Next Article in Journal
Correction: Shi et al. Parameter Estimation Method for Virtual Power Plant Frequency Response Model Based on SLP. Energies 2024, 17, 3124
Previous Article in Journal
Heat Transfer Enhancement in a 3D-Printed Compact Heat Exchanger
Previous Article in Special Issue
Photocatalytic Hydrogen Production Enhancement of NiTiO3 Perovskite through Cobalt Incorporation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Overview of the Recent Findings in the Perovskite-Type Structures Used for Solar Cells and Hydrogen Storage

1
Centre for Advanced Photovoltaics, Faculty of Electrical Engineering, Czech Technical University in Prague, Technická 1902/2, 166 27 Prague, Czech Republic
2
FZU—Institute of Physics of the Czech Academy of Sciences, Na Slovance 1999/2, 182 00 Prague, Czech Republic
*
Author to whom correspondence should be addressed.
I. Stachiv is currently Seconded at the European Research Council (ERC) Executive Agency of the European Commission. The views expressed in this article are purely those of the authors. They do not necessarily reflect the views or official positions of the European Commission and the ERC Executive Agency.
Energies 2024, 17(18), 4755; https://doi.org/10.3390/en17184755
Submission received: 17 July 2024 / Revised: 29 August 2024 / Accepted: 31 August 2024 / Published: 23 September 2024
(This article belongs to the Special Issue Advanced Materials and Technologies for Hydrogen Evolution)

Abstract

:
Perovskite-type structures have unique crystal architecture and chemical composition, which make them highly attractive for the design of solar cells. For instance, perovskite-based solar cells have been shown to perform better than silicon cells, capable of adsorbing a wide range of light wavelengths, and they can be relatively easily manufactured at a low cost. Importantly, the perovskite-based structures can also adsorb a significant amount of hydrogen atoms into their own structure; therefore, perovskite holds promise in the solid-state storage of hydrogen. It is widely expected by the scientific community that the controlled adsorption/desorption of the hydrogen atoms into/from perovskite-based structures can help to overcome the main hydrogen storage issues such as a low volumetric density and the safety concerns (i.e., the hydrogen embrittlement affects strongly the mechanical properties of metals and, as such, the storage or transport of the gaseous hydrogen in the vessels is, especially for large vessel volumes, challenging). The purpose of this review is to provide an updated overview of the recent results and studies focusing on the perovskite materials used for both solar cells and hydrogen storage applications. Particular attention is given to (i) the preparation and the achievable efficiency and stability of the perovskite solar cells and (ii) the structural, thermodynamic, and storage properties of perovskite hydrides and oxides. We show that the perovskite materials can not only reach the efficiency above current Si-based solar cells but also, due to good stability and reasonable price, can be preferable in the solid-state storage of hydrogen. Then, the future trends and directions in the research and application of perovskite in both solar cells and hydrogen storage are also highlighted.

