in Table 2); the precise mechanistic pathways are still a bit of a mystery in this field. This uncertainty largely stems from the intricate structural changes that MHPs undergo during cycling, the diverse range of perovskite compositions (such as lead-based, tin-based, and double perovskites), and the absence of systematic in situ characterization under realistic battery conditions. Therefore, this field of research requires further investigation.
Table 2 Li-transport mechanisms, SEI composition & evolution of different MHPs
Perovskite material |
Electrolyte |
Li storage mechanism/proposed reaction |
Initial SEI characteristics |
SEI evolution |
Ref |
CsPbCl3 |
1 M LiBF4 in PVDF-HFP + BMIMBF4 (1 : 3 w/w) |
Conversion → 2Li + CsPbCl3 → CsCl + 2LiCl + Pb; subsequent Li–Pb alloying (LixPb) during deeper discharge |
Likely contains LiCl, CsCl, Li–Pb alloys, solid electrolyte matrix |
Ex situ XRD after 10 cycles: loss of CsPbCl3 peaks; intensification of Pb + LixPby → conversion + alloying dominant |
45 |
CsPbBr3 |
1 M LiPF6 in EC : DEC : DMC (1 : 1 : 1) with 5% FEC |
(1) Intercalation → conversion→ alloying CsPbBr3 + 6.4Li+ + 6.4e− → CsBr + Li22Pb5+ 2LiBr (discharged) |
Likely contains FEC-derived FEC-rich SEI, LiBr, CsBr, LiPb alloys, and Pb residue |
Ex situ XRD and XPS after long-term cycling; reduced content of CsPbBr3 perovskite; pulverization → CsBr/Pb nanoparticles on CNTs |
44 |
(2) Partial reformation of the CsPbB3 on charge xCsBr + Li22Pb5 + 2xLiBr → xCsPbBr3 + (22 + 2x)Li+ + (22 + 2x)e− + (5 − x)Pb (residue) (charged) |
CsPbI3 |
1 M LiPF6 in EC/DMC (1 : 1) |
Intercalation forming LixCsPbI3 at higher V; deeper discharge yields conversion and Li–Pb alloying |
Probably contains LiF, LixPFy, LixPFyOz from salt decomposition; Li-alkoxy species (CH3OLi, (C2OLi)2) from EC/DMC reduction; possible Cs/Li iodides |
SEI forms in the first cycle and stabilizes; post-100-cycle XRD/SEM: δ-CsPbI3 peaks retained |
46 |
MAPbBr3/MAPbI3/MAPbIBr2 |
1.0 M LiPF6 in EC/DEC (1 : 1) vol ratio) |
Li intercalation followed by Pb(II) → Pb(0) reduction and conversion to LiX + Pb + LixPb alloys (LiPb, Li2.6Pb, Li4.4Pb, etc.), plus SEI formation (LiF, Li2CO3, organics); below ∼0.6 V, alloying–dealloying dominates |
LiF, Li2CO3, alkoxides; Pb/LixPb in SEI; possible LiI residues |
Pb, Li–Pb alloys, and halide products may remain embedded in SEI after 10 cycles |
169 |
MANiCl3 |
1.0 M LiPF6 in EC/DMC ((1 : 1) vol ratio) |
Intercalation into MANiCl3 lattice → conversion to MACl + LiCl + metallic Ni → alloying to Li–Ni alloys (LixNiy) + MACl |
SEI containing LiF, LiCl, and organic carbonate species from electrolyte decomposition |
May form thicker SEI after cycling; increased RSEI; chloride salt and Ni-rich interphase |
170 |
Cs4PbBr6 |
LiPF6 in EC/DEC (1 : 1) |
Cs4PbBr6 + LiPF6 → CsPF6 + PbBr + LiBr; CsPF6 unstable → Cs+ ions released; Cs+ participates in SEI formation (self-healing electrostatic shielding, SHES) to suppress Li dendrites; Pb2+ undergoes partial alloying with Li, not fully reversible → capacity fade |
Likely contains LiF, LiBr, Cs-containing species (CsPF6 decomposition products); Pb/Li–Pb alloy fragments; organic electrolyte reduction products |
SEI may become Cs-rich; the SHES mechanism promotes uniform Li deposition; Pb centers are trapped in the organic matrix after irreversible alloying |
171 |
Bz–Pb–I (1D hybrid) |
1 M LiPF6 in EC/DMC (1 : 1) + 5% |
Multi-step alloying resulting in metallic Pb and different LixPb alloy formation (Li8Pb3, Li22Pb5, LiPb, Li10Pb3, etc.) |
SEI is LiF-rich via FEC; halide residues; organic matrix interfaces |
— |
172 |
3. MHPs in LIBs
Due to the limited theoretical capacity, the risk of short circuits from dendrite formation, and the inadequate ability to sustain continuous high current discharge, currently available graphite anode materials present significant challenges. Consequently, there is a growing need for alternative anode materials that offer higher theoretical capacities, improved cycle stability, and greater cost efficiency.173,174 In this context, MHPs have recently gained attention as potential anode materials for LIBs. Several researchers have investigated different MHP compositions to develop high-performance anodes capable of replacing conventional graphite-based systems. Beyond their application as anode materials, MHPs have also been employed as ASEI layers on Li-metal anodes (LMAs) to mitigate detrimental side reactions with liquid electrolytes. This section aims to provide a comprehensive overview of the application of various MHPs in LIBs, focusing on their roles as both active anode materials and protective interfacial layers.
3.1. MHP-based anodes in LIBs
3.1.1. 3D MHP-based anodes. As mentioned earlier, the first use of 3D MHPs as anode materials was proposed by Xia et al.43 In 2015, where hydrothermally synthesized MAPbBr3 and MAPbI3 microcrystals displayed a first discharge capacity of nearly 330 mAh g−1 and 50 mAh g−1 at 200 mA g−1 current density (Fig. 9a). Moreover, the cycle stability of MAPbBr3 shows prominent improvement as compared to MAPbI3 (Fig. 9b). While the specific reasons for the enhanced performance of MAPbBr3 based batteries remain unclear, this study suggests that halide perovskites hold promising potential for lithium storage applications. Several years later, research also indicates that the charge/discharge capacity can be enhanced when I− is replaced with Br (MAPbBr3), as illustrated in Fig. 9c. However, an incomplete substitution of Br− (MAPbIBr2) leads to a decrease in charge/discharge capacity, as the lattice parameter shifts from tetragonal MAPbI3 to cubic MAPbBr3. This suggests that perovskites with varying halogens exhibit different specific capacities, primarily due to two factors: (1) the influence of halogens on the crystal structures, as we covered in part 3.1, and (2) the competition between the lower atomic mass of Br− and the larger lattice parameter of I−. The effect of crystal morphology on the electrochemical performance of MAPbBr3 was investigated by Q. Wang et al.175 In the study, they compared five different microcrystal sizes (∼2.9 mm, ∼2.4 mm, ∼1.9 mm, ∼1.5 mm, and ∼1.2 mm) and concluded that 1.2 mm-sized MAPbBr3 composite electrodes exhibit unexpectedly high cycling stability for more than 1000 cycles when tested as LIB anodes compared to other-sized samples due to high crystal quality and improved electrical conductivity.
 |
| Fig. 9 Electrochemical properties of the LIBs based of MAPbBr3 and MAPbI3: (a) charge/discharge curves, (b) cycle stability43 Copyright 2015, Chem. Commun. (c) Charge/discharge curves of MAPbI3, MAPbIBr2, and MAPbBr3 169 Copyright 2018, Inorg. Chem. (d) The cycling performance of CPB/o-CNT electrode at 0.1 A g−1 divided into three regions44 Copyright 2025, Energy Storage Mater. (e) SEM image of cubic CsPbCl3. CsPbCl3–graphite-based dual-ion batteries (f) galvanostatic CD curves at various current densities and (g) discharge capacity retention over 40 CD cycles at a current rate of 100 mA g−1 Copyright 2020, Phys. Rev. Appl.45 | |
As compared to organometallic halide perovskites, all-inorganic halide perovskites showed better stability, ultrahigh photoluminescence quantum yield, etc. In this regard, Jiang et al. proposed all inorganic CsPbBr3 as an active material for the LIB anode in 2017 with a first charging capacity of 94.8 mAh g−1 and the cyclic life of 32 rounds at 60 μA cm−2.176 In another report, the electrochemical performance and storage capacity were further enhanced by incorporating carbon nanotubes (CNTs), resulting in improved stability and rate capability due to the pseudo-capacitive effect.177 Moreover, X.-H. Wu et al.44 displays “negative fading” effect and a significant increase in capacity in CsPbBr3@CNT based electrodes. This electrode delivered a specific capacity of 630 mAh g−1 at 0.1 Ag−1 after 200 cycles (Fig. 9d). In addition to bromide, cubic CsPbCl3 (Fig. 9e) has been developed as an anode for LIB and dual-ion batteries. The findings indicate that the half-cell LIB exhibited specific discharge capacities of 612.3, 508.7, 362.4, and 275.2 mAh g−1 at varying current densities of 50, 100, 200, and 250 mA g−1, respectively (Fig. 9f), along with an average coulombic efficiency of 88%. Moreover, the combination of a 3D perovskite anode with a graphitic cathode provided insights into dual-ion batteries operating within a voltage range of 0–4.0 V, averaging 2.53 V, as illustrated in Fig. 9g. The summary of the performance of different MHPs used in LIBs is summarized in Table 3.
Table 3 Summary of different MHPs based on LIBs and their performancea
Perovskite material |
Counter electrode |
Electrolyte |
Potential range in V (vs. Li/Li+) |
Current density |
1st cycle capacity (mAh g−1) |
Discharge capacity after ‘n’ cycles (mAh g−1) |
Ref. |
Li-salt and solvent abbreviations in the electrolytes reported: LiPF6 = lithium hexafluorophosphate, LiBF6 = lithium tetrafluoroborate, LiTFSI = lithium bis(trifluoromethanesulfonyl)imide, LiCl = lithium chloride. EC = ethylene carbonate; DMC = dimethyl carbonate; DEC = diethyl carbonate; EMC = ethyl methyl carbonate; PC = propylene carbonate; DOL = 1,3-dioxolane; DME = dimethoxyethane. |
CsPbCl3 |
LiFePO4 |
1 M LiBF4 in PVDF-HFP + BMIMBF4 (1 : 3 w/w) |
3.60.01- |
50 mA g−1 |
612 |
∼300 (70 cycles) |
45 |
Cs4PbCl6 0DPG-A (zero-dimensional glass powder-A) |
Li-metal foil |
1.0 M LiPF6 in EC/DMC (1 : 1 vol ratio) |
0.01–3 |
5 A g−1 |
1387.9 |
510.5 (1000 cycles) |
178 |
CsPbBr3@CNTs |
Li-metal disks |
1.0 M LiPF6 in EC/DMC (1 : 1 vol ratio) |
0.001–3.0 |
100 mA g−1 |
644.6 |
470.2 (200 cycles) |
177 |
CsPbBr3@CNTs |
Li |
1.0 M LiPF6 in EC/DMC/EMC (1 : 1 : 1 vol ratio) |
0.2–3 V |
1 A g−1 |
629 |
∼376 after (900 cycles) |
44 |
CsPbBr3 |
Li-metal foil |
1.0 M LiTFSI in DOL/DME (1 : 1 vol ratio) |
2.8–0.05 |
60 μA cm−2 |
102.6 |
∼73.8 after (32 cycles) |
176 |
CsPbBr3 |
Li-metal foil |
LiTFSI + BMIMTFSI film |
3.0–0 |
60 mA g−1 |
259 |
∼200 (100 cycles) |
179 |
Cs4PbBr6 0DPG-A |
Li-metal foil |
1.0 M LiPF6 in EC/DMC (1 : 1 vol ratio) |
0.01–3.0 |
5 A g−1 |
195 |
429.6 (1000 cycles) |
178 |
CsPbI3 |
Li-metal foil |
1.0 M LiPF6 in a EC/DMC (1 : 1 vol ratio) |
0.1–3 |
40 mA g−1 |
151 |
235 (100 cycles) |
46 |
Cs4PbI6 0DPG-A |
Li-metal foil |
1.0 M LiPF6 in EC/DMC (1 : 1 vol ratio) |
0.01–3.0 |
5 A g−1 |
1589.7 |
36.1 (1000 cycles) |
178 |
MAPbI3 |
Li-metal foil |
1.0 M LiPF6 in EC/DMC/EMC (1 : 1 : 1 vol ratio) |
0.1–1.5 |
200 mA g−1 |
43.6 |
9 (200 cycles) |
43 |
MAPbI3 |
Li-metal foil |
LiPF6 in EC/DMC (1 : 1 vol ratio) |
0.01–2.5 |
100 mA g−1 |
476 |
202 (50 cycles) |
172 |
MAPbI3/ITO |
Li-metal foil |
1.0 M LiPF6 in EC/DEC (1 : 1) vol ratio) |
1.5–0.1 |
0.1 mA cm−2 |
∼340 |
∼50 (10 cycles) |
169 |
MAPbBr3 |
Li-metal foil |
1.0 M LiPF6 EC/DMC |
0.01–2.5 |
300 mA g−1 |
158.2 |
120.1 (200 cycles) |
175 |
MAPbBr3 |
Li sheet |
LiPF6 in EC/DMC |
0.1–1.5 |
300 mA g−1 |
480 |
50 (10 cycles) |
169 |
MAPbBr3 |
Li |
5 M LiTFSI in EC/PC |
0.1–2.7 |
30 mA g−1 |
∼510 |
80 (100 cycles) |
180 |
MAPbBr3 |
Li |
LiPF6 in EC/EMC/DMC (1 : 1 : 1 vol ratio) |
0.1–1.8 |
50 mA g−1 |
∼600 |
260 (50 cycles) |
167 |
MAPbBr3 |
Li-metal |
LiPF6 in EC/EMC/DMC(1 : 1 : 1 vol ratio) |
0.2–1.4 |
200 mA g−1 |
331.8 |
121 (200 cycles) |
43 |
MAPbIBr2/ITO |
Li-metal foil |
1.0 M LiPF6 in EC/DEC (1 : 1) vol ratio) |
1.5–1.0 |
0.1 mA cm−2 |
∼425 |
∼50 (10 cycles) |
169 |
MANiCl3 |
Li-metal foil |
1.2 M LiPF6 in EC/DMC (1 : 1) vol ratio) |
0.1–3.5 |
0.1 mA g−1 |
170 |
∼120 (50 cycles) |
181 |
MANiCl3 |
Li-metal foil |
1.0 M LiPF6 in EC/DMC (1 : 1) vol ratio) |
0.1–2.7 |
∼32 mA g−1 |
650 |
350 (19 cycles) |
170 |
Cs2NaErCl6 |
Li-metal foil |
1.0 M LiPF6 in EC/DMC/EMC (1 : 1 : 1 vol ratio) |
0.1–2.5 |
300 mA g−1 |
522 |
120 (500 cycles) |
5 |
MA2(CH3(CH2)2NH3)2Pb3Br10 |
Li-metal foil |
1.0 M LiPF6 in EC/DEC (1 : 1 vol ratio) |
0.1–1.2 |
0.1 mA cm−2 |
375 |
— |
169 |
Cs4PbBr6 |
Li-metal foil |
1.0 M LiPF6 in EC/DEC (1 : 1 vol ratio) |
0.01–3 |
0.5 A g−1 |
158.6 |
136.5 (1000 cycles) |
182 |
SnF2-doped Cs4PbBr6 (FLII glass matrix) |
Li-metal |
— |
0–2.0 |
0.5 A g−1 |
773 |
741 (350 cycles) |
183 |
(C4H9NH3)2PbI4 |
Li-metal |
1 M LiPF6 in an EC/DMC e (1: 1 vol ratio) with 5% FEC |
0.01–2.5 |
100 mA g−1 |
1605 |
213 (250 cycles) |
172 |
Cs4PbBr6 quantum dots silicate glass-ceramic |
Li-metal foil |
1.0 M LiPF6 in EC/DEC (1 : 1 vol ratio) |
0 to 3 |
50 mA g−1 |
1986.9 |
426.7 (100 cycles) |
184 |
C6H9NOPb |
Li-metal foil |
1 M LiPF6 in EC/DEC (1 : 1 vol ratio) with 5% FEC |
2.5–0.01 |
100 mA g−1 |
1580 |
585 (50 cycles) |
172 |
Li2(taurine)2CuCl4 (LTCC) |
Li-metal foil |
1 M LiPF6 in EC/DEC/DMC 1 : 1 : 1 vol ratio) + 5% FEC |
0.005–3.0 |
1.0 A g−1 |
— |
548 mAh g−1 (550 cycles) |
185 |
C4H20N4PbBr6 |
Li-metal foil |
1.0 M LiPF6 EC/DMC/EMC (1 : 1 : 1 vol ratio) |
0.1–2.5 |
150 mA g−1 |
1632.8 |
598.0 (50 cycles) |
186 |
(BA)2(MA)3Pb4Br13 |
Li-metal foil |
5 M LiTFSI in 1 mL EC/PC (1 : 1 vol ratio) |
2.8–1.8 |
30 mA g−1 |
108 |
∼42 (10 cycles) |
180 |
Mn-substituted Cs3Bi2Cl9 |
Li-foil |
1.0 M LiPF6 in EC/DMC/EMC (1 : 1 : 1 vol ratio) |
0.1–2.5 |
100 mA g−1 |
467 |
117.08 (500 cycles) |
187 |
Cs4PbBr6 nano hexagons/ITO |
Pt |
LiCl |
−1 to +0.1 |
45 mA g−1 |
377 |
— |
188 |
Cs2NaBiCl6 |
Li-metal foil |
1.0 M LiPF6 in EC/DMC/EMC (1 : 1 : 1 vol ratio) |
0.01–2.5 |
75 mA g−1 |
∼775 |
300 (25 cycles) |
189 |
Cs2NaBiCl6:xMn2+ |
Li-metal foil |
1.0 M LiPF6 in EC/DMC/EC (1 : 1 : 1 vol ratio) |
0.01–2.5V |
70 mA g−1 |
— |
240 (300 cycles) |
190 |