1. Introduction

The combination of the enormous demand for energy caused by rapid global economic development, population growth, and recent environmental challenges such as pollution or global warming yields the pursuit of suitable renewable energy sources such as hydropower, wind, geothermal, tidal, solar thermal, photovoltaics and/or hydrogen [1,2,3,4,5]. Hydropower, which generates electricity from fast-running or falling water, can be used for both electricity production and energy storage. However, despite the apparent advantages of hydrogen power over other renewable energy sources (e.g., relatively large and stable source of energy), it has many serious limitations, such as a high construction cost (i.e., it can be built on only specific land areas) and a large impact on the local ecosystem [6]. High construction and/or maintenance costs are also one of the main drawbacks of tidal or geothermal energy sources [7,8]. Then, it is particularly the relatively low installation and maintenance costs that make the photovoltaic systems a highly attractive source of renewable energy [9]. In general, photovoltaic systems utilize solar modules; each consists of a large number of solar cells [10,11]. Interestingly, the first solar cells were silicon-based and were developed at Bell Laboratories in 1954. They utilized the photoelectric effect in the p-n junction of the doped silicon and could reach an efficiency of ~6% [12]. We remind the reader that silicon is the second most common element in the Earth‘s lithosphere. It is also non-toxic and has exceptional stability and optimal bandgap for photovoltaic conversion. In addition, the current semiconductor technology is silicon-based. Therefore, it enables the easy preparation of high-quality silicon wafers at a reasonable cost. As a result, silicon has become fundamental to the development of photovoltaic technology [13,14,15]. Note that in the past 50 years, the efficiency of silicon solar cells has improved more than three times [16,17]. Nowadays, silicon-based technology is approaching its own intrinsic efficiency limits of ~29.1% [17]. For example, by combining the Si heterojunction architecture (i.e., amorphous Si/crystalline Si) with integrated back contacts, the superior photoconversion efficiency of ~26.7% for silicon solar cells has been reported [18]. On the other hand, silicon-based solar cells have several limitations, such as the high environmental impact (i.e., their fabrication requires a significant amount of energy, as well as their recycling is complicated) and rigidity (i.e., no possible bending of cells yielding a limited shape design), the production cost cannot be significantly reduced and, finally, the efficiency which is close to their theoretical limits (i.e., efficiency above 30% is not easily being reachable) is also strongly weather condition dependent (i.e., the amount of produced electricity notably reduces for lower light exposure) [19,20,21].
Hence, to overcome these limitations, next-generation materials such as quantum dots [22], organic [23], dye-sensitized [24], copper zinc tin sulfide [25], and perovskite [26], often referred to as the “third generation” of solar cells, have been proposed and extensively studied. Among them, perovskite solar cells (PSCs) have attracted significant attention from the community due to their excellent optoelectronic properties such as high absorption coefficients [27,28], long charge-carrier diffusion lengths [29,30], low exciton binding energies [31,32], tuneable bandgaps [33], and power conversion efficiency (PCE) [34,35] but also a low manufacturing cost (i.e., can be prepared by roll-to-roll technology), lightweight and flexibility [36,37,38]. In general, the PSCs can be separated based on the used materials (compounds) into two categories: all-inorganic perovskites [39,40] and hybrid organic-inorganic perovskites [41,42]. The standard formulation of these perovskite compounds is ABX3, where both A and B represent cations, and X denotes an anion (see Figure 1) [43,44,45]. Note that a variety of perovskite compounds are available in different forms, including carbides, nitrides, fluorides, chlorides, bromides, oxides, hydrides, and iodides [46,47]. When hydrogen, oxygen, halide atom, nitrogen, or fluorine replaces element X within a compound, the resulting entities are known as perovskite hydride, perovskite oxide, perovskite halide, perovskite nitride, and perovskite fluoride, respectively [48,49,50,51,52].
In 2009, the conversion efficiency of PSCs of 3.8% was reported in the pioneer study of Kojima et al. [53]. Since then, the efficiency of PSCs has increased several times. Recently, Yoo et al. [54] have shown that by utilizing the fluorine-doped tin dioxide with improved charge carrier management, an efficiency of 25.2% (i.e., efficiency is above other thin-film solar cells [55]) has been demonstrated. This study presented charge carrier management, and the optical advances will soon enable it to reach efficiency above 26% [56], that is, more than for Si-based solar cells. Another important use of perovskite is in the area of flexible solar cells [19]. We remind the reader that it is, particularly, a combination of lightweight and flexibility that makes flexible solar cells highly attractive for a large number of applications, including portable electrical chargers, electrical production on large-scale industrial roofs, powering various vehicles, bendable/wearable electronics, etc. In the flexible perovskite solar cells (F-PSCs), the main difficulties are related to the fabrication (i.e., the formation of the electron transport layer (ETL) is at a low temperature, below 150 °C, highly challenging) [57]. Hence, different fabrication approaches and strategies have been developed to design durable F-PSCs with high efficiency [58,59,60]. For example, the TiO2 ETL prepared by the plasma-enhanced atomic layer deposition has enabled it to achieve good durability and a conversion efficiency of 12.2% [58]. More recently, Chung et al. [59] developed a low-temperature fabrication process of the porous planar ETL, which, in turn, enabled it to reach a superior certified efficiency of F-PSCs of 19.9%. Another approach to achieving high efficiency of F-PSCs is utilizing hybrid electrodes. For example, Paik et al. [60] have demonstrated an efficiency of 21.02% when using the SnO2-TiO2 hybrid electrodes.
In addition, perovskite materials have also been extensively studied due to their excellent hydrogen storage capacity (i.e., solid-state storage) [61,62,63]. It is noteworthy that hydrogen is the most common element in the universe with a large energy density, non(low)-toxic, carbon-free, and can also be used with some modifications within the current combustion engines [64,65,66,67,68]. In contrast to batteries, hydrogen does not have deep discharge and other negative effects, and, as such, the storage capacity does not degrade in time [69]. The storage and transportation of hydrogen remain challenging, and researchers have been diligently working on the development of novel materials for hydrogen storage [70,71]. It can be stored in either a liquid state through liquefaction or a gaseous state via gas compression or a solid state [72,73,74,75]. The gravimetric density and hydrogen desorption temperature are two important performance indicators for evaluating hydrogen storage materials [76,77]. Higher gravimetric density means more hydrogen can be stored in a smaller volume or mass of the storage material [78,79]. Hydrogen desorption temperature refers to the temperature at which hydrogen molecules are released from a solid storage material [80]. To achieve a controlled release of hydrogen, maintaining a moderate hydrogen desorption temperature is crucial. Hence, the advancement of hydrogen storage materials possessing high gravimetric density along with moderate hydrogen desorption temperatures is the key to enhancing the storage and utilization of hydrogen energy [77,81,82]. Among various perovskite compounds, the oxides and hydrides are the predominant structural configurations with potential applicability for hydrogen storage (e.g., they have good stability and can be prepared at a relatively reasonable cost) [83,84,85,86,87].
It is evident from the present discussion that the field of perovskite materials is growing exponentially in terms of both the published papers and research groups being involved. The purpose of this review is to provide an overview of the recent findings and trends in perovskite materials used in both emergent areas: photovoltaic and solid-state hydrogen storage. The review is organized as follows: Section 2 focuses on the perovskite materials for solar cell applications. The solid-state storage of hydrogen by utilizing perovskite is given in Section 3. The fundamental properties of perovskite that are of importance in hydrogen storage are also discussed in Section 4 and Section 5. Finally, a brief discussion on possible future trends in the field of perovskite is given in Section 6. Importantly, to our knowledge, this is the first review that combines both areas of research in perovskite materials. As a result, our review would help researchers keep tracking the current trends and findings in the perovskite materials and, correspondingly, can help researchers share their knowledge in two separate studies of perovskite.

2. Perovskite Materials Used in Photovoltaic

2.1. Perovskite Solar Cells

As mentioned previously in the Introduction, a large number of perovskite compounds have already been prepared. Then, it is the outstanding optical properties compared to all-inorganic perovskites that made hybrid organic-inorganic being predominantly used in the PSCs [88,89]. Recent research has shown that metal halide perovskites are highly potential compounds of hybrid organic-inorganic perovskites [90,91]. It is because of a combination of the energy tunable bandgap [92,93] and a highly flexible crystal lattice, which allows the formation of dimensionalities ranging from three-dimensional down to zero-dimensional [94,95]. Noticing that in these compounds, cation A stands for cesium, methylammonium (MA), and/or formamidinium (FA), whereas B is a divalent metal cation such as lead or tin; and X is a halide such as iodide, bromide and/or chloride as also depicted in Figure 1 [96,97].
Importantly, all of these perovskite materials can be prepared by either wet or dry synthesis [98,99]. Because of its simplicity, high repeatability, and precise control over stoichiometry, wet synthesis is more often used in PSCs [100,101]. However, there are several issues associated with wet synthesis, such as solubility, toxicity, solvent compatibility, and higher cost [50,102]. In contrast, dry synthesis enables the insolubility problem to be overcome through solid-state reactions, mechanochemical synthesis, and vapor phase deposition methods [103,104]. The PSCs generally consist of five fundamental layers: the perovskite light-absorber layer, the transparent conductive oxide as a conducting substrate (typically tin-doped indium oxide or fluorine-doped tin oxide), the metal (mainly gold, silver, or copper), and two charge transport layers, typically the ETL and the hole transport layer [105]. PSCs have a similar operating principle as conventional heterojunction or p-i-n solar cells. The light absorbed by the perovskite absorber layer produces free electron-hole pairs (excitons are weakly bound), which are subsequently transported to the charge transport layer interfaces, producing an open-circuit voltage and a photocurrent [106,107,108]. Restricting the analysis to the function of the hole transport material in optimizing PCE, its highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) energy levels must be strategically aligned with those of the perovskite and the back electrode within the device. With vacuum level as the reference, the HOMO level of hole transport materials should be higher in energy than the perovskite valence band in order to facilitate hole transportation from the perovskite. Secondly, the LUMO of the hole transport materials should be as high as possible with respect to the perovskite conduction band, as shown in Figure 2 [109]. The latter conduction band offset would avoid electrons traveling from the perovskite to the hole transport materials and, as a result, reduce the current and voltage due to the recombination with holes in the hole-transporting layer or at the interfaces. Furthermore, the HOMO level of the hole transport materials should be positioned just below the Fermi level of the back electrode to ensure rapid charge collection [106,110,111].
PSC devices can also be divided based on their planar architecture into conventional (n-i-p) PSCs and inverted (p-i-n) PSCs. (see Figure 3) [112]. Many recent studies on high-performance conventional PSCs [113] utilize zinc oxide (ZnO) [114,115], tin oxide (SnO2) [116,117,118], or spiro-OMeTAD [119,120], whereas for high-performance p-i-n devices, the Poly[bis(4-phenyl)(2,4,6-trimethylphenyl)amine] (PTAA) [119,121] or self-assembled molecules [122,123], and PC61BM [124,125] or C60 [126,127] are primarily employed. In recent years, the research has shifted from the conventional PSCs to inverted PSCs due to their superior stability and efficiency approaching to of conventional PSCs [128]. Advances in the charge transport materials, investigating the impact of interface engineering, and defect passivation strategies are crucial for both the enhancement of PCE and the stability of PSCs. For example, Tan et al. [34] have proposed a dimethylacridine-based molecular doping method that can achieve a well-matched perovskite/tin-doped indium oxide (ITO) interface with comprehensive passivation of grain boundaries. This approach has enabled the achievement of a certified PCE of 25.39% and excellent stability by retaining 96.6% of the starting efficiency after 1000 h light soaking in inverted PSCs. Tang et al. [129] have revealed that the deposition of additional indium oxide on the ITO substrate by the atomic layer deposition enhances the anchoring of the self-assembled monolayers (SAMs), making them more resistant to detachment from the ITO substrate. The SAMs, containing a trimethoxysilane group, showed great anchoring to the substrate. These inverted PSCs have achieved a certified PCE of 24.8% and retained 98.9% of their initial PCE after 1200 h of operation at 85 °C. It has been demonstrated that by using a cyanoacrylic-acid-based molecular additive for NiOx-based inverted PSCs, their efficiency and stability can be easily improved [130]. In this study, a PCE of 23.48% and maintained 95.4% of its original performance after 1960 h of continuous operation at 65 °C have been reported. Caprioglio et al. [131] have studied the impact of different charge transport layers and passivation methods on Voc and short-circuit current density (Jsc) of wide-bandgap Br-rich PSCs. They demonstrated that adjustment of the perovskite n-interface layer (PCBM) can help reduce the quasi-fermi level splitting and the Voc mismatch. As a result, it increased the Voc. Then, by altering the perovskite p-interface layer (SAMs), the field screening effects can be suppressed, and correspondingly, the charge extraction and the Jsc can be enhanced. Namely, the PSCs had Voc up to 1.29 V, a fill factor of 80%, and Jsc of 17 mA/cm2. Importantly, their thermal stability with a T80 lifetime exceeding 3500 h at 85 °C. Imran et al. [132] have shown that by combining interface engineering and additive/passivation techniques, the highly efficient and stable RbCsFAPbI3-based inverted PSCs can be prepared. Here, the optimal additive content (2% GuHCl) improved film morphology and optoelectronic properties, while NiOx/PTAA bi-interfacial engineering and PEAI passivation helped to minimize defects and enhanced charge transfer resulting in Jsc of 24.52 mA/cm2, PCE of 22.78% and retained 95% of their initial performance after 500 h. It is important to note that an alternative strategy of PSC fabrication that enables the achieve an extraordinarily high certified PCE of ~25.8% has only recently been reported by Zhou et al. [35]. This strategy utilizes a modulation of anion–π interaction with halide in the AX component. As a result, it helps to establish the so-called dual-site regulation, and correspondingly, it improves phase and component purity at the nanoscale [133].