3.1.2 Low-dimensional MHPs-based anodes. In contrast to 3D perovskites, lower-dimensional variants are favored for improved lithium storage due to the greater space present in their layered architecture. These low-dimensional perovskites can enhance capacity and provide better reversibility than their 3D counterparts. The effect of dimensionality on electrochemical properties was first investigated in 2017 by Tathavadekar et al.172 They discovered that lowering the dimensionality of perovskites was effective in improving the lithium storage. They used 1D C6H9I3NOPb, 2D (C4H9NH3)2PbI4, and 3D MAPbI3 perovskites and revealed that the first discharge capacity for 1D and 2D perovskites is nearly 4 times higher than the 3D perovskite, making them potential active materials for anodes in LIBs (Fig. 10a). Similar results have been reported by Hong Kong et al.186 in 2022, they fabricated three different dimensional perovskites as anodes in the Li-ion battery. Among the anodes, the 1D C4H20N4PbBr6-based one offers the best performance, providing a stable capacity value of 598.0 mAh g−1 (Fig. 10b–d). The reason for this is that 2D and 1D perovskites have a greater spacing between crystal planes than close-packed 3D structures, which allows more Li-ions to intercalate. As a type of low-dimensional material, RP perovskites possess a quasi-2D structure, which allows for the optimization of energy storage capacity through the arrangement of layers. De Volder et al.180 examined the electrochemical capacity and cyclic stability of a series of (BA)2(MA)n−1PbnX3n+1 as anodes for LIBs by adjusting the n values (Fig. 10e and f). It is evident from Fig. 10g that the sample with n = 4 exhibited the best performance in terms of battery capacity, while a noticeable decline in capacity was observed as per units increased. And the spacer layers that facilitate ion diffusion and provide structural constraint. Recently, Maity et al. compared the electrochemical performance of 3D MHP (CsMAPbIBr) and the 2D–3D hybrid MHP (CsMABPAPbIBr) as shown in Fig. 10h.171 They demonstrated that the mixed dimensionality furnishes more accessible sites for Li-ions compared to the control 3D MHP (CsMAPbIBr), thereby increasing both capacity and rate capability. Li-ion cells utilizing the 2D–3D hybrid MHP–CNTs composite demonstrated an outstanding discharge capacity of 221 mAh g−1, an energy density of 508 Wh kg−1, and 84% of capacity retention after 100 cycles, marking a significant enhancement compared to 3D MHP–CNTs (Fig. 10i and j). Due to the environmental issues associated with Pb-based perovskites stemming from the recognized toxicity of lead, it is crucial to establish stable lead-free perovskite alternatives for LIBs. Research focused on lead-free perovskite materials and their photoelectric applications has shown that lead-free perovskites featuring intricate crystal structures are capable of accommodating defects and intercalated ions, which could lead to effective Li-ion storage.191,192 Pandey et al.193 reported that the 2D (MA)2CuBr4 and 3D Cs2CuBr4 possess a reversible capacity of ∼480mAh g−1 and 420 mAh g−1, respectively. In their study, it was observed that 2D material capacity increased gradually with subsequent charge–discharge cycles, achieving 630 mAh g−1 at 140 cycles. The reason for this observation stems from the slow percolation of electrolyte into the electrode as well as the appearance of Li-ion intercalation sites in response to cycling. Similarly, Yang and co-workers introduced lead-free all-inorganic double perovskite Cs2NaBiCl6
188 and Cs2NaErCl6
5 as anode material for LIB. It was revealed that Cs2NaBiCl6-based battery properties are not beneficial, especially poor battery cycle stability, which limits their practical use. Nevertheless, Cs2NaErCl6 as a negative electrode material showed high cycle stability, with a specific capacity of 120 mAh g−1 after 500 cycles at a current density of 300 mA g−1 with a Coulomb efficiency of nearly 100%. Recently, a new class of low-dimensional lead-free Cs3Bi2Cl9 was synthesized by Jia et al.187 The authors demonstrated that the initial discharge specific capacity can be enhanced from 263.39 mAh g−1 to 467 mAh g−1 and stabilized to values from 89.32 mAh g−1 to 125.3 mAh g−1 after 100 cycles, with a coulombic efficiency of more than 99%, and stabilized at 117.08 mAh g−1 after 500 cycles. Recently, Wu et al.185 presented a layered perovskite Li2(C2H7NO3S)2CuCl4 (LTCC). The LTCC anode can achieve a remarkable specific capacity of 861 mAh g−1 at 0.1 A g−1 after 100 cycles. Moreover, it retains a high discharge capacity of 548 mAh g−1 over 550 cycles at 1.0 A g−1, exhibiting outstanding cycling stability among perovskite-type anode materials for LIBs. However, compared to traditional 3D perovskites, there is a limited number of studies on low-dimensional, lead-free, and non-perovskite MHPs in LIBs. Nevertheless, emerging research in this area highlights significant potential for future exploration and development. The quantitative comparison of energy density and cycle life of different MHPs in LIBs is given in Table 4.
 |
| Fig. 10 (a) Cyclic stability for 1–2–3 D hybrid perovskite at a current density of 100 mA g−1 in the potential window of 2.5–0.01 V 172 Copyright 2017, J. Mater. Chem. A Galvanostatic charge–discharge profiles of first three cycles of (b–d) 1D, 2D and 3D perovskite at a current density of 100 mA g−1 between 0.01 and 2.50 V 186 Copyright 2022, J. Energy Chem. (e) Quasi 2D (BA)2(MA)n−1MnX3n+1 (n = 1) perovskite structure with individual layers of PbX4 intercalated between two BA organic chains. (f) n = 2-layered perovskite structure. (g) Gravimetric charge–discharge capacities of the bromide-based layered perovskite (BA)2(MA)n−1PbnBr3n+1 from n = 1–n = 4 and the respective bulk MAPbBr3 perovskite (equivalent in structure to n = ∞) as a function of cycle number from cycle 11–100; the first 10 cycles are highlighted inset. Specific charge capacities are shown shaded and specific discharge capacities block colour. A current density of 30 mA g−1 and a potential window of 2.85–0.1 V are used180 Copyright 2021, Mater. Adv. (h) Schematic structure of 2D–3D hybrid PSK (CsMABPAPbIBr), showing the interactions between Br(Pr)NH2·HBr and the unit cell of the PSK (perovskite). (i) GCD plots acquired intermittently during cycling at 20 mA g−1, (j) capacity/CE variation with cycling171 Copyright 2025, ACS Appl. Energy Mater. | |
Table 4 Comparison of energy density and cycle life of different MHPs in LIBs
Perovskite material |
Year of publication |
First discharge capacity (mAh g−1) |
Average voltage (V) |
First discharge energy density (Wh kg−1) |
Stable capacity (mAh g−1) |
Stable energy density (Wh kg−1) |
Cycling life |
Ref. |
CsPbCl3 |
2020 |
612 |
1.805 |
1105.26 |
∼300 |
541.50 |
70 cycles |
45 |
CsPbBr3@CNT |
2025 |
∼640 |
1.50 |
961.6 |
376 |
564 |
900 cycles |
44 |
CsPbBr3 |
2021 |
259 |
1.5 |
388.5 |
∼200 |
300 |
100 cycles |
179 |
CsPbI3 |
2021 |
151 |
1.55 |
234 |
235 |
364 |
100 cycles |
46 |
MAPbI3 |
2017 |
476 |
1.255 |
597 |
202 |
254 |
50 cycles |
169 |
MAPbBr3 |
2021 |
∼510 |
1.4 |
714 |
80 |
112 |
100 cycles |
180 |
MAPbBr3 |
2015 |
331.8 |
0.8 |
265.4 |
121 |
96.8 |
200 cycles |
43 |
MANiCl3 |
2024 |
170 |
1.8 |
306 |
∼120 |
216 |
50 cycles |
181 |
MANiCl3 |
2020 |
650 |
1.4 |
910 |
350 |
490 |
19 cycles |
170 |
Cs4PbCl6 0DPG-A |
2022 |
1387.9 |
1.505 |
2088.8 |
510.5 |
768.3 |
1000 cycles |
178 |
Cs4PbI6 0DPG-A |
2022 |
1589.7 |
1.505 |
2392.5 |
36.1 |
54.3 |
1000 cycles |
178 |
Cs4PbBr6 0DPG-A |
2022 |
195 |
1.505 |
293.5 |
429.6 |
646.5 |
1000 cycles |
178 |
MA2(CH3(CH2)2NH3)2Pb3Br10 |
2018 |
375 |
0.65 |
243.8 |
— |
— |
— |
180 |
(C4H9NH3)2PbI4 |
2017 |
1605 |
1.25 |
2014.3 |
213 |
267.3 |
250 cycles |
172 |
C6H9NOPb |
2017 |
1580 |
1.25 |
1982.9 |
585 |
734.2 |
50 cycles |
172 |
Li2(taurine)2CuCl4 (LTCC) |
2025 |
— |
1.5 |
— |
548 |
823.4 |
550 cycles |
185 |
C4H20N4PbBr6 |
2022 |
1632.8 |
1.3 |
2122.6 |
598.0 |
777.4 |
50 cycles |
186 |
(BA)2(MA)3Pb4Br13 |
2021 |
108 |
2.3 |
248.4 |
∼42 |
96.6 |
10 cycles |
180 |
Mn-substituted Cs3Bi2Cl9 |
2025 |
467 |
1.3 |
607.1 |
117.08 |
152.2 |
500 cycles |
187 |
Cs2NaBiCl6 |
2021 |
775 |
1.25 |
972.6 |
300 |
376.5 |
25 cycles |
189 |
Cs2NaErCl6 |
2022 |
300 |
1.3 |
390 |
120 |
156 |
500 cycles |
5 |
Cs4PbBr6 quantum dots silicate glass-ceramic |
2022 |
1986.9 |
1.5 |
2980.35 |
426.7 |
640.05 |
100 cycles |
184 |
3.2. MHPs as ASEI
The major issue with Li-metal batteries (LMBs) is the self-derived unstable SEI, which possesses low Li+ conductivity, low mechanical modulus, and inhomogeneous composition, which makes it difficult to achieve smooth and stable deposition/stripping of Li metal anode.194 Due to an inhomogeneous SEI, a non-uniform Li deposition will result, which facilitates the formation of Li dendrites. Li dendrites are capable of piercing low mechanical modulus SEI, leading to the formation of new SEIs derived from Li metal reactions with electrolytes. During Li stripping, Li dendrites are likely to isolate from the Li bulk and turn into “dead’’ Li, resulting in low coulombic efficiency. After repeated cycling, the dendrites may grow to several hundred microns and penetrate the separators, causing short circuits and safety hazards.195 In this regard, researchers are exploring various strategies for inhibiting dendrite growth: (a) Li-alloy anodes,196 (b) solid-state electrolyte,197 (c) structured anodes,195 (d) ASEI,198 (e) electrolyte additives,199 and (f) interface modifications.200 Among them, one of the most effective approaches to inhibit dendrite problems has been reported to be regulating Li-ion distribution via the application of an artificial protective layer on the electrode surface, which reduces current density and strengthens the electrode/electrolyte interface stability.201 So far, a variety of protective materials, including inorganics, polymers, and hybrids, have been applied to LMAs through various deposition methods, such as solid gas reactions, atomic layer deposition, or wet chemical emulsion coating.202–204
Choosing the materials for ASEI coating depends on the preparation conditions of thin films on LMAs, as well as the properties of the materials, such as stability against electrolytes, electron insulation, high Li-ion conduction, and flexibility confirm the changes in the volume of the lithium anode.201 Currently used materials for SEI are capable of effectively separating LMAs from electrolytes and preventing spontaneous side reactions. However, structural stability and high ion conductivity are incompatible, which makes Li plating/stripping tough, further restricting LMB capacity and long-term performance. It is therefore highly desirable to develop ASEI materials that have good structural stability and Li-ion conductivity.205
Due to its adjustable 3D framework structure and bandgap, MHPs can achieve Li-ion conduction and electronic insulation, which is expected to become a promising candidate for constructing a high-performance SEI layer. In 2020, an interfacial layer composed of solution-processed MASnCl3 and MAPbCl3 perovskites was developed by Yin et al.47 as a new type of interfacial layer for the LMA through a solid-state transfer process. They demonstrated through galvanostatic Li plating and stripping that MSC-Li cell can be cycled for more than 800 h, much longer than the 100 h cycling life of the cell using bare Li (Fig. 11a). They also evaluated the electrochemical performance of Li4Ti5O12 (LTO)/perovskite coated Li metal batteries at a high rate of 5C to demonstrate the efficiency of perovskite protection in LMBs. It was shown that the bare Li metal anodes exhibit a drastic capacity decay, with a capacity of 28.3 mAh g−1 at the 400th cycle. In contrast, the cells using MASnCl3 and MAPbCl3 coated LMA show stable cycling for more than 500 cycles with a low capacity decay rate of 0.07% per cycle (Fig. 11b). It was concluded that the metal chloride perovskite protection layer can ensure stable cycling of LMBs under strict conditions. Furthermore, based on DFT calculations, the researchers proposed a Li-ion transport gradient layer model that illustrated the shielding mechanism for dense deposition of Li-metal using perovskite thin films, shown in Fig. 11c.
 |
| Fig. 11 (a) Galvanostatic voltage curves (top) of bare Li and MSC-Li tested with a current density of 1 mA cm−2 for 1 mAh cm−2, and the enlarged voltage curves during different periods (bottom). (b) Galvanostatic cycling performances of cells using LTO as cathode and bare Li, MSC-Li or MPC-Li as anode. (c) Schematic illustration of the mechanism of Li-ions' intercalation into perovskite lattice, the formation of perovskite-alloy gradient Li-ion conductor, and the deposition process47 Copyright 2020, Nat. Commun. (d) Voltage profiles of pristine Li and Li: CsPbI3 symmetric cells measured at a current density of 1 mA cm−2 and a discharge capacity of 1 mA h cm−2 206 Copyright 2020, J. Mater. Chem. A (e) Voltage–time curves of Li/Li and Li@Li–CsPbCl3/Li@Li–CsPbCl3 symmetric cells with an areal capacity of 1 mAh cm−2 at a current density of 1 mA cm−2. (f) Long-term cycling stabilities of Li/LiFePO4 cells with bare Li and Li@ Li@Li–CsPbCl3 anodes at a current density of 3C48 Copyright 2024, Energy Adv. (g) Schematic illustration of Li plating on KNiF3 perovskite SEI through octahedral structure and on bare Li205 Copyright 2022, Chem. Eng. J. | |
Step 1: only Li-ions absorption within the perovskite framework, leaving solvent molecules outside the perovskite framework.
Step 2: intercalation and migration of Li-ions into the highly symmetric perovskite framework.
Step 3: electrochemical conversion reaction at the interface of the perovskite layer and the substrate.
Step 4: formation of both the insulating LiCl layer and the Li–M alloy layer. A Li–M layer will facilitate homogeneous Li deposition. Nevertheless, the generated LiCl can insulate electrons, preventing the perovskite from further conversion reactions, ensuring that the perovskite remains stable in its top state.
In the same year, Kaisar et al.206 fabricated δ-CsPbI3 as an electrochemical intercalation layer through an inexpensive and facile spray-coating method that stabilizes Li electrodes for LMBs. DFT calculations confirmed the Li-ion intercalation into the δ-CsPbI3 framework, forming Li:CsPbI3. Electrochemical testing of a Li:CsPbI3 symmetric cell revealed dendrite-free plating after 1000 h at a current density of 1 mA cm−2 and discharge capacity of 1 mAh cm−2 (Fig. 11d). In this new and simple method, derogatory dendrites are avoided, thereby enabling the preparation of LMAs for practical application in high-density LMBs. Recently, Liu et al. developed lithium-doped CsPbCl3 ASEI. They demonstrated that Li–CsPbCl3 not only successfully inhibits the formation of lithium dendrites but also promotes the movement of Li-ions at the interface and encourages the ultra-dense and even deposition of Li, creating a beneficial setting for stable Li electroplating/stripping and greatly enhancing the electrochemical performance of LMBs (Fig. 11e and f). However, hybrid perovskites such as MAPbCl3 and MASnCl3 are prone to decomposition when exposed to photo-, thermal-, or moisture-stresses.207,208 Furthermore, the cubic phase of CsPbI3 is thermodynamically unstable at room temperature, which may result in a blocking of Li-ion transfer channels.209 In addition to that, lead-based perovskites are highly toxic, which makes them unsuitable for LMBs. Therefore, perovskite materials that are non-toxic and have a high moisture- and thermal stability would be better suited to Li-metal SEI. In this regard, Y. Zhang et al.205 developed air-stable fluoride perovskite (KNiF3), which is applied as an SEI layer to induce uniform Li-ion deposition for an air-stable and dendrite-free LMA. In their study, they demonstrated that symmetric cells protected by KNiF3 SEI maintain high cycling stability over a period of 3000 hours at a capacity of 4 mAh cm−2. When coupled with commercial LiFePO4 cathodes (LFP, 13.3 mg cm−2), LFP‖Li–KNiF3 batteries show promoted cycling stability and rate capability, much better than the bare Li. Even though MHPs are capable of improving cycling performance, rate capability, and stability. Moreover, they also reveal the protection mechanism of the perovskite SEI through DFT calculation. They demonstrated that Li-ion migrated along the octahedral structure of the perovskite while maintaining the 3D cubic framework without decomposition or phase transition (Fig. 11g).