2.2. Other Solar Cells Systems Utilizing Properties of Perovskite Materials

First of all, we remind the reader that the open-circuit voltage (Voc) of a single-junction solar cell is limited by the ratio of the bandgap of the absorber over the elementary charge. Hence, in a single junction cell, an absorber with a narrower bandgap is unable to generate a high Voc due to limitations imposed by its bandgap. Conversely, an absorber with a wider bandgap has the potential to yield a higher Voc; however, the short-circuit current is restricted because photons with energy lower than the absorber’s bandgap do not contribute to the photocurrent. The limited efficiency of these solar cells (i.e., one p-n junction) can also be overcome by combining more cells together with different bandgap energy (i.e., two/multiple p-n junctions), and they are known as tandem/triple, etc. solar cells [134,135].
These tandem solar cells usually consist of two types of solar cells: perovskite-silicon (or perovskite—CIGS) and all-perovskite solar cells. In this case, the top layer is usually made of a wide-bandgap perovskite material that absorbs high-energy photons, while the bottom layer consists of a narrow-bandgap material like silicon or copper indium gallium diselenide (CIGS) that absorbs lower-energy photons (see Figure 3) [112,136,137]. For instance, Tockhorn et al. [138] fabricated perovskite-silicon tandem solar cells with periodic nanotextures that reduced the reflection losses compared to planar tandems and increased fabrication yield from 50% to 95%. Moreover, the enhanced optoelectronic properties of the perovskite top cell resulted in a significant improvement in Voc by 15 mV. Finally, they have reported a certified power conversion efficiency of 29.8%. To improve the perovskite interface, Liu et al. [139] employed a sequential interface engineering technique that enhanced the conduction band offset and recombination rate at the perovskite/C60 interface. This process involved initially depositing ethylenediamine diodide, followed by a sequential deposition of 4-Fluoro-Phenethylammonium chloride. This approach resulted in a wide bandgap (1.67 eV) PSC with a Voc of 1.26 V, and PCE of 21.8%. Furthermore, the perovskite-silicon tandem solar cell achieved a certified PCE of 29.0%. Moreover, Zheng et al. [140] found that the solvent’s properties substantially impact the degree of moisture interference. They employed n-Butanol as part of a solvent engineering strategy, leveraging its low polarity and moderate volatilization rate to improve perovskite film uniformity. The resulting wide bandgap (1.68 eV) PSC achieved 20.79% of PCE, and the perovskite-silicon tandem solar cell achieved 25.9% of PCE for 16 cm2 by slot-die deposition. In addition, Priyanka et al. [141] demonstrated that incorporating a barium stannate (BaSnO3) charge transport layer enhanced the performance of perovskite/CIGS tandem solar cells. They reported that using a high work function metal/Cs0.15MA0.15FA0.70Pb(I0.85Br0.15)3/BaSnO3/ITO configuration as an HTL-free top perovskite solar cell resulted in an impressive PCE of 39%.
Interface engineering of perovskite layers plays a crucial role in the performance of all-perovskite tandem solar cells. Yu et al. [142] revealed that the energy levels and redox potentials of tin adducts are significantly influenced by the presence of electron-withdrawing ligands. As a result, they investigated both a narrow bandgap FA0.7MA0.25Cs0.05Sn0.5Pb0.5I3 (1.26 eV) perovskite and a wide bandgap Cs0.2FA0.8PbI1.8Br1.2 (1.77 eV) perovskite, resulting in all-perovskite tandem solar cells with a certified PCE of 26.96%. Li et al. [143] reported that the solution-processed ETL ink, which consists of hybrid fullerenes (i.e., C60, PCBM, and indene-C60 bisadduct), enhanced interface morphology, conductivity, and passivation. Correspondingly, it led to the development of an all-perovskite tandem solar cell with a narrow bandgap FA0.7MA0.3Sn0.5Pb0.5I3 (~1.25 eV) perovskite and a wide bandgap Cs0.35FA0.65PbI1.8Br1.2 (~1.80 eV) perovskite, achieving a PCE of 23.3% with an active area of 20.25 cm2.
As we mentioned in the Introduction, perovskite materials are also used in flexible, semi-transparent, or space solar cells. In the case of flexible PSCs, the perovskite is fabricated on a flexible substrate such as poly(ethylene terephthalate) (PET) using thermal radiation annealing [144]. It has shown that a high power conversion efficiency of ~22.61% and a remarkable fill factor of 83.42% can be achieved by means of dual-sided annealing [145]. Briefly, dual-side annealing allows the PET to remain undistorted during high-temperature annealing, reduces the grain boundaries, decreases lattice mismatch, and achieves cleaner vertical stacking of grains in perovskite. Defects in the perovskite films can be suppressed by utilizing the ETL in SnO2:OH ETL processed at room temperature [146]. In this study, the ETL showed a lower defect density, especially a reduced concentration of oxygen vacancies, yielding improved energy band alignment and enhanced surface wettability. Then, by using a MAPbI3 base perovskite, they achieved a PCE of 18.71 % (active area: 36.50 cm2) and retained over 83% of their initial PCE after multiple flexing test cycles of flexible PSCs. Another method to suppress defects from the interface utilized the proline hydrochloride, which is rich in –NH3+ groups and possesses a conjugated rigid structure [137]. It not only stabilized the α-phase FAPbI3 template but also protected it from degradation caused by the phase transitions. By means of this method, the flexible FAPbI3 perovskite solar cell achieved a certified 23.51% of PCE (active area of 0.0601 cm2) and retained 83% of its initial PCE undergoing a flexing test involving 6000 bending cycles at a 5 mm radius [137].
Semi-transparent solar cells are used in building windows and/or cladding, as well as vehicle integration. The transparency of perovskite solar cells relies on both the bandgap of the perovskite layer and the transmittance of the electrode. A thin metal layer frequently serves as the transparent electrode in semi-transparent solar cells. For instance, Ju et al. [147] reported a semi-transparent PSC with a 3D-structured fluorine-doped tin oxide enhanced the diffuse transmittance and shortened the carrier travel distance. They have reported that transparent PSC devices achieved a power conversion efficiency between 12.0% and 14.6% and an average visible transmittance (AVT) of 13.4–17.0%. Jin et al. [148] reported that using Spiro-OMeTAD and PBDB-T as the HTL can achieve higher performance, stable thermal stability, and moisture resistance in semi-transparent solar cells. PBDB-T improved the stability of the perovskite/HTL interface via interacting with lead ions in the perovskite through its thiophene units and carbonyl functional groups, effectively passivating interfacial defects. They obtained PCE, AVT, and light utilization efficiency of ~13.71%, 36.04%, and 4.94%, respectively. Yoo et al. [149] demonstrated that the potassium pyrophosphate interlayer between the ETL SnO2 and the perovskite layer enhanced both the performance and the stability. This improvement was attributed to the passivation of the SnO2 surface, better perovskite film morphology, and reduced non-radiative recombination at the perovskite interface. Their semi-transparent perovskite solar cells achieved a PCE of ~16%, an AVT of ~36%, and long-term stability.
Solar cells face numerous challenges in the space environment, particularly from cosmic radiation. PSCs have demonstrated impressive resilience against different types of radiation, including electrons, protons, ultraviolet rays, and gamma rays [150,151,152]. The effect of high doses of 1 MeV e-beam radiation up to an accumulated fluence to 1016 ecm−2 on perovskite thin films and solar cells have been reported by Pérez-del-Rey et al. [153]. They found that PSCs based on quartz substrates remain stable even under high doses of 1 MeV e-beam irradiation. Analysis of time-resolved microwave conductivity in both pristine and irradiated films revealed a slight decrease in the charge carrier diffusion length following irradiation.

3. Perovskite-Type Hydrogen Storage—Overview

In the case of hydrogen storage, it is necessary that materials can fulfill the following conditions: (1) the hydrogen bonds in these materials are of remarkable strength; (2) these materials encompass a substantial volume, facilitating the storage of a significant amount of hydrogen; (3) these materials exhibit catalytic properties that enhance hydrogen uptake; and, (4) also exhibit acceptable gravimetric hydrogen storage capabilities, contributing to enhanced hydrogen storage capacity [154,155]. The perovskite for hydrogen storage utilizes hydride, oxide, and halide.

3.1. Perovskite-Type Hydride

The compounds formulated as ABH3, referred to as perovskite-type hydrides, have stable structures and distinctive properties; therefore, they have gained considerable attention from the scientific community [156,157]. The perovskite hydrides are usually categorized into two groups based on the selected elements A and B in the perovskite structure. In the first group, the elements A and B could be selected from groups I and II of the periodic table [158]. Typical members of this structured group are BaLiH3, CsCaH3, KMgH3, NaMgH3, RbCaH3, or SrLiH3. Among different types of ABH3, the magnesium-based perovskite hydrides have recently received particular attention from researchers due to their high hydrogen storage capacity. For example, the superior gravimetric and volumetric hydrogen storage densities of 6% and 88 kg/m3 of NaMgH3 have already been demonstrated [74]. The second group consists of those of perovskite-type hydrides, where element A is either a monovalent alkali metal or a divalent metal, and element B is a transition metal. Some examples of the compounds of this structural group include CaCoH3, CaNiH3, LiTH3 (T: Fe, Co, Ni, Cu), SrPdH3, and SrLiH3. Another area of research involves studying the elemental contributions towards the band structure of S-block perovskites, which exhibit similarities to Pb-perovskites [159]. Building on previous efforts, Adeyinka et al. [160] utilized density functional theory calculations to determine the hydrogen storage capacities of KBeH3, KMgH3, and KCaH3, which were found to be 5.866%, 4.516% and 3.649%, respectively.

3.2. Perovskite-Type Oxides

The formula of typically perovskite-type oxides is composed of ABO3. The ABO3 perovskite-type oxides possess not only a remarkable hydrogen storage capacity but also long-term stability [161,162]. Nowadays, perovskite-type oxides have emerged as suitable materials for efficient electrocatalysis in hydrogen conversion and storage, owing to their distinctive physical and chemical attributes, that is, superior redox activity, flexible structure, remarkable ionic/electronic conductivity, and outstanding thermal and chemical stability [163,164]. Researchers are devoted to studying the performance of perovskite-type oxides for hydrogen storage at room temperature [165]. The pioneering study of Sakaguchi et al. [166] has revealed that perovskites (SrCe0.95Yb0.05O3) can store hydrogen at even room temperature in the context of Ni/MH systems. A few years later, the maximum hydrogen absorption capacity of La0.6Sr0.4Co0.2Fe0.8O3 was determined to be 1.72 wt% at 333 K and 6 mol∙L−1 KOH, as recalculated from the electrochemical capacity [167].