4. MHPs-based photo-induced LIBs
It is essential to integrate energy storage systems with photovoltaic technology to efficiently and widely utilize solar energy. Conventional photo-rechargeable batteries are made up of a photovoltaic cell and a storage battery, which are separate systems linked by an external wire (Fig. 12a).210 Nevertheless, this type of system is known for its high expense, size, and various other issues.42 To address this problem, integrated photovoltaic rechargeable batteries were developed that, unlike non-integrated systems, merge solar energy collection and storage into one device, potentially resulting in more efficient and streamlined solutions.211 These integrated systems can be implemented in two configurations:51
 |
| Fig. 12 Working principle of the photo-assisted energy storage device (a) conventional three electrode system210 reproduced with permission 2024, Adv. Funct. Mater., (b and c) the charging and discharging process of a two-electrode device. (d and e) The charging and discharging process of a three-electrode device211 reproduced with permission 2024, Nano Energy. | |
(a) Three-electrode system: this type of system includes a photoelectrode for light conversion, a counter electrode for storing energy, and a common electrode (as a positive or negative electrode) between the photovoltaic cell and the battery (Fig. 12b).
(b) Two-electrode system: the positive electrode has an integrated function, i.e., both photoconversion and energy storage (Fig. 12c).
In this section, we will be discussing the configuration and working principle of photo-induced batteries, followed by the role of MHPs in photo-induced LIBs.
4.1. Working principle
The fundamental operating principle of a photo-battery remains largely consistent, regardless of the type or configuration of the device. When illuminated, the photo-active electrode produces electron–hole pairs as a result of the photovoltaic effect. The electrons and holes generated play a role in the reactions that take place during the charge and discharge cycle. During the charging process (Fig. 12b), a link is formed between the external circuit and both the anode and photoelectrode. When the photoelectrode material is exposed to light, it gets excited, generating high-energy photoelectrons that jump from the VB of the semiconductor to the CB. Concurrently, positive holes are created in the VB of the material. As these holes migrate to the outer layer of the active material on the photoelectrode/cathode side, an oxidation reaction takes place. In this reaction, the reduction product from the photoelectrode (CR) transforms into the oxidation product (CO), as demonstrated by the following equation:
At the same time, the photoelectrons originating from the VB of the semiconductor will move through the external circuit toward the anode and interact with the migrating metal ions Mn+ from the cathode side. This interaction leads to a reduction reaction that restores the metal M, as depicted in the following equation:
As a result, the battery device converts electrical energy into chemical energy, which is subsequently stored.
During the discharge process (Fig. 12c), a link is formed between the photoelectrode and the anode to enable integration with the external load. Photoelectrons that are excited on the surface of the semiconductor engage in a reduction reaction with the charging product (CO) from the photoelectrode, resulting in the production of the discharge product (CR), as indicated in eqn (5). At the same time, the photo-holes interact with the electrons released from the oxidation reaction taking place on the anode side, leading to their recombination, as illustrated by the eqn (6):
In this procedure, chemical energy is converted into electrical energy and discharged to supply power to the external load. In more intricate three-electrode systems, the photo-assisted charging mode frequently entails a multi-step reaction occurring at the shared electrode side. This reaction involves additional holes to oxidize the shuttle mediator, as shown in eqn (7)
|
Mred + h+ →Mox
| (7) |
Then Mox will oxidize the reduction product (CR) as shown in eqn (8)
Solar energy is stored as chemical energy in the form of Co at the photoelectrode and the Mred at the counter electrode. During the discharge process (Fig. 12e), the Mred is oxidized, and electrons are transferred from the external circuit to the shared electrode. Simultaneously, the Co reduced back to its reduced state. This completed the charge–discharge cycle of the integrated three-electrode photo-rechargeable battery.
4.2. Role of MHPs in photo-induced batteries
Perovskite halides have become significant in the domains of photovoltaics and energy storage, and they are now being explored as photoactive materials for photo-batteries. This is due to the same characteristics that make them suitable for photovoltaic technologies and batteries: an adjustable bandgap, high mobility of charge carriers, a low rate of non-radiative recombination, an extensive absorption spectrum, long charge diffusion lengths, and minor effective masses of carriers, as previously mentioned. Connecting a LIB directly to a solar cell allows for self-charging capabilities. Solar cells are capable of converting solar energy into electrical energy, which can then be stored in LIBs for later use. In a study by Dai et al.,41 a PSC was directly linked to four individual MAPbI3 packs with a LIB, achieving a notable overall conversion and storage efficiency of 7.80% along with outstanding cycling stability. Nevertheless, external photo-rechargeable batteries have several drawbacks, including the tendency to incur ohmic loss during energy transfer between solar cells and storage batteries, as well as their bulky size and high expense, which do not align with the evolving needs of portable electronic devices. As a result, integrated photo-rechargeable batteries have increasingly emerged as a focal point for research. In this regard, Ahmad et al.212 explored the application of 2D lead-based perovskites, specifically (PEA)2PbI4 (PEA = C6H9C2H4NH3) (Fig. 13a and b), as a photo-active electrode material for LIBs. The battery utilizing the iodide perovskite achieved a specific capacity of up to 100 mAh g−1 at a current rate of 30 mA g−1. Incorporating reduced graphene oxide (rGO) as a conductive additive enabled it to demonstrate photo-charging when illuminated, without needing an external load, across a voltage range of 1.4–3.0 V. This was then succeeded by a 25-hour discharge period with a 21.5 kΩ resistor acting as the load, thus allowing the device to operate as a genuine photo-battery. When the voltage range was lowered to below 1.4 V, the perovskite experienced irreversible degradation due to the reduction of Pb2+ to Pb0. Additionally, the battery provided a higher voltage output when discharged in light compared to darkness. With a 21.5 kΩ resistor as the load, Ahmad et al. recorded a photo-conversion efficiency of 0.034%. Additionally, He et al.213 proposed a hypothesis involving polarons to clarify the process of photo-rechargeability in the (PEA)2PbI4 perovskite system. Importantly, the movement of the Li-ion is indicated to occur following the formation of a hole polaron, which is believed to reflect the photo-charging activity that triggers the one-way flow of Li-ions on the surface of the electrode. This illustrates that MHPs are viable electrode materials and can function as the active layer for photo-charging in photo-rechargeable perovskite batteries. Due to toxicity issues related to Pb-based MHPs, lead-free MHPS also gained a lot of attention in photo-induced batteries. Tewari et al.214 created a photo-rechargeable LIB utilizing a lead-free, fully inorganic perovskite material known as Cs3Bi2I9 for the photo-electrode (Fig. 13c and d). In their experiments, the battery initially exhibited a discharge capacity of 413 mAh g−1 at a current of 50 mA g−1. Nevertheless, after several cycles, the performance deteriorated due to the transformation of bismuth in the perovskite from Bi3+ to metallic Bi0, which influenced the material's structure and functionality. To investigate its photo-charging characteristics, they fabricated the photo-electrode on either FTO-coated glass or porous carbon felt and assessed the battery under light conditions. They discovered that the perovskite was the key component facilitating the light-induced charging process. Upon exposure to light with energy surpassing the bandgap of the perovskite, electrons were excited and traversed a layer of PCBM to reach the current collector, resulting in an accumulation of positive charges (holes) within the perovskite. These holes caused Li-ions to be repelled back into the electrolyte, aiding in the battery's charging process.212,215 However, further investigation is needed to understand the fate of these photo-generated electrons fully (Table 5).
 |
| Fig. 13 (a) First photo-charge (at 100 mW cm−2) and discharge (dark, 21.5 kΩ load) voltage profile of the 2D (PEA)2PbI4 (CHPI). (b) The discharge curves of CHPI based photo-batteries in dark and illuminated conditions212 Copyright 2018, Nano Lett. (c) Cycles of discharge of the Cs3Bi2I9 photo-battery in dark at 100 mA g−1 followed by photo-charging under 100 mW cm−2. (d) First discharge curves of the Cs3Bi2I9 photo-battery in dark and under light. Electrochemical performance of PR-LIB under illumination214 Copyright 2021, Nano Lett. (e) Capacity variation of the PE at 50 and 100 mA g−1 under dark and light states. (f) I–T curves with alternating light and dark states217 Copyright 2025, Sci. China Mater. | |
Table 5 Summary of different MHP-based photo-induced batteries and their performance
Configuration of PSCs |
Configuration of energy storage units |
Integrated strategy |
Photocharge capacity [mAh g− 1] |
Photo conversion efficiency (PCE) % |
Ref. |
ITO/PEDOT:PSS/MAPbI3 |
LiFePO4/Li4Ti5O12 |
Wire connection of 4 cells |
140.4 at 0.5 °C |
7.8 |
41 |
FTO/(PEA)2PbI4/rGO |
PEAPbBr4/Li |
Dual functional electrode |
100 |
0.034 |
212 |
ITO/PTAA/MAPbI3/PCBM/BCP/Ag |
LiFePO4/Li4Ti5O12 |
DC–DC |
0.372 mAh at 6C |
9.9 |
218 |
ITO/PEDOT:PSS/MAPbI3/PCBM/Ag |
LiFePO4/Li4Ti5O12 |
DC–DC |
151.3 at 0.5 °C |
9.36 |
219 |
FTO/rGO (PCBM)/Cs3Bi2I9/Cu |
Cs3Bi2I9/Cu |
Dual functional electrode |
410 |
0.43 |
214 |
ITO/PYAA/MAPbI3/PCBM/C60/BCP/Al |
Al/graphite |
Dual functional electrode |
757–770 |
12.04 |
220 |
FTO/TiO2/ZrO2/MAPbI3/carbon |
FTO/NiO/rGO/WO3/FTO |
Wire connection |
63 at 5C |
— |
221 |
FTO/TiO2/MAPbI3/carbon |
Li/carbon |
Joint electrode mode |
750 |
5.14 |
222 |
(MA)3Bi2I9/rGO |
MA3Bi2I9/rGO/Li |
Dual functional electrode |
282.4 |
11.8 |
217 |
(BA)2(MA)3Pb4I13/MoS2/rGO |
(BA)2(MA)3Pb4I13/MoS2/rGO/Li |
Dual functional electrode |
180.67 |
0.52 |
223 |
FTO/CsPbI2Br |
S–Li7P2.9Sb0.1S10.75O0.25–C |
Dual functional electrode |
∼1500 at 0.1C |
11.2 |
224 |
FTO/TiO2/MAPbI3 |
FTO/TiO2/MAPbI3/Zn foil |
Dual functional electrode |
362 |
0.31 |
225 |
MAPbI3/PEDOT/ZrO2 |
MAPbI3/PEDOT/ZrO2/Zn anode |
Dual functional electrode |
555 |
0.51 |
226 |
As noted by Paolella et al., during electrical charging, these electrons could flow through the external circuit to the opposite electrode, where they reduce Li-ions to Li-metal. In scenarios lacking an external circuit, the electrons might interact with the battery's electrolyte (ethylene carbonate/dimethyl carbonate), generating reactive oxygen species that ultimately contribute to the formation of the SEI on Li-metal.216 Furthermore, Tewari and Shivarudraiah214 discharged the battery to 0.9 V and then photo-charged it up to approximately 2.5 V under light without any external load. They measured a photo-conversion efficiency of 0.43% during the initial cycle, but this efficiency declined to about 0.1% in subsequent cycles. Despite the existing limitations, their findings indicate that Cs3Bi2I9 perovskites have significant potential for photo-rechargeable batteries, and future enhancements in efficiency may stem from a better comprehension of how light-driven charge separation and transport function in these systems. Yin and colleagues have recently developed a photo-rechargeable LIB utilizing a bismuth-based hybrid perovskite known as (MA)3Bi2I9, commonly called MBI, as the light-sensitive electrode.217 Their research demonstrated that this material enhances charge separation when exposed to light, facilitating better Li-ion mobility in the battery and improving its overall performance. Investigations using ex situ XRD indicated that Li-ion storage within MBI occurs through a combination of two mechanisms: the insertion of Li-ion into the structure and a chemical conversion reaction. When the battery was evaluated under light conditions, its discharge capacity increased from 236 mAh g−1 in darkness to 282.4 mAh g−1, marking a 19.7% enhancement at a current rate of 50 mA g−1 (Fig. 13e). This improvement was attributed to the light-generated electrons and holes in MBI, which accelerated charge movement and the flow of Li-ions. This not only increased the battery's capacity but also lowered the charge flow resistance. From the perspective of energy efficiency, the battery exhibited better performance in light conditions. It required 0.1 V less for charging, leading to a 6% reduction in energy input, and delivered 0.1 V more during discharge, resulting in an 11.8% increase in energy output (Fig. 13f). In summary, this research indicates that MBI-based batteries could pave the way for the next generation of energy storage solutions devices that are efficient, environmentally friendly, and well-suited for portable electronics powered by light.