4. Perovskite-Type Hydrides Structural, Thermodynamic, and Hydrogen Storage Properties

The fundamental properties of perovskite-type hydrides are routinely obtained from the first-principles calculations using the density-functional theory (DFT). In this section, we will present the recent data on the properties of the currently considered perovskite-hydrides materials, that is, Mg-based perovskite-type hydrides MgXH3 (X = Co, Cu, Ni) [168], Ca-based perovskite-type hydrides CaXH3 (X = Mn, Fe, Co) [47], and metal base perovskite-hydrides AeSiH3 (Ae = K, Li, Na, Mg) [169], AeVH3 (Ae = Be, Mg, Ca, Sr) [170], XSrH3 (X = K and Rb) [73], XCuH3 (X = Co, Ni, Zn) [171], XCoH3 (X = In, Mn, Sr, Sn, Cd) [172], XPtH3 (X = Li, Na, K, Rb) [173], XAlH3 (X = Na, K) [174], XTiH3 (X = K, Rb, Cs) [77] and KXH3 (X = Be, Ca, Mg) [160].
The calculated structure of MgXH3 (X = Co, Cu, Ni) hydride perovskite is given in Figure 4. In the unit cell, the Mg atoms are located at the corners with coordinates, the metal atoms X- are in the center of the unit cell with coordinates, and the H atoms are at the octahedral sites at the center of faces. The assessment of the stability of materials MgCoH3, MgCuH3, and MgNiH3 has been determined utilizing the Birch-Murnaghan equation of state in conjugation with energy volume cures [175,176]. Mathematically, it can be obtained from
E t o t a l V = E 0 V + B 0 V B 0 1 B 0 1 V 0 V + V 0 V B 0 1 ,
where E0(V) is the total energy of the ground state of a unite cell at a given volume V, V0 is the equilibrium volume, B0 is the bulk modulus and B 0 stands its derivate with respect to pressure. Then, the gravimetric storage capacity of hydride perovskite, which is of importance in hydrogen storage applications, is estimated from the gravimetric ratio [177]
Cwt   [ % ]   = A H A M m H m M + A H A M m H × 100 [ % ] .
Here m and A stand for the molar mass and atoms, respectively; and subscript H stands for hydrogen and subscript M means the host perovskite material. Theoretically achievable amounts of hydrogen stored per unit mass of the host perovskite for all three Mg-based perovskites (i.e., Co, Cu, Ni) are summarized in Table 1. Note that the hydrogen desorption temperature (Tdes), which is an essential parameter for assessing the temperature required to release the stored hydrogen, can be determined from
T d e s = Δ H Δ S ,
where ΔH and ΔS are the formation energy and entropy change of hydrogen (130.7 J mol−1 K−1), respectively.
The structure of CaXH3 is similar to the above-discussed perovskite-type hydride (i.e., in the present case, Mg is just replaced by Ca in Figure 3). The fundamental properties of the CaXH3 are given in Table 2. After calculation, the lattice constant (Å), volume (V in Å3), density (ρ in g/cm3), formation enthalpies (ΔHf), and gravimetric hydrogen storage capacity (Cwt% in wt%) of the hydrides studied are given in Table 2. The negative values of the formation enthalpies listed in Table 2 indicate that CaXH3 (X: Mn, Fe, Co) perovskite-type hydrides are thermodynamically stable and can be synthesized. Among them, CaCoH3 has the highest stability, which is given by
Δ H f = E t C a X H 3 [ E C a + E X + 3 × E H ] ,
where Et (CaXH3) symbolize total energy of CaXH3 and E(Ca), E(X), and E(H) are the ground state energies for one Ca atom, one X atom, and one H atom, respectively [47].
The crystal structure of cubic AeSiH3 (Ae = K, Li, Na, Mg) compounds is shown in Figure 5. In the unit cell, the atoms Ae-are at the corners with coordinates, the Si atoms are located in the center of the unit cell with coordinates, and the H atoms are at the octahedral sites at the center of faces. The formation enthalpies of AeSiH3 hydrides are estimated as follows [169]:
Δ H f ( A e S i H 3 ) = [ E t A e S i H 3 [ E s A e + E s S i 3 E s H ]
where Es (Ae) shows the energy of Ae (Ae = Li, K, Na, Mg) atoms, Es (Si) shows the energy of Si atoms, and Es(H) shows the ground state energy of H atoms. Et shows the total energy of the compound. After calculation, the lattice constant (Å), volume (V in Å3), density (ρ in g/cm3), formation energy (ΔHf in eV/atom), and gravimetric hydrogen storage capacity (Cwt% in wt%) of the hydrides studied are summarized in Table 3. We emphasize here that the higher entropy values contribute to the release of hydrogen from perovskite materials. Perovskite materials with lower hydrogen adsorption-free energy values have a higher affinity for hydrogen gas, making them more likely to store hydrogen efficiently.