5. Perovskites in supercapacitors and photo-induced supercapacitors
The integration of perovskite materials into SCs has garnered considerable attention, given the ongoing efforts to improve energy density, efficiency, and charge storage for bridging the performance differences between standard capacitors and batteries, since they deliver high power density with quick charge–discharge rates. Nonetheless, traditional SCs typically struggle with energy density and stability, while different structures of perovskite materials can enhance their electrochemical performance. In SCs, the spotlight is on identifying the mechanisms by which perovskites affect charge storage and transfer processes and investigating their operational efficiency and cyclic stability.227,228 Studies demonstrate that ionic and electronic conductivity are essential for supercapacitor performance. In addition, the properties of MHP-based materials can be enhanced by simply making modifications in the crystal structure. Outcomes suggest that perovskite-based supercapacitors significantly increased specific capacitance, reaching over 200 F g−1, compared to the 60–120 F g−1 seen in typical carbon-based materials, which stems from the high surface area and superior ionic conductivity of perovskite materials, which allow for more effective ion transport and charge storage.229 The perovskite SCs performed well over time, retaining over 90% of their initial capacitance after 5.000 charge–discharge cycles.230 NCs combined with reduced graphene oxide (rGO) have delivered a specific capacitance 178 times better than rGO electrodes on their own.231 Furthermore, cycle life tests lasting over 10.000 cycles with minimal capacity fade strengthen the case for perovskite-based SCs in practical applications, which closely aligns with the stability problems in other advanced materials.232 In general, the room temperature laser-triggered technique selected for the conjugation of the two components presents a distinct avenue for the cost-effective and large-scale synthesis of precisely tailored perovskite-2D conjugates.231 However, limitations of perovskites' usage in SCs are due to their inherent instability and sensitivity to moisture and heat changes. These contentious issues must be systematically addressed to hint at a factual potential for commercial applications in high-performance energy storage. Furthermore, these devices maintain robust capacitance retention, generally above 80% following numerous cycles, suggesting good operational stability. The data suggest that perovskite materials integrated with carbon composites in hybrid electrode designs can create synergistic effects, improving both electron and ion transport, and thus, performance, and reinforcing the potential benefits of perovskite-based approaches. This can be attributed to the synergy of the EDLC and pseudo capacitance originating from the different components. In particular, they showed a specific capacitance value of 106 F g−1 with excellent stability, remaining 97.2% after 100 continuous intercalation/deintercalation scans. This route allows one to create NCs with various morphologies and chemical phases, along with multiple 2D materials, to discover the optimal combinations.230 Different 2D materials with large electrolyte contact areas and numerous energy storage active sites may serve as alternatives to the low capacitance NH3-functionalized rGO for enhancing the capacitance of the conjugated systems.231 For example, layered perovskites are associated with better ion transport dynamics, giving them distinct advantages over materials like manganese oxides and carbon-based electrodes.232
When these results focus on nanostructured materials, the performance improvement appears noticeably. These conclusions are also consistent with studies that highlight how scalable synthesis methods have been key in reaching these performance levels with perovskites. Nevertheless, it is also important to note the challenges, particularly concerning the environmental impact and the operational stability of perovskite materials over longer periods. The fine-tuning of the anionic species in the perovskite framework has noticeably improved ionic conductivity, which helps charge move better within the SC.233 For example, changing the lead content in lead-based perovskites leads to different electrochemical behaviors and stabilities.230 Similarly, the incorporation of dopants can effectively tweak the bandgap of perovskite materials, increasing their performance in SC applications. Moreover, looking at various perovskite formulations, lower-dimensional perovskites tend to have better capacitance and energy density than their 3D counterparts.231 The insights could lead to new perovskite composites that further enhance their electrochemical properties, keeping them relevant in the area of energy storage.234 Sol–gel and hydrothermal methods, for example, give different structures that boost ion conductivity and surface area-both vital for storing charge effectively in SCs.235 The synthesis method of perovskites may lead to the conclusion that using hydrothermal methods can deliver better electrical performance, with more capacitance and better stability.228 Adding other elements and materials during production is also essential for the construction of perovskites for supercapacitors.236 For example, perovskites mixed with conducting polymers can work together to improve both electrical conductivity and mechanical stability. This helps to solve some significant problems when using perovskites in energy storage. Perovskite materials, especially the organic–inorganic type like MAPbI3, are showing power conversion and absorbing light.231 In general, light intensity and wavelength play a crucial role in getting the most out of perovskites. If light intensity changes, it can significantly enhance charge separation and the speed at which electrons move. Some perovskite structures, such as thin microbelts, convert light into energy, particularly when exposed to visible light. SC energy storage largely depends on what happens electrostatically and electrochemically at the electrode material interface. Photo-induced SCs are based on solar cells for photoelectricity energy conversion and SCs for energy storage. Liu and co-workers have incorporated an all-solid-state photo-charging capacitor based on MAPbI3 and SCs (polyaniline (PANI)/carbon nano tube (CNT)), where the CNT bridge was devoted to preventing water from the aqueous gel electrolytes (Fig. 14). Indeed, the CNT bridge could be a path for holes to transport between the two electrodes.237
 |
| Fig. 14 Representation of the photo capacitor and energy level schematic.237 Copyright 2017, J. Mater. Chem. A. | |
The MHP-based PSCs achieve excellent charging performance, recording energy, with efficient storage mechanisms, and power densities of 30.71 Wh kg−1 and 1875 Wh kg−1.237 Especially when applied in portable electronics, PSCs may have to operate under complex illumination conditions, such as fluctuating sunlight. By creating high-performance supercapacitors, nanostructured electrode materials have outlined tremendous electrochemical characteristics. The improvements in performance and lifespan could be achieved by optimizing the integration of perovskite materials into existing energy storage systems, particularly for energy-harvesting applications, which emphasize the need for scalable and affordable solutions (Table 6). The implications of this research are quite significant, both from a theoretical and a practical point of view, and help to understand the charge storage and transport mechanisms within perovskite materials.238
Table 6 The MHP-based electrodes in electrochemical supercapacitors and photo-induced supercapacitors
Perovskites in electrodes |
Specific capacitance [mF cm−2] |
Energy density |
Capacitance retention |
Reference |
MAPbBr3 |
507 [F g−1] |
Power density 764 W kg−1 |
— |
239 |
MAPbBr3 |
98.36 [F g−1] |
— |
93% after 2000 cycles |
240 |
2D BA2PbBr4 |
148.25 [F g−1] |
— |
98% after 2000 cycles |
240 |
Quasi-2D BA2MAPb2Br7 |
138.35 [F g−1] |
— |
96% after 2000 cycles |
240 |
CsPbBr2.9I0.1 |
150 |
— |
— |
241 |
MAPbI3 |
432 |
34.2 Wh kg−1 |
— |
242 |
CsPbI3 |
7.23 |
— |
65.5% after 1000 cycles |
243 |
MAPbI3 |
422 |
— |
70% after 500 cycles |
237 |
CdS-MAPI3 |
372 |
23.8 |
96.6% after 4000 cycles |
244 |
MBI |
350 |
2.98 μF cm−2 |
94.79% after 5000 cycles |
245 |
In summary, the improved electrochemical performance of SCs using perovskite materials is directly linked to their distinct structural and compositional benefits, pointing to a clear path for the advancement of energy storage solutions. Perovskites continue to be explored for use in supercapacitors, yet several challenges have surfaced that keep them from being fully used in actual applications. Though the creation of perovskite materials has shown promise, like better electrochemical performance and energy density, some key problems still need to be taken into account. For instance, how stable are perovskite SCs when the environment changes, like with different humidity or temperatures. While they seem to cycle better than older electrode materials, they can still be destroyed over time.246
Furthermore, empirical data reveal that perovskite films often have defects and grain boundaries despite improvements in material design, which can harm charge mobility and performance.231 Prior research supports this, suggesting that refining synthetic techniques is key to reducing defects and improving the electrochemical stability of perovskite systems. Thus, research into perovskite materials for SCs keeps enriching the academic world but also points out the many complexities that must be addressed to reach their full potential. Future research should closely match real-world applications, ensuring that the knowledge gained leads to real progress in sustainable energy technologies.246
On the other hand, research toward perovskites' photo-induced functionalities demonstrates how light-induced impacts could enhance energy storage mechanisms in supercapacitors by synergistic interactions between light absorption and the charge storage qualities of perovskite materials.234 The oxygen vacancies in the perovskite are deemed to remain the charge storage region of the pseudo capacitance. Charge storage of oxygen intercalation and energy densities of perovskite supercapacitors will be enhanced by expanding the vacancies of oxygen.
Mechanisms of photo-induced charge generation and understanding how perovskites generate charge when exposed to light are super important for boosting. Basically, when perovskites absorb light, electrons jump from one energy level (the VB) to a higher one (the CB), and this creates electron–hole pairs. How long these charge carriers stick around depends on the material's band gap, how crystalline it is, and its interface properties. For the best energy storage, there is a need for good charge separation and transport, which lets you capture and release energy in a more dynamic way. Getting a handle on these complex mechanisms is key to using perovskites in energy devices. As we keep researching, tweaking the design and makeup of perovskite materials might solve current problems and help them become more useful in photo-induced SCs.
6. Conclusion and future perspective
MHPs have come a long way from being niche materials for photovoltaics to now a dynamic platform with the potential of revolutionizing energy conversion and storage technologies. Their standout properties, like high ionic conductivity, adjustable optoelectronic properties, expansive surface areas, and structural flexibility, have opened doors for their use in LIBs, SCs, ASEIs, and integrated photo-rechargeable systems. Acting as electrodes and interfacial layers, MHPs enhance energy storage and conversion performance. By advancing compositional engineering, surface passivation, and interface design, perovskite-based devices have been able to achieve improved operational stability and functional versatility, allowing for moving beyond proof-of-concept. Combining carbon with perovskite in hybrid electrodes can enhance electronic conductivity, while approaches like compositional and bandgap engineering (for instance, partially substituting Pb2+ with Sn2+) may provide avenues to optimize charge carrier mobility and address toxicity issues. However, as this review points out, there are still several hurdles that need to be overcome before perovskite-based energy technologies can hit the market. These challenges include long-term operational instability when faced with real-world stressors, environmental and health issues tied to lead-containing systems, difficulties in scaling up high-quality large-area films, and the use of toxic or hard-to-handle solvents in material synthesis. To address these issues, not only will materials chemistry have to be bettered gradually, but processing methods will also need to be integrated, sustainable, and scalable. Ultimately, MHPs offer a remarkable opportunity to create self-sustaining, multifunctional energy platforms by integrating photovoltaics with energy storage. The following should be considered when prioritizing future research:
(1) Environmentally friendly chemistries: lead-free MHP formulations that maintain electronic and ionic performance while ensuring long-term chemical stability need to be explored further. Compositional bandgap engineering requires the partial replacement of Pb2+ with Sn2+ or with other stable metals and surface passivation techniques (by widening the bandgap, it suppresses electron migration).
(2) Scalable and green fabrication methods: pushing forward with solvent-free or hybrid deposition strategies that work well with industrial roll-to-roll processing, ensuring everything stays uniform and reproducible over large areas. Undisputedly, a significant mission in the preparation of MHPs and deposition procedures of perovskite thin film also lies in translating the lab-scale technique to industrially applicable manufacturing methods.
(3) Dimensionality and microstructural engineering: the study of the influence of modifying grain size, controlling defects, optimizing phase transitions, and nano-structuring on the electrochemical performance is required. Moreover, the enhanced electrochemical performance of low-dimensional MHPs in energy storage systems should be explored further.
(4) Solid-electrolyte: development of advanced methods of dense perovskite-based solid electrolytes, which would lower impedance and boost performance.
(5) Li-transport mechanisms: understanding of Li-ion movement through vacancy, interstitial, or exchange routes within perovskite frameworks is essential for facilitating uniform ion flow and even lithium deposition.
(6) Interface study: diving into advanced characterization and computational modelling to get a grip on defect formation, ion migration, and interfacial degradation, all while keeping it real under actual operating conditions.
(7) ASEI: the MHP interfacial layer protection approach could open a promising avenue for shielding lithium metal from the liquid electrolyte-Li-ion transport gradient layer model.
(8) Fluorinated MHPs: future research should aim at creating open-framework perovskite-derived electrodes and SEIs with fluorinated heterogeneous nanodomains to improve stability and prevent dendrite growth in LIBs.
(9) Circular economy integration: setting up closed-loop recycling protocols for MHP-based devices to tackle end-of-life management, especially for those lead-containing systems. Also, integrate sustainable supply chain practices.
(10) Collaborative pathways: promote interdisciplinary collaboration across materials science, electrochemistry, and device engineering to achieve stable, efficient, and scalable next-generation perovskite-based energy storage solutions.
To conclude, MHPs are poised to redefine energy harvesting, storage, and optoelectronics. In order to transform the energy landscape, MHP-based technologies must strategically tackle current challenges and embrace interdisciplinary innovations. We aim to be both a comprehensive reference and a springboard for such forward-thinking research directions. In turn, profound insights offered by the scientists in the near future will pave the way for discoveries and developments in fertile landscapes of halide perovskite materials and energy storage systems. With profound insights offered by the scientists in the future, further exciting discoveries are certainly to be realized.
Conflicts of interest
There is no conflict of interest.
Data availability
No primary research results, software or code have been included and no new data were generated or analysed as part of this review.
Acknowledgements
As a part of the DESTINY PhD program, this publication is acknowledged by funding from the European Union's Horizon 2020 research and innovation program under the Marie Skłodowska-Curie Actions COFUND (Grant Agreement #945357). The authors also gratefully acknowledge the National Science Centre, Poland, Grants OPUS 2021/41/B/ST5/04450 (W. W. and M. K. K.) and MAESTRO 11, No. 2019/34/A/ST5/00416 (J. L., M. S.), for financial support.
References
- A. T. Nguyen, V. D. Phung, V. O. Mittova, H. D. Ngo, T. N. Vo, M. L. Le Thi, V. H. Nguyen, I. Y. Mittova, M. L. P. Le, Y. N. Ahn, I. T. Kim and T. L. Nguyen, Fabricating nanostructured HoFeO3 perovskite for lithium-ion battery anodes via co-precipitation, Scr. Mater., 2022, 207, 114259 CrossRef CAS.
- Q. Wei, F. Xiong, S. Tan, L. Huang, E. H. Lan, B. Dunn and L. Mai, Porous One-Dimensional Nanomaterials: Design, Fabrication and Applications in Electrochemical Energy Storage, Adv. Mater., 2017, 29, 1602300 CrossRef PubMed.
- C.-Y. Wang, T. Liu, X.-G. Yang, S. Ge, N. V. Stanley, E. S. Rountree, Y. Leng and B. D. McCarthy, Fast charging of energy-dense lithium-ion batteries, Nature, 2022, 611, 485–490 CrossRef CAS PubMed.
- B. Nykvist and M. Nilsson, Rapidly falling costs of battery packs for electric vehicles, Nat. Clim. Change, 2015, 5, 329–332 CrossRef.
- S. Yang, Q. Liang, H. Wu, J. Pi, Z. Wang, Y. Luo, Y. Liu, Z. Long, D. Zhou, Y. Wen, Q. Wang, J. Guo and J. Qiu, Lead-Free Double Perovskite Cs 2 NaErCl6 : Li+ as High-Stability Anodes for Li-Ion Batteries, J. Phys. Chem. Lett., 2022, 13, 4981–4987 CrossRef CAS PubMed.
- C. P. Grey and D. S. Hall, Prospects for lithium-ion batteries and beyond—a 2030 vision, Nat. Commun., 2020, 11, 6279 CrossRef CAS PubMed.
- R. K. Sahu, S. Gangil, V. K. Bhargav, P. Sahu and B. Ghritalahre, Synthesizing biomass into nano carbon for use in high-performance supercapacitors - A brief critical review, J. Energy Storage, 2023, 72, 108348 CrossRef.
- W. Zhang, G. E. Eperon and H. J. Snaith, Metal halide perovskites for energy applications, Nat. Energy, 2016, 1, 16048 CrossRef CAS.
- J. Y. Kim, J.-W. Lee, H. S. Jung, H. Shin and N.-G. Park, High-Efficiency Perovskite Solar Cells, Chem. Rev., 2020, 120, 7867–7918 CrossRef CAS PubMed.
- C. Yang, W. Hu, J. Liu, C. Han, Q. Gao, A. Mei, Y. Zhou, F. Guo and H. Han, Achievements, challenges, and future prospects for industrialization of perovskite solar cells, Light: Sci. Appl., 2024, 13, 227 CrossRef CAS PubMed.
- J. Han, K. Park, S. Tan, Y. Vaynzof, J. Xue, E. W.-G. Diau, M. G. Bawendi, J.-W. Lee and I. Jeon, Perovskite solar cells, Nat. Rev. Methods Primers, 2025, 5, 3 CrossRef CAS.
- L. Yuan, Q. Xue, F. Wang, N. Li, G. I. N. Waterhouse, C. J. Brabec, F. Gao and K. Yan, Perovskite Solar Cells and Light Emitting Diodes: Materials Chemistry, Device Physics and Relationship, Chem. Rev., 2025, 125, 5057–5162 CrossRef CAS PubMed.
- L. Liang, T. Ma, Z. Chen, J. Wang, J. Hu, Y. Ji, W. Shen and J. Chen, Patterning Technologies for Metal Halide Perovskites: A Review, Adv. Mater. Technol., 2023, 8, 2200419 CrossRef.
- M. A. Green, E. D. Dunlop, M. Yoshita, N. Kopidakis, K. Bothe, G. Siefer, X. Hao and J. Y. Jiang, Solar Cell Efficiency Tables (Version 65), Prog. Photovoltaics Res. Appl., 2025, 33, 3–15 CrossRef.
- Z.-A. Nan, L. Chen, Q. Liu, S.-H. Wang, Z.-X. Chen, S.-Y. Kang, J.-B. Ji, Y.-Y. Tan, Y. Hui, J.-W. Yan, Z.-X. Xie, W.-Z. Liang, B.-W. Mao and Z.-Q. Tian, Revealing phase evolution mechanism for stabilizing formamidinium-based lead halide perovskites by a key intermediate phase, Chem, 2021, 7, 2513–2526 CAS.
- Z. Liang, Y. Zhang, H. Xu, W. Chen, B. Liu, J. Zhang, H. Zhang, Z. Wang, D.-H. Kang, J. Zeng, X. Gao, Q. Wang, H. Hu, H. Zhou, X. Cai, X. Tian, P. Reiss, B. Xu, T. Kirchartz, Z. Xiao, S. Dai, N.-G. Park, J. Ye and X. Pan, Homogenizing out-of-plane cation composition in perovskite solar cells, Nature, 2023, 624, 557–563 CrossRef CAS PubMed.
- J. Kim, J. Park, J. Lim, J. Kim, J. Kim, N. Shin, J. S. Yun, J. Im and S. Il Seok, Susceptible organic cations enable stable and efficient perovskite solar cells, Joule, 2025, 9, 101879 CrossRef CAS.
- M. Cheng, Y. Duan, D. Zhang, Z. Xie, H. Li, Q. Cao, Z. Qiu, Y. Chen and Q. Peng, Tailoring Buried Interface and Minimizing Energy Loss Enable Efficient Narrow and Wide Bandgap Inverted Perovskite Solar Cells by Aluminum Glycinate Based Organometallic Molecule, Adv. Mater., 2025, 37, 2419413 CrossRef CAS PubMed.
- M. Saliba, T. Matsui, J.-Y. Seo, K. Domanski, J.-P. Correa-Baena, M. K. Nazeeruddin, S. M. Zakeeruddin, W. Tress, A. Abate, A. Hagfeldt and M. Grätzel, Cesium-containing triple cation perovskite solar cells: improved stability, reproducibility and high efficiency, Energy Environ. Sci., 2016, 9, 1989–1997 RSC.
- A. Kojima, K. Teshima, Y. Shirai and T. Miyasaka, Organometal Halide Perovskites as Visible-Light Sensitizers for Photovoltaic Cells, J. Am. Chem. Soc., 2009, 131, 6050–6051 CrossRef CAS PubMed.
- S. Liu, J. Li, W. Xiao, R. Chen, Z. Sun, Y. Zhang, X. Lei, S. Hu, M. Kober-Czerny, J. Wang, F. Ren, Q. Zhou, H. Raza, Y. Gao, Y. Ji, S. Li, H. Li, L. Qiu, W. Huang, Y. Zhao, B. Xu, Z. Liu, H. J. Snaith, N.-G. Park and W. Chen, Buried interface molecular hybrid for inverted perovskite solar cells, Nature, 2024, 632, 536–542 CrossRef PubMed.
- X. Zhang, S. Wu, H. Zhang, A. K. Y. Jen, Y. Zhan and J. Chu, Advances in inverted perovskite solar cells, Nat. Photonics, 2024, 18, 1243–1253 CrossRef CAS.
- N. Ren, L. Tan, M. Li, J. Zhou, Y. Ye, B. Jiao, L. Ding and C. Yi, 25% - Efficiency flexible perovskite solar cells via controllable growth of SnO 2, iEnergy, 2024, 3, 39–45 Search PubMed.
- Y. Gao, K. Huang, C. Long, Y. Ding, J. Chang, D. Zhang, L. Etgar, M. Liu, J. Zhang and J. Yang, Flexible Perovskite Solar Cells: From Materials and Device Architectures to Applications, ACS Energy Lett., 2022, 7, 1412–1445 CrossRef CAS.
- Y. Cheng and L. Ding, Pushing commercialization of perovskite solar cells by improving their intrinsic stability, Energy Environ. Sci., 2021, 14, 3233–3255 RSC.
- H. Zhu, S. Teale, M. N. Lintangpradipto, S. Mahesh, B. Chen, M. D. McGehee, E. H. Sargent and O. M. Bakr, Long-term operating stability in perovskite photovoltaics, Nat. Rev. Mater., 2023, 8, 569–586 CrossRef.
- P. Zhu, C. Chen, J. Dai, Y. Zhang, R. Mao, S. Chen, J. Huang and J. Zhu, Toward the Commercialization of Perovskite Solar Modules, Adv. Mater., 2024, 36, 2307357 CrossRef CAS.
- C. Jing, Z. Lin, Y. Wu and X. Ouyang, Air-Processed Perovskite Solar Cells: Challenges, Progress, and Industrial Strategies, Small, 2025, 2504448 CrossRef CAS PubMed.
- X.-K. Liu, W. Xu, S. Bai, Y. Jin, J. Wang, R. H. Friend and F. Gao, Metal halide perovskites for light-emitting diodes, Nat. Mater., 2021, 20, 10–21 CrossRef CAS PubMed.
- D. Chen, G. Zou, Y. Wu, B. Tang, A. L. Rogach and H. Yip, Metal Halide Perovskite LEDs for Visible Light Communication and Lasing Applications, Adv. Mater., 2024, 37, 2414745 CrossRef PubMed.
- Y. Fu, H. Zhu, J. Chen, M. P. Hautzinger, X.-Y. Zhu and S. Jin, Metal halide perovskite nanostructures for optoelectronic applications and the study of physical properties, Nat. Rev. Mater., 2019, 4, 169–188 CrossRef CAS.
- H. Dong, C. Ran, W. Gao, M. Li, Y. Xia and W. Huang, Metal Halide Perovskite for next-generation optoelectronics: progresses and prospects, eLight, 2023, 3, 3 CrossRef.
- L. A. Muscarella and E. M. Hutter, Halide Double-Perovskite Semiconductors beyond Photovoltaics, ACS Energy Lett., 2022, 7, 2128–2135 CrossRef CAS.
- X. Guan, Z. Lei, X. Yu, C. Lin, J. Huang, C. Huang, L. Hu, F. Li, A. Vinu, J. Yi and T. Wu, Low-Dimensional Metal-Halide Perovskites as High-Performance Materials for Memory Applications, Small, 2022, 18, 2203311 CrossRef CAS PubMed.
- W. Peixiong, C. Xiang, P. Xiaoxin, J. Bowen, L. Xiaoqing, P. Yanqin, P. Jie, T. Li, D. Jinxia, Z. Jun and W. Hao, A mini review of recent progress on halide perovskite memristor devices: materials science, challenges and applications, Mater. Today Energy, 2024, 45, 101692 CrossRef.
- B. Jiang, X. Chen, X. Pan, L. Tao, Y. Huang, J. Tang, X. Li, P. Wang, G. Ma, J. Zhang and H. Wang, Advances in Metal Halide Perovskite Memristors: A Review from a Co-Design Perspective, Adv. Sci., 2024, 12, 2409291 CrossRef PubMed.
- W. Xu, H. Cho, Y. Kim, Y. Kim, C. Wolf, C. Park and T. Lee, Organometal Halide Perovskite Artificial Synapses, Adv. Mater., 2016, 28, 5916–5922 CrossRef CAS PubMed.
- M. Yu, X. Ren, L. Ma and Y. Wu, Integrating a redox-coupled dye-sensitized photoelectrode into a lithium–oxygen battery for photoassisted charging, Nat. Commun., 2014, 5, 5111 CrossRef CAS PubMed.
- X. Liu, Y. Yuan, J. Liu, B. Liu, X. Chen, J. Ding, X. Han, Y. Deng, C. Zhong and W. Hu, Utilizing solar energy to improve the oxygen evolution reaction kinetics in zinc–air battery, Nat. Commun., 2019, 10, 4767 CrossRef PubMed.
- J. Han, S. Lee, C. Youn, J. Lee, Y. Kim and T. Choi, Hybrid photoelectrochemical-rechargeable seawater battery for efficient solar energy storage systems, Electrochim. Acta, 2020, 332, 135443 CrossRef CAS.
- J. Xu, Y. Chen and L. Dai, Efficiently photo-charging lithium-ion battery by perovskite solar cell, Nat. Commun., 2015, 6, 8103 CrossRef CAS PubMed.
- H. Xue, H. Gong, Y. Yamauchi, T. Sasaki and R. Ma, Photo-enhanced rechargeable high-energy-density metal batteries for solar energy conversion and storage, Nano Res. Energy, 2022, 1, e9120007 CrossRef.
- H.-R. Xia, W.-T. Sun and L.-M. Peng, Hydrothermal synthesis of organometal halide perovskites for Li-ion batteries, Chem. Commun., 2015, 51, 13787–13790 RSC.
- X.-H. Wu, M.-J. Zhao, Y. Chai, Z. Liu, W.-J. Jiang, L.-B. Yang, B.-J. Feng, J.-J. Liu, Q. Yu, K.-Z. Du and Y. Zhao, Unlocking the atomic-scale mechanism of structural evolutions during (de)lithiation and negative-fading in CsPbBr3 anodes, Energy Storage Mater., 2025, 75, 104043 CrossRef.
- P. Pal and A. Ghosh, Three-Dimensional CsPbCl3 Perovskite Anode for Quasi-Solid-State Li-Ion and Dual-Ion Batteries: Mechanism of Li+ Conversion Process in Perovskite, Phys. Rev. Appl., 2020, 14, 064010 CrossRef CAS.
- N. Kaisar, T. Paul, P.-W. Chi, Y.-H. Su, A. Singh, C.-W. Chu, M.-K. Wu and P. M. Wu, Electrochemical Performance of Orthorhombic CsPbI3 Perovskite in Li-Ion Batteries, Materials, 2021, 14, 5718 CrossRef CAS PubMed.
- Y.-C. Yin, Q. Wang, J.-T. Yang, F. Li, G. Zhang, C.-H. Jiang, H.-S. Mo, J.-S. Yao, K.-H. Wang, F. Zhou, H.-X. Ju and H.-B. Yao, Metal chloride perovskite thin film-based interfacial layer for shielding lithium metal from liquid electrolyte, Nat. Commun., 2020, 11, 1761 CrossRef CAS PubMed.
- R. Liu, W. Feng, L. Fang, H. Deng, L. Lin, M. Chen, J.-X. Zhong and W. Yin, An ultrathin Li-doped perovskite SEI film with high Li ion flux for a fast charging lithium metal battery, Energy Adv., 2024, 3, 2999–3006 RSC.
- A. Sandhu and M. K. Chini, 2D and 3D Halide Perovskite-Based Supercapacitors, ChemistrySelect, 2024, 9, e202304441 CrossRef CAS.
- M. Riaz, S. M. Ali, R. Alotaibi, S. D. Ali and A. Mehmood, Mesoporous structure of tin (Sn) based inorganic halide perovskite CsSnBr3 binary composite with PANI and rGO for energy storage as supercapacitor electrodes, Inorg. Chem. Commun., 2025, 176, 114338 CrossRef CAS.
- X. Zhang, W. Song, J. Tu, J. Wang, M. Wang and S. Jiao, A Review of Integrated Systems Based on Perovskite Solar Cells and Energy Storage Units: Fundamental, Progresses, Challenges, and Perspectives, Adv. Sci., 2021, 8, 2100552 CrossRef CAS PubMed.
- Y. Chen, Z. Yue, S.-W. Tsang and Y. Cheng, Metal halide perovskites for efficient solar energy conversion and storage systems: Principles, recent advances, challenges and prospects, Nano Energy, 2025, 137, 110782 CrossRef CAS.
- F. Li, Y. Liang and R. Zheng, A balanced view of ion migration in halide perovskite electronics, Newton, 2025, 1, 100096 CrossRef.
- T. Oku, Crystal structures of perovskite halide compounds used for solar cells, Rev. Adv. Mater. Sci., 2020, 59, 264–305 CAS.
- H. Tanaka, T. Oku and N. Ueoka, Structural stabilities of organic–inorganic perovskite crystals, Jpn. J. Appl. Phys., 2018, 57, 08RE12 CrossRef.
- Z. Li, M. Yang, J.-S. Park, S.-H. Wei, J. J. Berry and K. Zhu, Stabilizing Perovskite Structures by Tuning Tolerance Factor: Formation of Formamidinium and Cesium Lead Iodide Solid-State Alloys, Chem. Mater., 2016, 28, 284–292 CrossRef CAS.
- R. Kour, S. Arya, S. Verma, J. Gupta, P. Bandhoria, V. Bharti, R. Datt and V. Gupta, Potential Substitutes for Replacement of Lead in Perovskite Solar Cells: A Review, Global Challenges, 2019, 3, 1900050 CrossRef PubMed.
- Y. Wang, J. Ren, X. Zhou and G. Zhang, Stability improvements of metal halide perovskite nanocrystals and their optoelectrical applications, Mater. Chem. Front., 2023, 7, 2175–2207 RSC.
- W. Gao, C. Chen, C. Ran, H. Zheng, H. Dong, Y. Xia, Y. Chen and W. Huang, A-Site Cation Engineering of Metal Halide Perovskites: Version 3.0 of Efficient Tin-Based Lead-Free Perovskite Solar Cells, Adv. Funct. Mater., 2020, 30, 2000794 CrossRef CAS.
- D. Duan, C. Ge, M. Z. Rahaman, C.-H. Lin, Y. Shi, H. Lin, H. Hu and T. Wu, Recent progress with one-dimensional metal halide perovskites: from rational synthesis to optoelectronic applications, NPG Asia Mater., 2023, 15, 8 CrossRef CAS.
- T. Ibn-Mohammed, S. C. L. Koh, I. M. Reaney, A. Acquaye, G. Schileo, K. B. Mustapha and R. Greenough, Perovskite solar cells: An integrated hybrid lifecycle assessment and review in comparison with other photovoltaic technologies, Renewable Sustainable Energy Rev., 2017, 80, 1321–1344 CrossRef CAS.
- J. Huang, Y. Yuan, Y. Shao and Y. Yan, Understanding the physical properties of hybrid perovskites for photovoltaic applications, Nat. Rev. Mater., 2017, 2, 17042 CrossRef CAS.
- J.-W. Lee, S. Tan, S. Il Seok, Y. Yang and N.-G. Park, Rethinking the A cation in halide perovskites, Science, 2022, 375 Search PubMed.
- M. G. Goesten and R. Hoffmann, Mirrors of Bonding in Metal Halide Perovskites, J. Am. Chem. Soc., 2018, 140, 12996–13010 CrossRef CAS PubMed.
- J. Han, Y. Li, P. Zhang, B. Xu, X. Xu and Z. Quan, Cooperative Regulation of ns2 Lone-Pair Expression Realizes Distinct Excitonic Emissions in Hybrid Germanium, Tin, and Lead Halides, J. Am. Chem. Soc., 2025, 147, 1291–1299 CrossRef CAS PubMed.
- I. Y. H Chang, C. H. Park and K. Matsuishi, First-Principles Study of the Structural and the Electronic Properties of the Lead-Halide-Based Inorganic-Organic Perovskites, Inorg. Chem., 2019, 58, 4134–4140 CrossRef.
- K. J. Savill, A. M. Ulatowski and L. M. Herz, Optoelectronic Properties of Tin–Lead Halide Perovskites, ACS Energy Lett., 2021, 6, 2413–2426 CrossRef CAS.
- M. Awais, R. L. Kirsch, V. Yeddu and M. I. Saidaminov, Tin Halide Perovskites Going Forward: Frost Diagrams Offer Hints, ACS Mater. Lett., 2021, 3, 299–307 CrossRef CAS.
- Y. Zhang, J. D. Lin, V. Vijayaragavan, K. K. Bhakoo and T. T. Y. Tan, Tuning sub-10 nm single-phase NaMnF3 nanocrystals as ultrasensitive hosts for pure intense fluorescence and excellent T1 magnetic resonance imaging, Chem. Commun., 2012, 48, 10322 RSC.
- C. W. Myung, J. Yun, G. Lee and K. S. Kim, A New Perspective on the Role of A-Site Cations in Perovskite Solar Cells, Adv. Energy Mater., 2018, 8, 1702898 CrossRef.
- G. E. Eperon, S. D. Stranks, C. Menelaou, M. B. Johnston, L. M. Herz and H. J. Snaith, Formamidinium lead trihalide: a broadly tunable perovskite for efficient planar heterojunction solar cells, Energy Environ. Sci., 2014, 7, 982 RSC.
- P. Wu, D. Li, S. Wang and F. Zhang, Magic guanidinium cations in perovskite solar cells: from bulk to interface, Mater. Chem. Front., 2023, 7, 2507–2527 RSC.
- D. J. Kubicki, D. Prochowicz, A. Hofstetter, M. Saski, P. Yadav, D. Bi, N. Pellet, J. Lewiński, S. M. Zakeeruddin, M. Grätzel and L. Emsley, Formation of Stable Mixed Guanidinium–Methylammonium Phases with Exceptionally Long Carrier Lifetimes for High-Efficiency Lead Iodide-Based Perovskite Photovoltaics, J. Am. Chem. Soc., 2018, 140, 3345–3351 CrossRef CAS PubMed.
- J. Qin, Z. Che, Y. Kang, C. Liu, D. Wu, H. Yang, X. Hu and Y. Zhan, Towards operation-stabilizing perovskite solar cells: Fundamental materials, device designs, and commercial applications, InfoMat, 2024, 6, e12522 CrossRef CAS.
- C. C. Stoumpos, C. D. Malliakas and M. G. Kanatzidis, Semiconducting Tin and Lead Iodide Perovskites with Organic Cations: Phase Transitions, High Mobilities, and Near-Infrared Photoluminescent Properties, Inorg. Chem., 2013, 52, 9019–9038 CrossRef CAS PubMed.
- D. M. Trots and S. V. Myagkota, High-temperature structural evolution of caesium and rubidium triiodoplumbates, J. Phys. Chem. Solids, 2008, 69, 2520–2526 CrossRef CAS.
- R. Panetta, G. Righini, M. Colapietro, L. Barba, D. Tedeschi, A. Polimeni, A. Ciccioli and A. Latini, Azetidinium lead iodide: synthesis, structural and physico-chemical characterization, J. Mater. Chem. A, 2018, 6, 10135–10148 RSC.
- M. Maçzka, M. Ptak, A. Gągor, D. Stefańska, J. K. Zaręba and A. Sieradzki, Methylhydrazinium Lead Bromide: Noncentrosymmetric Three-Dimensional Perovskite with Exceptionally Large Framework Distortion and Green Photoluminescence, Chem. Mater., 2020, 32, 1667–1673 CrossRef.
- C. Yi, J. Luo, S. Meloni, A. Boziki, N. Ashari-Astani, C. Grätzel, S. M. Zakeeruddin, U. Röthlisberger and M. Grätzel, Entropic stabilization of mixed A-cation ABX3 metal halide perovskites for high performance perovskite solar cells, Energy Environ. Sci., 2016, 9, 656–662 RSC.
- D. Drozdowski, A. Gągor and M. Mączka, Methylhydrazinium lead iodide – one dimensional chain phase with excitonic absorption and large energy band gap, J. Mol. Struct., 2022, 1249, 131660 CrossRef CAS.
- Y. Zhang, G. Grancini, Z. Fei, E. Shirzadi, X. Liu, E. Oveisi, F. F. Tirani, R. Scopelliti, Y. Feng, M. K. Nazeeruddin and P. J. Dyson, Auto-passivation of crystal defects in hybrid imidazolium/methylammonium lead iodide films by fumigation with methylamine affords high efficiency perovskite solar cells, Nano Energy, 2019, 58, 105–111 CrossRef CAS.
- K. Opała, M. Saski, W. Marynowski, A. Borkenhagen and J. Lewiński, Exploring Polytype Formation in Double-Cation DMAxFA1– xPb3 Perovskites: Compositional Engineering and Phase Transitions, Chem. Mater., 2025, 37, 897–911 CrossRef.