We only note that the structure of AeVH3, where Ae = Be, Mg, Ca, and Sr, is similar to the perovskite-type hydride presented in Figure 5 [170]. The lattice parameters (Å) are calculated as 3.48, 3.66, 3.73, and 3.83; The volumes (Å3) are calculated as 42.21, 49.22, 53.28, and 56.22, respectively, for AeVH3 with Ae = Be, Mg, Ca, and Sr. For the perovskite hydrides AeVH3 (where Ae = Be, Mg, Ca, Sr), the gravimetric hydrogen storage capacity decreases with increasing alkaline metal (Ae) mass, with values of 4.6, 3.7, 3.1, and 2.0 wt%, respectively. This trend shows that the decreasing gravimetric hydrogen storage capacity is due to the increasing mass of alkaline metals. Consequently, BeVH3 stands out as the most suitable material for hydrogen storage applications, with a higher gravimetric density of 4.6 wt% compared to the other three materials studied [170].
The following are perovskite-type hybrids with identical structures but with different atomic arrangements. A brief description of their structures is given below, followed by the calculated values. All calculated gravimetric hydrogen storage capacities of XSrH3 (X = K and Rb) are given in Table 4. KSrH3 has a greater capacity to store hydrogen (2.33%) compared to RbSrH3 (1.71%). In the context of illustrating thermodynamic properties, The Debye temperature is a crucial parameter in the discussion of the quasiharmonic Debye model and various thermodynamic properties. As the temperature increases, the Debye temperature decreases. Similarly, the Debye temperature increases at higher pressures. KSrH3 has a higher Debye temperature value compared to RbSrH3. It can, therefore, be said that KSrH3 is more stable than RbSrH3 [73].
All calculated gravimetric hydrogen storage capacities of XCuH3 (X = Co, Ni, Zn) are given in Table 5. NiCuH3 has a greater capacity to store hydrogen (3.0%) compared to CoCuH3 (2.8%) and ZnCuH3 (2.7%). In the thermodynamic properties, the negative values of free energy deliberate the thermodynamic stability of any material; the free energy of these three materials is all negative, indicating that they are stable materials in relation to each other. CoCuH3 is more stable than NiCuH3 and ZnCuH3 because CoCuH3 has the lowest free energy of them [171].
All calculated gravimetric hydrogen storage capacities of XCoH3 (X = In, Mn, Sr, Sn, Cd) are given in Table 6. MnCoH3 is most favored for hydrogen storage properties because its gravimetric ratio for hydrogen storage is greater than all these studied materials [172].
All calculated gravimetric hydrogen storage capacities of XPtH3 (X = Li, Na, K, Rb) are given in Table 7. Compared to other materials, LiPtH3 exhibits the highest gravimetric hydrogen storage capacities and stability (Tdes) [173].
All calculated gravimetric hydrogen storage capacities of XAlH3 (X = Na, K) are given in Table 8. NaAlH3 has a higher gravimetric hydrogen storage capacity than KAlH3. The formation energies of these compounds are negative, demonstrating their thermodynamic stability [174].
All calculated gravimetric hydrogen storage capacities of XTiH3 (X = K, Rb, Cs) are given in Table 9. The weight hydrogen storage capacities of KTiH3, RbTiH3, and CsTiH3 were determined to be 3.36, 2.22, and 1.65 wt%, respectively. The hydrogen desorption temperatures for KTiH3, RbTiH3, and CsTiH3 were determined to be 209 K, 161 K, and 107 K, respectively. These results indicate that KTiH3 has the potential to be a hydrogen storage material [77].
Mbonu et al. [160] used DFT to calculate the structural properties of the S-block halide perovskite KXCl3 (X = Be, Ca, Mg). In terms of thermodynamic performance, KMgCl3 has the highest heat capacity, while KCaCl3 has the lowest. In order to investigate the hydrogen storage capabilities of this perovskite system, functionalization was carried out by replacing chlorine (Cl) atoms in the crystal lattice with hydrogen (H) atoms. The resulting functionalized compounds, KBeH3, KMgH3, and KCaH3, are detailed in Table 10. KBeH3 has the best gravimetric hydrogen storage capacity of 5.866%, surpassing KMgH3 and KCaH3.