- M. Šimėnas, S. Balčiūnas, A. Gągor, A. Pieniążek, K. Tolborg, M. Kinka, V. Klimavicius, Š. Svirskas, V. Kalendra, M. Ptak, D. Szewczyk, A. P. Herman, R. Kudrawiec, A. Sieradzki, R. Grigalaitis, A. Walsh, M. Maçzka and J. Banys, Mixology of MA1– xEAxPbI3 Hybrid Perovskites: Phase Transitions, Cation Dynamics, and Photoluminescence, Chem. Mater., 2022, 34, 10104–10112 CrossRef PubMed.
- P. Singh, R. Mukherjee and S. Avasthi, Acetamidinium-Substituted Methylammonium Lead Iodide Perovskite Solar Cells with Higher Open-Circuit Voltage and Improved Intrinsic Stability, ACS Appl. Mater. Interfaces, 2020, 12, 13982–13987 CrossRef CAS.
- M. Nishat, M. K. Hossain, M. R. Hossain, S. Khanom, F. Ahmed and M. A. Hossain, Role of metal and anions in organo-metal halide perovskites CH3NH3MX3 (M: Cu, Zn, Ga, Ge, Sn, Pb; X: Cl, Br, I) on structural and optoelectronic properties for photovoltaic applications, RSC Adv., 2022, 12, 13281–13294 RSC.
- Y. Liang, F. Li, X. Cui, C. Stampfl, S. P. Ringer, X. Yang, J. Huang and R. Zheng, Multiple B-site doping suppresses ion migration in halide perovskites, Sci. Adv., 2025, 11, eads7054 CrossRef CAS PubMed.
- M. B. Gray, E. T. McClure and P. M. Woodward, Cs2AgBiBr6−xClx solid solutions – band gap engineering with halide double perovskites, J. Mater. Chem. C, 2019, 7, 9686–9689 RSC.
- A. H. Slavney, T. Hu, A. M. Lindenberg and H. I. Karunadasa, A Bismuth-Halide Double Perovskite with Long Carrier Recombination Lifetime for Photovoltaic Applications, J. Am. Chem. Soc., 2016, 138, 2138–2141 CrossRef CAS PubMed.
- G. García-Espejo, D. Rodríguez-Padrón, R. Luque, L. Camacho and G. de Miguel, Mechanochemical synthesis of three double perovskites: Cs2AgBiBr6 , (CH3NH3)2TlBiBr6 and Cs2AgSbBr6, Nanoscale, 2019, 11, 16650–16657 RSC.
- W. Meng, X. Wang, Z. Xiao, J. Wang, D. B. Mitzi and Y. Yan, Parity-Forbidden Transitions and Their Impact on the Optical Absorption Properties of Lead-Free Metal Halide Perovskites and Double Perovskites, J. Phys. Chem. Lett., 2017, 8, 2999–3007 CrossRef CAS PubMed.
- G. Meyer, Halogen-Elpasolithe, VI [1] Erste Iod-Elpasolithe, Cs2I MIIII6 (BI = Li, Na)/Halo-Elpasolites, VI [1] The First Iodo-Elpasolites, Cs 2 B I MIII I6 ( BI = Li, Na), Z. Naturforsch., B, 1980, 35, 394–396 CrossRef.
- Z. Deng, F. Wei, F. Brivio, Y. Wu, S. Sun, P. D. Bristowe and A. K. Cheetham, Synthesis and Characterization of the Rare-Earth Hybrid Double Perovskites: (CH3NH3)2 KGdCl6 and (CH3NH3 )2KYCl6, J. Phys. Chem. Lett., 2017, 8, 5015–5020 CrossRef CAS PubMed.
- G. T. Kent, E. Morgan, K. R. Albanese, A. Kallistova, A. Brumberg, L. Kautzsch, G. Wu, P. Vishnoi, R. Seshadri and A. K. Cheetham, Elusive Double Perovskite Iodides: Structural, Optical, and Magnetic Properties, Angew. Chem., Int. Ed., 2023, 62, e202306000 CrossRef CAS.
- S. Ghosh, H. Shankar and P. Kar, Recent developments of lead-free halide double perovskites: a new superstar in the optoelectronic field, Mater. Adv., 2022, 3, 3742–3765 RSC.
- D. J. Kubicki, M. Saski, S. MacPherson, K. Gałkowski, J. Lewiński, D. Prochowicz, J. J. Titman and S. D. Stranks, Halide Mixing and Phase Segregation in Cs2AgBiX6 (X = Cl, Br, and I) Double Perovskites from Cesium-133 Solid-State NMR and Optical Spectroscopy, Chem. Mater., 2020, 32, 8129–8138 CrossRef CAS PubMed.
- F. Wei, Z. Deng, S. Sun, F. Zhang, D. M. Evans, G. Kieslich, S. Tominaka, M. A. Carpenter, J. Zhang, P. D. Bristowe and A. K. Cheetham, Synthesis and Properties of a Lead-Free Hybrid Double Perovskite: (CH 3 NH 3 ) 2 AgBiBr 6, Chem. Mater., 2017, 29, 1089–1094 CrossRef CAS.
- C. Kupfer, J. Elia, M. Kato, A. Osvet and C. J. Brabec, Mechanochemical Synthesis of Cesium Titanium Halide Perovskites Cs2TiBr6-xIx (x = 0, 2, 4, 6), Cryst. Res. Technol., 2023, 58, 2200150 CrossRef CAS.
- P. Sebastia-Luna, U. Pokharel, B. A. H. Huisman, L. J. A. Koster, F. Palazon and H. J. Bolink, Vacuum-Deposited Cesium Tin Iodide Thin Films with Tunable Thermoelectric Properties, ACS Appl. Energy Mater., 2022, 5, 10216–10223 CrossRef CAS PubMed.
- S.-T. Ha, R. Su, J. Xing, Q. Zhang and Q. Xiong, Metal halide perovskite nanomaterials: synthesis and applications, Chem. Sci., 2017, 8, 2522–2536 RSC.
- A. Soultati, M. Tountas, K. K. Armadorou, A. R. b. M. Yusoff, M. Vasilopoulou and M. K. Nazeeruddin, Synthetic approaches for perovskite thin films and single-crystals, Energy Adv., 2023, 2, 1075–1115 RSC.
- S. Tao, I. Schmidt, G. Brocks, J. Jiang, I. Tranca, K. Meerholz and S. Olthof, Absolute energy level positions in tin- and lead-based halide perovskites, Nat. Commun., 2019, 10, 2560 CrossRef PubMed.
- K. Chen, S. Schünemann, S. Song and H. Tüysüz, Structural effects on optoelectronic properties of halide perovskites, Chem. Soc. Rev., 2018, 47, 7045–7077 RSC.
- A. Alaei, A. Circelli, Y. Yuan, Y. Yang and S. S. Lee, Polymorphism in metal halide perovskites, Mater. Adv., 2021, 2, 47–63 RSC.
- M. Saski, S. Sobczak, P. Ratajczyk, M. Terlecki, W. Marynowski, A. Borkenhagen, I. Justyniak, A. Katrusiak and J. Lewiński, Unprecedented Richness of Temperature- and Pressure-Induced Polymorphism in 1D Lead Iodide Perovskite, Small, 2024, 20, 2403685 CrossRef CAS PubMed.
- M. Simenas, A. Gagor, J. Banys and M. Maczka, Phase Transitions and Dynamics in Mixed Three- and Low-Dimensional Lead Halide Perovskites, Chem. Rev., 2024, 124, 2281–2326 CrossRef CAS PubMed.
- M. T. Weller, O. J. Weber, P. F. Henry, A. M. Di Pumpo and T. C. Hansen, Complete structure and cation orientation in the perovskite photovoltaic methylammonium lead iodide between 100 and 352 K, Chem. Commun., 2015, 51, 4180–4183 RSC.
- R. Shi, Q. Fang, A. S. Vasenko, R. Long, W.-H. Fang and O. V. Prezhdo, Structural Disorder in Higher-Temperature Phases Increases Charge Carrier Lifetimes in Metal Halide Perovskites, J. Am. Chem. Soc., 2022, 144, 19137–19149 CrossRef CAS PubMed.
- H. Jin, Y.-J. Zeng, J. A. Steele, M. B. J. Roeffaers, J. Hofkens and E. Debroye, Phase stabilization of cesium lead iodide perovskites for use in efficient optoelectronic devices, NPG Asia Mater., 2024, 16, 24 CrossRef CAS PubMed.
- A. Marronnier, G. Roma, S. Boyer-Richard, L. Pedesseau, J.-M. Jancu, Y. Bonnassieux, C. Katan, C. C. Stoumpos, M. G. Kanatzidis and J. Even, Anharmonicity and Disorder in the Black Phases of Cesium Lead Iodide Used for Stable Inorganic Perovskite Solar Cells, ACS Nano, 2018, 12, 3477–3486 CrossRef CAS PubMed.
- M. Szafrański and A. Katrusiak, Photovoltaic Hybrid Perovskites under Pressure, J. Phys. Chem. Lett., 2017, 8, 2496–2506 CrossRef PubMed.
- L. Kong, G. Liu, J. Gong, Q. Hu, R. D. Schaller, P. Dera, D. Zhang, Z. Liu, W. Yang, K. Zhu, Y. Tang, C. Wang, S.-H. Wei, T. Xu and H. Mao, Simultaneous band-gap narrowing and carrier-lifetime prolongation of organic–inorganic trihalide perovskites, Proc. Natl. Acad. Sci. U. S. A., 2016, 113, 8910–8915 CrossRef CAS PubMed.
- N. Wang, S. Zhang, S. Wang, X. Yang, F. Guo, Y. Zhang, Z. Gu and Y. Song, Pressure Engineering on Perovskite Structures, Properties, and Devices, Adv. Funct. Mater., 2024, 34, 2315918 CrossRef CAS.
- D. D. Nematov, A. S. Burkhonzoda, M. S. Kurboniyon, U. Zafari, K. T. Kholmurodov, M. G. Brik, T. Yamamoto and F. Shokir, The Effect of Phase Changes on Optoelectronic Properties of Lead-Free CsSnI3 Perovskites, J. Electron. Mater., 2025, 54, 1634–1644 CrossRef CAS.
- K. Yamada, Y. Kuranaga, K. Ueda, S. Goto, T. Okuda and Y. Furukawa, Phase Transition and Electric Conductivity of ASnCl3 (A = Cs and CH3NH3), Bull. Chem. Soc. Jpn., 1998, 71, 127–134 CrossRef CAS.
- F. Hao, C. C. Stoumpos, D. H. Cao, R. P. H. Chang and M. G. Kanatzidis, Lead-free solid-state organic–inorganic halide perovskite solar cells, Nat. Photonics, 2014, 8, 489–494 CrossRef CAS.
- Y. Kawamura, H. Mashiyama and K. Hasebe, Structural Study on Cubic–Tetragonal Transition of CH3NH3PbI3, J. Phys. Soc. Jpn., 2002, 71, 1694–1697 CrossRef CAS.
- H. Mashiyama, Y. Kawamura, E. Magome and Y. Kubota, Displacive character of the cubic-tetragonal transition in CH3NH3PbX3, J. Korean Phys. Soc., 2003, 42, S1026–S1029 CAS.
- A. Poglitsch and D. Weber, Dynamic disorder in methylammoniumtrihalogenoplumbates (II) observed by millimeter-wave spectroscopy, J. Chem. Phys., 1987, 87, 6373–6378 CrossRef CAS.
- E. C. Schueller, G. Laurita, D. H. Fabini, C. C. Stoumpos, M. G. Kanatzidis and R. Seshadri, Crystal Structure Evolution and Notable Thermal Expansion in Hybrid Perovskites Formamidinium Tin Iodide and Formamidinium Lead Bromide, Inorg. Chem., 2018, 57, 695–701 CrossRef CAS PubMed.
- X. Wang, Q. Wang, Z. Chai and W. Wu, The thermal stability of FAPbBr 3 nanocrystals from temperature-dependent photoluminescence and first-principles calculations, RSC Adv., 2020, 10, 44373–44381 RSC.
- R. J. Sutton, M. R. Filip, A. A. Haghighirad, N. Sakai, B. Wenger, F. Giustino and H. J. Snaith, Cubic or Orthorhombic? Revealing the Crystal Structure of Metastable Black-Phase CsPbI3 by Theory and Experiment, ACS Energy Lett., 2018, 3, 1787–1794 CrossRef CAS.
- M. Rodová, J. Brožek, K. Knížek and K. Nitsch, Phase transitions in ternary caesium lead bromide, J. Therm. Anal. Calorim., 2003, 71, 667–673 CrossRef.
- I. Chung, J.-H. Song, J. Im, J. Androulakis, C. D. Malliakas, H. Li, A. J. Freeman, J. T. Kenney and M. G. Kanatzidis, CsSnI3 : Semiconductor or Metal? High Electrical Conductivity and Strong Near-Infrared Photoluminescence from a Single Material. High Hole Mobility and Phase-Transitions, J. Am. Chem. Soc., 2012, 134, 8579–8587 CrossRef CAS PubMed.
- D. E. Scaife, P. F. Weller and W. G. Fisher, Crystal preparation and properties of cesium tin(II) trihalides, J. Solid State Chem., 1974, 9, 308–314 CrossRef CAS.
- G. Thiele, H. Wilhelm Rotter and K. Schmidt, Kristallstrukturen und Phasentransformationen von Caesiumtrihalogenogermanaten(II) CsGeX3 (X = Cl, Br, I), Z. Anorg. Allg. Chem., 1987, 545, 148–156 CrossRef CAS.
- A. Celeste and F. Capitani, Hybrid perovskites under pressure: Present and future directions, J. Appl. Phys., 2022, 132, 220903 CrossRef CAS.
- P. Vishnoi and C. N. R. Rao, Temperature and pressure induced structural transitions of lead iodide perovskites, J. Mater. Chem. A, 2024, 12, 19–37 RSC.
- P. Gratia, I. Zimmermann, P. Schouwink, J.-H. Yum, J.-N. Audinot, K. Sivula, T. Wirtz and M. K. Nazeeruddin, The Many Faces of Mixed Ion Perovskites: Unraveling and Understanding the Crystallization Process, ACS Energy Lett., 2017, 2, 2686–2693 CrossRef CAS.
- S. Shao and M. A. Loi, Advances and Prospective in Metal Halide Ruddlesen–Popper Perovskite Solar Cells, Adv. Energy Mater., 2021, 11, 2003907 CrossRef CAS.
- P. Zhu and J. Zhu, Low-dimensional metal halide perovskites and related optoelectronic applications, InfoMat, 2020, 2, 341–378 CrossRef CAS.
- G. Wang, S. Mei, J. Liao, W. Wang, Y. Tang, Q. Zhang, Z. Tang, B. Wu and G. Xing, Advances of Nonlinear Photonics in Low-Dimensional Halide Perovskites, Small, 2021, 17, 2100809 CrossRef CAS PubMed.
- M. Li, R. Begum, J. Fu, Q. Xu, T. M. Koh, S. A. Veldhuis, M. Grätzel, N. Mathews, S. Mhaisalkar and T. C. Sum, Low threshold and efficient multiple exciton generation in halide perovskite nanocrystals, Nat. Commun., 2018, 9, 4197 CrossRef.
- M. C. Brennan, J. E. Herr, T. S. Nguyen-Beck, J. Zinna, S. Draguta, S. Rouvimov, J. Parkhill and M. Kuno, Origin of the Size-Dependent Stokes Shift in CsPbBr3 Perovskite Nanocrystals, J. Am. Chem. Soc., 2017, 139, 12201–12208 CrossRef CAS PubMed.
- P. Vashishtha, D. Z. Metin, M. E. Cryer, K. Chen, J. M. Hodgkiss, N. Gaston and J. E. Halpert, Shape-, Size-, and Composition-Controlled Thallium Lead Halide Perovskite Nanowires and Nanocrystals with Tunable Band Gaps, Chem. Mater., 2018, 30, 2973–2982 CrossRef CAS.
- K. Leng, I. Abdelwahab, I. Verzhbitskiy, M. Telychko, L. Chu, W. Fu, X. Chi, N. Guo, Z. Chen, Z. Chen, C. Zhang, Q.-H. Xu, J. Lu, M. Chhowalla, G. Eda and K. P. Loh, Molecularly thin two-dimensional hybrid perovskites with tunable optoelectronic properties due to reversible surface relaxation, Nat. Mater., 2018, 17, 908–914 CrossRef CAS PubMed.
- J. Xing, Y. Zhao, M. Askerka, L. N. Quan, X. Gong, W. Zhao, J. Zhao, H. Tan, G. Long, L. Gao, Z. Yang, O. Voznyy, J. Tang, Z.-H. Lu, Q. Xiong and E. H. Sargent, Color-stable highly luminescent sky-blue perovskite light-emitting diodes, Nat. Commun., 2018, 9, 3541 CrossRef PubMed.
- D. Prochowicz, M. Saski, P. Yadav, M. Grätzel and J. Lewiński, Mechanoperovskites for Photovoltaic Applications: Preparation, Characterization, and Device Fabrication, Acc. Chem. Res., 2019, 52, 3233–3243 CrossRef CAS PubMed.
- J. Shamsi, A. S. Urban, M. Imran, L. De Trizio and L. Manna, Metal Halide Perovskite Nanocrystals: Synthesis, Post-Synthesis Modifications, and Their Optical Properties, Chem. Rev., 2019, 119, 3296–3348 CrossRef CAS PubMed.
- Y. Zhang, T. D. Siegler, C. J. Thomas, M. K. Abney, T. Shah, A. De Gorostiza, R. M. Greene and B. A. Korgel, A “Tips and Tricks” Practical Guide to the Synthesis of Metal Halide Perovskite Nanocrystals, Chem. Mater., 2020, 32, 5410–5423 CrossRef CAS.
- R. Vidal, J.-A. Alberola-Borràs, S. N. Habisreutinger, J.-L. Gimeno-Molina, D. T. Moore, T. H. Schloemer, I. Mora-Seró, J. J. Berry and J. M. Luther, Assessing health and environmental impacts of solvents for producing perovskite solar cells, Nat. Sustainability, 2020, 4, 277–285 CrossRef.
- W. Zuo, M. M. Byranvand, T. Kodalle, M. Zohdi, J. Lim, B. Carlsen, T. Magorian Friedlmeier, M. Kot, C. Das, J. I. Flege, W. Zong, A. Abate, C. M. Sutter-Fella, M. Li and M. Saliba, Coordination Chemistry as a Universal Strategy for a Controlled Perovskite Crystallization, Adv. Mater., 2023, 35, 2302889 CrossRef CAS PubMed.
- D. Prochowicz, M. Franckevičius, A. M. Cieślak, S. M. Zakeeruddin, M. Grätzel and J. Lewiński, Mechanosynthesis of the hybrid perovskite CH 3 NH 3 PbI 3 : characterization and the corresponding solar cell efficiency, J. Mater. Chem. A, 2015, 3, 20772–20777 RSC.
- D. Prochowicz, P. Yadav, M. Saliba, D. J. Kubicki, M. M. Tavakoli, S. M. Zakeeruddin, J. Lewiński, L. Emsley and M. Grätzel, One-step mechanochemical incorporation of an insoluble cesium additive for high performance planar heterojunction solar cells, Nano Energy, 2018, 49, 523–528 CrossRef CAS.
- D. Prochowicz, P. Yadav, M. Saliba, M. Saski, S. M. Zakeeruddin, J. Lewiński and M. Grätzel, Mechanosynthesis of pure phase mixed-cation MAxFA1−xPbI3 hybrid perovskites: photovoltaic performance and electrochemical properties, Sustainable Energy Fuels, 2017, 1, 689–693 RSC.
- D. Prochowicz, P. Yadav, M. Saliba, M. Saski, S. M. Zakeeruddin, J. Lewiński and M. Grätzel, Reduction in the Interfacial Trap Density of Mechanochemically Synthesized MAPbI 3, ACS Appl. Mater. Interfaces, 2017, 9, 28418–28425 CrossRef CAS PubMed.
- F. Palazon, Y. El Ajjouri and H. J. Bolink, Making by Grinding: Mechanochemistry Boosts the Development of Halide Perovskites and Other Multinary Metal Halides, Adv. Energy Mater., 2019, 10, 1902499 CrossRef.
- M. F. Mohamad Noh, N. A. Arzaee, I. N. Nawas Mumthas, N. A. Mohamed, S. N. F. Mohd Nasir, J. Safaei, A. R. bin M. Yusoff, M. K. Nazeeruddin and M. A. Mat Teridi, High-humidity processed perovskite solar cells, J. Mater. Chem. A, 2020, 8, 10481–10518 RSC.
- A. Khorasani, F. Mohamadkhani, M. Marandi, H. Luo and M. Abdi-Jalebi, Opportunities, Challenges, and Strategies for Scalable Deposition of Metal Halide Perovskite Solar Cells and Modules, Adv. Energy Sustainability Res., 2024, 5, 2300275 CrossRef CAS.
- J. Burschka, N. Pellet, S.-J. Moon, R. Humphry-Baker, P. Gao, M. K. Nazeeruddin and M. Grätzel, Sequential deposition as a route to high-performance perovskite-sensitized solar cells, Nature, 2013, 499, 316–319 CrossRef CAS.
- A. Ummadisingu and M. Grätzel, Revealing the detailed path of sequential deposition for metal halide perovskite formation, Sci. Adv., 2018, 4, e1701402 CrossRef PubMed.
- A. Ummadisingu, S. Meloni, A. Mattoni, W. Tress and M. Grätzel, Crystal-Size-Induced Band Gap Tuning in Perovskite Films, Angew. Chem., Int. Ed., 2021, 60, 21368–21376 CrossRef CAS PubMed.
- K. J. Prince, H. M. Mirletz, E. A. Gaulding, L. M. Wheeler, R. A. Kerner, X. Zheng, L. T. Schelhas, P. Tracy, C. A. Wolden, J. J. Berry, S. Ovaitt, T. M. Barnes and J. M. Luther, Sustainability pathways for perovskite photovoltaics, Nat. Mater., 2025, 24, 22–33 CrossRef CAS PubMed.
- Y. Zhang, M. Chen, Y. Zhou, W. Li, Y. Lee, H. Kanda, X. Gao, R. Hu, K. G. Brooks, R. Zia, S. Kinge, N. P. Padture and M. K. Nazeeruddin, The Synergism of DMSO and Diethyl Ether for Highly Reproducible and Efficient MA0.5FA0.5PbI3 Perovskite Solar Cells, Adv. Energy Mater., 2020, 10, 2001300 CrossRef CAS.
- J. Küffner, J. Hanisch, T. Wahl, J. Zillner, E. Ahlswede and M. Powalla, One-Step Blade Coating of Inverted Double-Cation Perovskite Solar Cells from a Green Precursor Solvent, ACS Appl. Energy Mater., 2021, 4, 11700–11710 CrossRef.
- H.-S. Yun, H. W. Kwon, M. J. Paik, S. Hong, J. Kim, E. Noh, J. Park, Y. Lee and S. Il Seok, Ethanol-based green-solution processing of α-formamidinium lead triiodide perovskite layers, Nat. Energy, 2022, 7, 828–834 CrossRef CAS.
- J. Ávila, C. Momblona, P. P. Boix, M. Sessolo and H. J. Bolink, Vapor-Deposited Perovskites: The Route to High-Performance Solar Cell Production?, Joule, 2017, 1, 431–442 CrossRef.
- F. U. Kosasih, E. Erdenebileg, N. Mathews, S. G. Mhaisalkar and A. Bruno, Thermal evaporation and hybrid deposition of perovskite solar cells and mini-modules, Joule, 2022, 6, 2692–2734 CrossRef CAS.
- Z. Zhang, R. Ji, M. Kroll, Y. J. Hofstetter, X. Jia, D. Becker-Koch, F. Paulus, M. Löffler, F. Nehm, K. Leo and Y. Vaynzof, Efficient Thermally Evaporated γ-CsPbI3 Perovskite Solar Cells, Adv. Energy Mater., 2021, 11, 2100299 CrossRef CAS.
- H. Li, J. Zhou, L. Tan, M. Li, C. Jiang, S. Wang, X. Zhao, Y. Liu, Y. Zhang, Y. Ye, W. Tress and C. Yi, Sequential vacuum-evaporated perovskite solar cells with more than 24% efficiency, Sci. Adv., 2022, 8, eabo7422 CrossRef PubMed.
- A. Singh, J. Hieulle, J. F. Machado, S. Gharabeiki, W. Zuo, M. U. Farooq, H. Phirke, M. Saliba and A. Redinger, Coevaporation Stabilizes Tin-Based Perovskites in a Single Sn-Oxidation State, Nano Lett., 2022, 22, 7112–7118 CrossRef CAS PubMed.
- A. Z. Afshord, B. E. Uzuner, W. Soltanpoor, S. H. Sedani, T. Aernouts, G. Gunbas, Y. Kuang and S. Yerci, Efficient and Stable Inverted Wide-Bandgap Perovskite Solar Cells and Modules Enabled by Hybrid Evaporation-Solution Method, Adv. Funct. Mater., 2023, 33, 2301695 CrossRef CAS.
- W. Chen, S. Zhou, J. Cao, L. Yuan and W. Liu, Controlled Crystallization and Enhanced Performance of γ-CsPbI 3 Perovskite Through Methylammonium Iodide-Assisted Coevaporation, Small Methods, 2024, 9, 2400796 CrossRef.
- N. Leupold and F. Panzer, Recent Advances and Perspectives on Powder-Based Halide Perovskite Film Processing, Adv. Funct. Mater., 2021, 31, 2007350 CrossRef CAS.
- E. Couderc, Perovskite photovoltaics: Manufacturing costs, Nat. Energy, 2017, 2, 17080 CrossRef.
- Y. Liu, Z. Zhang, T. Wu, W. Xiang, Z. Qin, X. Shen, Y. Peng, W. Shen, Y. Li and L. Han, Cost Effectivities Analysis of Perovskite Solar Cells: Will it Outperform Crystalline Silicon Ones?, Nano-Micro Lett., 2025, 17, 219 CrossRef CAS PubMed.
- J. A. Dawson, A. J. Naylor, C. Eames, M. Roberts, W. Zhang, H. J. Snaith, P. G. Bruce and M. S. Islam, Mechanisms of Lithium Intercalation and Conversion Processes in Organic–Inorganic Halide Perovskites, ACS Energy Lett., 2017, 2, 1818–1824 CrossRef CAS.
- N. Vicente and G. Garcia-Belmonte, Methylammonium Lead Bromide Perovskite Battery Anodes Reversibly Host High Li-Ion Concentrations, J. Phys. Chem. Lett., 2017, 8, 1371–1374 CrossRef CAS PubMed.
- N. Vicente, D. Bresser, S. Passerini and G. Garcia-Belmonte, Probing the 3-step Lithium Storage Mechanism in CH 3 NH 3 PbBr 3 Perovskite Electrode by Operando -XRD Analysis, ChemElectroChem, 2019, 6, 456–460 CrossRef CAS.
- D. Ramirez, Y. Suto, N. C. Rosero-Navarro, A. Miura, K. Tadanaga and F. Jaramillo, Structural and Electrochemical Evaluation of Three- and Two-Dimensional Organohalide Perovskites and Their Influence on the Reversibility of Lithium Intercalation, Inorg. Chem., 2018, 57, 4181–4188 CrossRef CAS PubMed.
- L. T. López, D. Ramírez, F. Jaramillo and J. A. Calderón, Novel hybrid organic-inorganic CH3NH3NiCl3 active material for high-capacity and sustainable lithium-ion batteries, Electrochim. Acta, 2020, 357, 136882 CrossRef.
- D. Maity, B. Kaur, P. Ghosal and M. Deepa, Large Alkylammonium Cation Based 2D-3D Hybrid Perovskite with Fast Charge Conduction for a Li-Ion Battery Anode, ACS Appl. Energy Mater., 2025, 8, 76–86 CrossRef CAS.
- M. Tathavadekar, S. Krishnamurthy, A. Banerjee, S. Nagane, Y. Gawli, A. Suryawanshi, S. Bhat, D. Puthusseri, A. D. Mohite and S. Ogale, Low-dimensional hybrid perovskites as high performance anodes for alkali-ion batteries, J. Mater. Chem. A, 2017, 5, 18634–18642 RSC.
- K. Roy, T. Li, S. Ogale and N. Robertson, Hybrid perovskite-like iodobismuthates as low-cost and stable anode materials for lithium-ion battery applications, J. Mater. Chem. A, 2021, 9, 2689–2693 RSC.
- F. Moradi and A. Iagaru, Dual-tracer imaging of malignant bone involvement using PET, Clin. Transl. Imaging, 2015, 3, 123–131 CrossRef.
- Q. Wang, T. Yang, H. Wang, J. Zhang, X. Guo, Z. Yang, S. Lu and W. Qin, Morphological and chemical tuning of lead halide perovskite mesocrystals as long-life anode materials in lithium-ion batteries, CrystEngComm, 2019, 21, 1048–1059 RSC.
- Q. Jiang, M. Chen, J. Li, M. Wang, X. Zeng, T. Besara, J. Lu, Y. Xin, X. Shan, B. Pan, C. Wang, S. Lin, T. Siegrist, Q. Xiao and Z. Yu, Electrochemical Doping of Halide Perovskites with Ion Intercalation, ACS Nano, 2017, 11, 1073–1079 CrossRef CAS PubMed.
- S. Liu, K. Zhang, L. Tan, S. Qi, G. Liu, J. Chen and Y. Lou, All-inorganic halide perovskite CsPbBr3@CNTs composite enabling superior lithium storage performance with pseudocapacitive contribution, Electrochim. Acta, 2021, 367, 137352 CrossRef CAS.
- W. Xie, J. Cao, P. Li, M. Fan, S. Xu, J. Du and J. Zhang, Stabilized anode-electrolyte interfaces via Cs4Pb(Cl/Br/I)6 perovskite crystal based glass-ceramics for fast and long cycle-life lithium ion batteries, Mater. Des., 2022, 220, 110860 CrossRef CAS.
- T. Paul, S. Maiti, B. K. Chatterjee, P. Bairi, B. K. Das, S. Thakur and K. K. Chattopadhyay, Electrochemical Performance of 3D Network CsPbBr3 Perovskite Anodes for Li-Ion Batteries: Experimental Venture with Theoretical Expedition, J. Phys. Chem. C, 2021, 125, 16892–16902 CrossRef CAS.
- A. Mathieson, M. Rahil, Y. Zhang, W. M. Dose, J. T. Lee, F. Deschler, S. Ahmad and M. De Volder, Ruddlesden Popper 2D perovskites as Li-ion battery electrodes, Mater. Adv., 2021, 2, 3370–3377 RSC.
- L. T. C. Lopez, J. C. Alvarez Quiceno, F. Jaramillo, J. M. Osorio-Guillén and J. A. Calderón, Enhancement of electrochemical stability by molecular cation vacancies of the electrode material CH3NH3NiCl3 for lithium-ion batteries, J. Electroanal. Chem., 2024, 968, 118505 CrossRef.
- W. Zhao, C. Liu and X. Yin, Cs4PbBr6 Combined with Graphite as Anode for High-Performance Lithium Batteries, Metals, 2022, 12, 1584 CrossRef CAS.
- J. Cao, W. Xie, K. Liu, X. Meng, J. Zhang, J. Zhang, Z. Yao and M. Fan, SnF2 -Doped Cs4PbBr6 Glass Ceramic as a High-Performance Anode for Li-Ion Batteries, J. Phys. Chem. C, 2022, 126, 3359–3365 CrossRef CAS.
- X. Liu, R. Wei, S. Fu, W. Xie, S. Xu and J. Zhang, Cs4PbBr6 QDs silicate glass-ceramic: A potential anode material for LIBs, Ceram. Int., 2022, 48, 23293–23299 CrossRef CAS.
- X.-H. Wu, Y. Chai, J. Shen, P.-W. Huang, S.-Y. Xu, H.-Y. Zhong, B.-C. Chen, Y. Zhao, B. Sa and K.-Z. Du, Deciphering the role of van der Waals heterostructures in enhancing layered perovskite anodes for high-performance lithium-ion batteries, J. Mater. Chem. A, 2025, 13, 8750–8760 RSC.
- H. Kong, J. Wu, Y. Han, Y. Zhang, N. Zhou, Q. Chen, W. Sun, H. Zhou and L.-M. Peng, One-dimensional perovskite-based Li-ion battery anodes with high capacity and cycling stability, J. Energy Chem., 2022, 72, 73–80 CrossRef CAS.
- W. Jia, J. Bao, H. Zhang, M. Wu, J. Qiu, H. Wu and Y. Zhang, Effect of Mn-ion Reconstructed Lattice on Lead-Free Halide Perovskite Cs3Bi2-XMnxCl9 Anode in Li-ion Batteries, J. Environ. Sci., 2025, 159, 154–165 CrossRef PubMed.
- A. Kostopoulou, D. Vernardou, K. Savva and E. Stratakis, All-inorganic lead halide perovskite nanohexagons for high performance air-stable lithium batteries, Nanoscale, 2019, 11, 882–889 RSC.
- H. Wu, J. Pi, Q. Liu, Q. Liang, J. Qiu, J. Guo, Z. Long, D. Zhou and Q. Wang, All-Inorganic Lead Free Double Perovskite Li-Battery Anode Material Hosting High Li+ Ion Concentrations, J. Phys. Chem. Lett., 2021, 12, 4125–4129 CrossRef CAS PubMed.
- J. Bao, W. Jia, H. Zhang, M. Wu, J. Qiu, H. Wu and Y. Zhang, Synergistic Enhancement of Mn2+ Doping and Thermal Field Treatment on Cs2NaBiCl6 Double Perovskite Anode Performance for Lithium-Ion Batteries, J. Phys. Chem. C, 2025, 129, 8521–8528 CrossRef CAS.
- Z. Shi, J. Guo, Y. Chen, Q. Li, Y. Pan, H. Zhang, Y. Xia and W. Huang, Lead-Free Organic–Inorganic Hybrid Perovskites for Photovoltaic Applications: Recent Advances and Perspectives, Adv. Mater., 2017, 186, 113649 Search PubMed.
- S. F. Hoefler, G. Trimmel and T. Rath, Progress on lead-free metal halide perovskites for photovoltaic applications: a review, Monatsh. Chem., 2017, 148, 795–826 CrossRef CAS PubMed.
- P. Pandey, N. Sharma, R. A. Panchal, S. W. Gosavi and S. Ogale, Realization of High Capacity and Cycling Stability in Pb-Free A2CuBr4 (A=CH3NH3/Cs, 2D/3D) Perovskite-Based Li-Ion Battery Anodes, ChemSusChem, 2019, 12, 3742–3746 CrossRef CAS PubMed.
- X.-Q. Zhang, X.-B. Cheng, X. Chen, C. Yan and Q. Zhang, Fluoroethylene Carbonate Additives to Render Uniform Li Deposits in Lithium Metal Batteries, Adv. Funct. Mater., 2017, 27, 1605989 CrossRef.
- X.-B. Cheng, T.-Z. Hou, R. Zhang, H.-J. Peng, C.-Z. Zhao, J.-Q. Huang and Q. Zhang, Dendrite-Free Lithium Deposition Induced by Uniformly Distributed Lithium Ions for Efficient Lithium Metal Batteries, Adv. Mater., 2016, 28, 2888–2895 CrossRef CAS PubMed.
- D. Rehnlund, F. Lindgren, S. Böhme, T. Nordh, Y. Zou, J. Pettersson, U. Bexell, M. Boman, K. Edström and L. Nyholm, Lithium trapping in alloy forming electrodes and current collectors for lithium based batteries, Energy Environ. Sci., 2017, 10, 1350–1357 RSC.
- S. Rajendran, N. K. Thangavel, K. Mahankali and L. M. R. Arava, Toward Moisture-Stable and Dendrite-Free Garnet-Type Solid-State Electrolytes, ACS Appl. Energy Mater., 2020, 3, 6775–6784 CrossRef CAS.
- X.-B. Cheng, C. Yan, X. Chen, C. Guan, J.-Q. Huang, H.-J. Peng, R. Zhang, S.-T. Yang and Q. Zhang, Implantable Solid Electrolyte Interphase in
Lithium-Metal Batteries, Chem, 2017, 2, 258–270 CAS.
- F. Ding, W. Xu, G. L. Graff, J. Zhang, M. L. Sushko, X. Chen, Y. Shao, M. H. Engelhard, Z. Nie, J. Xiao, X. Liu, P. V. Sushko, J. Liu and J.-G. Zhang, Dendrite-Free Lithium Deposition via Self-Healing Electrostatic Shield Mechanism, J. Am. Chem. Soc., 2013, 135, 4450–4456 CrossRef CAS PubMed.
- N. Liu, J. Baek, S. M. Kim, S. Hong, Y. K. Hong, Y. S. Kim, H.-S. Kim, S. Kim and J. Park, Improving the Stability of High-Performance Multilayer MoS2 Field-Effect Transistors, ACS Appl. Mater. Interfaces, 2017, 9, 42943–42950 CrossRef CAS PubMed.
- Q. Wu, G. Wang, Q. Huang, S. Xie, J. Huang and X. Wang, Perovskite-type La0.6Sr0.4Co0.2Fe0.8O3−δ as an artificial interphase layer for dendrite-free Li metal anodes, Chem. Eng. J., 2022, 444, 136340 CrossRef CAS.
- Z. Yu, Y. Cui and Z. Bao, Design Principles of Artificial Solid Electrolyte Interphases for Lithium-Metal Anodes, Cell Rep. Phys. Sci., 2020, 1, 100119 CrossRef.
- R. Xu, X.-B. Cheng, C. Yan, X.-Q. Zhang, Y. Xiao, C.-Z. Zhao, J.-Q. Huang and Q. Zhang, Artificial Interphases for Highly Stable Lithium Metal Anode, Matter, 2019, 1, 317–344 CrossRef.
- X.-B. Cheng, R. Zhang, C.-Z. Zhao and Q. Zhang, Toward Safe Lithium Metal Anode in Rechargeable Batteries: A Review, Chem. Rev., 2017, 117, 10403–10473 CrossRef CAS PubMed.
- Y. Zhang, Y. Liu, J. Zhou, D. Wang, L. Tan and C. Yi, 3D cubic framework of fluoride perovskite SEI inducing uniform lithium deposition for air-stable and dendrite-free lithium metal anodes, Chem. Eng. J., 2022, 431, 134266 CrossRef CAS.
- N. Kaisar, A. Singh, P.-Y. Yang, Y.-T. Chen, S. Li, C.-W. Pao, S. Jou and C.-W. Chu, Long-lifespan lithium–metal batteries obtained using a perovskite intercalation layer to stabilize the lithium electrode, J. Mater. Chem. A, 2020, 8, 9137–9145 RSC.
- S. Bai, P. Da, C. Li, Z. Wang, Z. Yuan, F. Fu, M. Kawecki, X. Liu, N. Sakai, J. T.-W. Wang, S. Huettner, S. Buecheler, M. Fahlman, F. Gao and H. J. Snaith, Planar perovskite solar cells with long-term stability using ionic liquid additives, Nature, 2019, 571, 245–250 CrossRef CAS PubMed.
- M. Kim, G.-H. Kim, T. K. Lee, I. W. Choi, H. W. Choi, Y. Jo, Y. J. Yoon, J. W. Kim, J. Lee, D. Huh, H. Lee, S. K. Kwak, J. Y. Kim and D. S. Kim, Methylammonium Chloride Induces Intermediate Phase Stabilization for Efficient Perovskite Solar Cells, Joule, 2019, 3, 2179–2192 CrossRef CAS.
- Y. Wang, M. I. Dar, L. K. Ono, T. Zhang, M. Kan, Y. Li, L. Zhang, X. Wang, Y. Yang, X. Gao, Y. Qi, M. Grätzel and Y. Zhao, Thermodynamically stabilized β-CsPbI3–based perovskite solar cells with efficiencies >18%, Science, 2019, 365, 591–595 CrossRef CAS PubMed.
- M. Yang, D. Wang, Y. Ling, X. Guo and W. Chen, Emerging Advanced Photo-Rechargeable Batteries, Adv. Funct. Mater., 2024, 34, 2410398 CrossRef CAS.
- L. Song, Y. Fan, H. Fan, X. Yang, K. Yan, X. Wang and L. Ma, Photo-assisted rechargeable metal batteries, Nano Energy, 2024, 125, 109538 CrossRef CAS.
- S. Ahmad, C. George, D. J. Beesley, J. J. Baumberg and M. De Volder, Photo-Rechargeable Organo-Halide Perovskite Batteries, Nano Lett., 2018, 18, 1856–1862 CrossRef CAS PubMed.
- M. He, L. Zhang and J. Li, Theoretical investigation on interactions between lithium ions and two-dimensional halide perovskite for solar-rechargeable batteries, Appl. Surf. Sci., 2021, 541, 148509 CrossRef CAS.
- N. Tewari, S. B. Shivarudraiah and J. E. Halpert, Photorechargeable Lead-Free Perovskite Lithium-Ion Batteries Using Hexagonal Cs3Bi2I9 Nanosheets, Nano Lett., 2021, 21, 5578–5585 CrossRef CAS PubMed.
- B. D. Boruah, B. Wen and M. De Volder, Light Rechargeable Lithium-Ion Batteries Using V 2 O 5 Cathodes, Nano Lett., 2021, 21, 3527–3532 CrossRef CAS PubMed.
- A. Paolella, C. Faure, G. Bertoni, S. Marras, A. Guerfi, A. Darwiche, P. Hovington, B. Commarieu, Z. Wang, M. Prato, M. Colombo, S. Monaco, W. Zhu, Z. Feng, A. Vijh, C. George, G. P. Demopoulos, M. Armand and K. Zaghib, Light-assisted delithiation of lithium iron phosphate nanocrystals towards photo-rechargeable lithium ion batteries, Nat. Commun., 2017, 8, 14643 CrossRef CAS PubMed.
- X. Yin, G. Di, Y. Liu, G. Wang, C. Mi, Y. Kuang, X. Xiang, X. Sun, E. Edri, X. Lv and M. Li, An energy-saving photo-rechargeable lithium-ion battery based on lead-free hybrid perovskite, Sci. China Mater., 2025, 68, 1091–1099 CrossRef.
- L. Kin, Z. Liu, O. Astakhov, S. N. Agbo, H. Tempel, S. Yu, H. Kungl, R.-A. Eichel, U. Rau, T. Kirchartz and T. Merdzhanova, Efficient Area Matched Converter Aided Solar Charging of Lithium Ion Batteries Using High Voltage Perovskite Solar Cells, ACS Appl. Energy Mater., 2020, 3, 431–439 CrossRef CAS.
- A. Gurung, K. Chen, R. Khan, S. S. Abdulkarim, G. Varnekar, R. Pathak, R. Naderi and Q. Qiao, Highly Efficient Perovskite Solar Cell Photocharging of Lithium Ion Battery Using DC–DC Booster, Adv. Energy Mater., 2017, 7, 1602105 CrossRef.
- Y. Hu, Y. Bai, B. Luo, S. Wang, H. Hu, P. Chen, M. Lyu, J. Shapter, A. Rowan and L. Wang, A Portable and Efficient Solar-Rechargeable Battery with Ultrafast Photo-Charge/Discharge Rate, Adv. Energy Mater., 2019, 9, 1900872 CrossRef.
- X. Xia, Z. Ku, D. Zhou, Y. Zhong, Y. Zhang, Y. Wang, M. J. Huang, J. Tu and H. J. Fan, Perovskite solar cell powered electrochromic batteries for smart windows, Mater. Horiz., 2016, 3, 588–595 RSC.
- P. Chen, G. Li, T. Li and X. Gao, Solar-Driven Rechargeable Lithium–Sulfur Battery, Adv. Sci., 2019, 6, 1900620 CrossRef PubMed.
- T.-T. Li, Y.-B. Yang, B.-S. Zhao, Y. Wu, X.-W. Wu, P. Chen and X.-P. Gao, Photo-rechargeable all-solid-state lithium − sulfur batteries based on perovskite indoor photovoltaic modules, Chem. Eng. J., 2023, 455, 140684 CrossRef CAS.
- R. M. Ansari, S. Chamola and S. Ahmad, Ruddlesden–Popper 2D Perovskite-MoS 2 Hybrid Heterojunction
Photocathodes for Efficient and Scalable Photo-Rechargeable Li-Ion Batteries, Small, 2024, 20, 2401350 CrossRef CAS PubMed.
- H. Liu, P. Wu, R. Wang, H. Meng, Y. Zhang, W. Bao and J. Li, A Photo-rechargeable Aqueous Zinc–Tellurium Battery Enabled by the Janus-Jointed Perovskite/Te Photocathode, ACS Nano, 2023, 17, 1560–1569 CrossRef CAS PubMed.
- H. Hassan, A. Althobaiti, A. Mohammad, Z. Ahmad, I. Barsoum, M. Sohail, S. Mumtaz and A. A. Rafi, A hybrid MAPbI3/PEDOT-ZrO2 perovskedot composite for enhanced stability and charge transport in photo-batteries, Inorg. Chem. Commun., 2024, 170, 113380 CrossRef CAS.
- Y. Qian, Q. Ruan, M. Xue and L. Chen, Emerging perovskite materials for supercapacitors: Structure, synthesis, modification, advanced characterization, theoretical calculation and electrochemical performance, J. Energy Chem., 2024, 89, 41–70 CrossRef CAS.
- P. Simon and Y. Gogotsi, Materials for electrochemical capacitors, Nat. Mater., 2008, 7, 845–854 CrossRef CAS PubMed.
- M. Mohan, N. P. Shetti and T. M. Aminabhavi, Perovskites: A new generation electrode materials for storage applications, J. Power Sources, 2023, 574, 233166 CrossRef CAS.
- R. J. Kashtiban, C. E. Patrick, Q. Ramasse, R. I. Walton and J. Sloan, Picoperovskites: The Smallest Conceivable Isolated Halide Perovskite Structures Formed within Carbon Nanotubes, Adv. Mater., 2022, 35, 2208575 CrossRef PubMed.
- A. Kostopoulou, D. Vernardou, N. Livakas, K. Brintakis, S. Daskalakis and E. Stratakis, Harnessing laser technology to create stable metal halide perovskite–rGO conjugates as promising electrodes for Zn-ion capacitors, Nanoscale, 2024, 16, 6455–6463 RSC.
- A. Vilanova, P. Dias, T. Lopes and A. Mendes, The route for commercial photoelectrochemical water splitting: a review of large-area devices and key upscaling challenges, Chem. Soc. Rev., 2024, 53, 2388–2434 RSC.
- M. A. Kuzina, D. D. Kartsev, A. V. Stratonovich and P. A. Levkin, Organogels versus Hydrogels: Advantages, Challenges, and Applications, Adv. Funct. Mater., 2023, 33, 2301421 CrossRef CAS.
- M. Titirici, P. Johansson, M. Crespo Ribadeneyra, H. Au, A. Innocenti, S. Passerini, E. Petavratzi, P. Lusty, A. A. Tidblad, A. J. Naylor, R. Younesi, Y. A. Chart, J. Aspinall, M. Pasta, J. Orive, L. M. Babulal, M. Reynaud, K. G. Latham, T. Hosaka, S. Komaba, J. Bitenc, A. Ponrouch, H. Zhang, M. Armand, R. Kerr, P. C. Howlett, M. Forsyth, J. Brown, A. Grimaud, M. Vilkman, K. B. Dermenci, S. Mousavihashemi, M. Berecibar, J. E. Marshall, C. R. McElroy, E. Kendrick, T. Safdar, C. Huang, F. M. Zanotto, J. F. Troncoso, D. Z. Dominguez, M. Alabdali, U. Vijay, A. A. Franco, S. Pazhaniswamy, P. S. Grant, S. López Guzman, M. Fehse, M. Galceran and N. Antuñano, 2024 roadmap for sustainable batteries, J. Phys. Energy, 2024, 6, 041502 CrossRef CAS.
- A. S. R. Bati, Y. L. Zhong, P. L. Burn, M. K. Nazeeruddin, P. E. Shaw and M. Batmunkh, Next-generation applications for integrated perovskite solar cells, Commun. Mater., 2023, 4, 2 CrossRef CAS.
- E. Davari and D. G. Ivey, Bifunctional electrocatalysts for Zn–air batteries, Sustainable Energy Fuels, 2018, 2, 39–67 RSC.
- R. Liu, C. Liu and S. Fan, A photocapacitor based on organometal halide perovskite and PANI/CNT composites integrated using a CNT bridge, J. Mater. Chem. A, 2017, 5, 23078–23084 RSC.
- D. Meng, H. Gu, Q. Lu, Y. Zhao, G. Zhu, Y. Zhang, Q. Zhong and Y. Bu, Advances and Perspectives for the Application of Perovskite Oxides in Supercapacitors, Energy Fuels, 2021, 35, 17353–17371 CrossRef CAS.
- P. Andričević, X. Mettan, M. Kollár, B. Náfrádi, A. Sienkiewicz, T. Garma, L. Rossi, L. Forró and E. Horváth, Light-Emitting Electrochemical Cells of Single Crystal Hybrid Halide Perovskite with Vertically Aligned Carbon Nanotubes Contacts, ACS Photonics, 2019, 6, 967–975 CrossRef.
- A. Mahapatra, M. Mandal, A. Das Mahapatra, V. Anilkumar, J. Nawrocki, R. D. Chavan, P. Yadav and D. Prochowicz, Mechanochemically-assisted synthesis of 3D, 2D and quasi 2D lead halide perovskites for supercapacitor applications, Mater. Adv., 2024, 5, 3881–3889 RSC.
- C. H. Ng, H. N. Lim, S. Hayase, Z. Zainal, S. Shafie, H. W. Lee and N. M. Huang, Cesium Lead Halide Inorganic-Based Perovskite-Sensitized Solar Cell for Photo-Supercapacitor Application under High Humidity Condition, ACS Appl. Energy Mater., 2018, 1, 692–699 CrossRef CAS.
- A. Slonopas, H. Ryan and P. Norris, Ultrahigh energy density CH3NH3PbI3 perovskite based supercapacitor with fast discharge, Electrochim. Acta, 2019, 307, 334–340 CrossRef CAS.
- P. Maji, A. Ray, P. Sadhukhan, A. Roy and S. Das, Fabrication of symmetric supercapacitor using cesium lead iodide (CsPbI3) microwire, Mater. Lett., 2018, 227, 268–271 CrossRef CAS.
- L. E. Oloore, M. A. Gondal, I. K. Popoola and A. Popoola, Cadmium Sulfide Quantum Dots–Organometallic Halide Perovskite Bilayer Electrode Structures for Supercapacitor Applications, ChemElectroChem, 2020, 7, 486–492 CrossRef CAS.
- I. K. Popoola, M. A. Gondal, A. Popoola and L. E. Oloore, Bismuth-based organometallic-halide perovskite photo-supercapacitor utilizing novel polymer gel electrolyte for hybrid energy harvesting and storage applications, J. Energy Storage, 2022, 53, 105167 CrossRef.
- S. Güz, M. Buldu-Akturk, H. Göçmez and E. Erdem, All-in-One Electric Double Layer Supercapacitors Based on CH 3 NH 3 PbI 3 Perovskite Electrodes, ACS Omega, 2022, 7, 47306–47316 CrossRef PubMed.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.