5. Perovskite-Type Oxides Structural, Thermodynamic, and Hydrogen Storage Properties

Ostadebrahim et al. [163] used a single-phase sol-gel method to synthesize ternary metal perovskite-type oxides LaMO3 (M = Cr, Mn, Fe, Co, Ni) and investigated their surface area, pore size, and electrochemical properties. The nanostructured materials parameters of the surface area and the pore size are crucial reference indicators for the electrochemical storage of hydrogen. According to the analysis results of BET-BJH, the BET surface area (SABET), total pore volume (Vt), average pore diameter (Davg), and SBJH of the samples are shown in Table 11. In the electrochemical performance of LaMO3 in a three-electrode electrochemical cell, LaFeO3 achieved the highest current density at the anodic peak. As shown in Table 10, LaFeO3 exhibited a morphology characterized by a spherical shape, the maximum number of nanoparticles per unit area, the highest total pore volume (Vt), and the lowest average pore diameter (Davg). This morphology is indicative of superior electrode response.
The electrochemical hydrogen storage capacity of LaMO3 nanocrystals is shown in Table 12. LaFeO3 shows the most favorable performance in terms of electrochemical hydrogen storage capacity. The lifetime and stability of the electrochemical hydrogen storage were investigated by galvanostatic charge-discharge tests and Tafel polarisation analysis. The LaFeO3 electrode exhibited stable discharge capacity over a current range of 1 to 4 mA. The Tafel test results showed the lowest corrosion rate for LaFeO3 nano-perovskites. These results show that LaFeO3 nano-perovskites are a promising candidate for energy storage applications [163].
Bhardwaj et al. [178] used perovskite oxide materials, specifically the proton insertion anode Sm1−xSrxCoO3−δ (x = 0, 0.5, 1), for nickel/oxide rechargeable batteries. Sm1−xSrxCoO3−δ (x = 0, 0.5, 1) was synthesized via solid-state reactions and its electrochemical hydrogen storage performance in alkaline electrolytes. The oxygen-deficient Sm0.5Sr0.5CoO3−δ (SSC) exhibits a maximum reversible discharge capacity of 182 mAh g−1, which increases with increasing temperature. Partial substitution of the acceptor dopant strontium with samarium induces lattice expansion, a greater extent of Co-O interactions, and increases the reducibility of cobalt ions, thereby promoting higher hydrogen storage.
Nabil et al. [179] studied the electronic, mechanical, and elastic properties of LaCrO3Hx (x = 0, 6) for hydrogen storage applications, which have been investigated. The determination included cell parameters, crystal structure, and mechanical performance. In terms of electronic properties, the Fermi level density of states (DOS) for LaCrO3 and LaCrO3H6 is exceptionally low. This indicates that the lower the DOS of the Fermi level, the more stable the compound. Table 13 shows that LaCrO3 has a higher bulk modulus value, and when hydrogen binds to LaCrO3, the bulk modulus decreases, indicating greater material compressibility. In conclusion, the LaCrO3 compound has a remarkable hydrogen storage capacity of about 4 wt%.
All calculated gravimetric hydrogen storage capacities of perovskite-type hydrides/oxide KMO3−xHx (x = 0, 0.3, 0.6, 0.9, 1.2, 1.5, 1.8, 2.1, 2.4, 2.7, and 3.0) are given in Table 14. The calculated lattice parameters indicate that KMg has a cubic crystal structure in KMgO3 and KMgH3. However, for other doping concentrations, the cubic crystal structure is distorted when anions are doped on the anion side [180].

6. Future Trends

Perovskite solar cells show immense potential to become the next photovoltaic commercial technology. Future trends in perovskite solar cells focus on enhancing their efficiency, stability, scalability, and sustainability. A possible way to improve efficiency is by using tandem solar cells. For instance, there is a trend in integrating perovskite layers with silicon or materials like CIGS (copper indium gallium selenide) to form tandem solar cells. These configurations have already surpassed 30% efficiency. Another challenge for PSCs is their stability. To address this, researchers must focus on interface engineering to overcome various defects, non-radiative recombination, and energy band alignment. Scalability might be overcome by using roll-to-roll processing, which is a high-speed and low-cost process and results in the fabrication of flexible and lightweight perovskite modules [181]. The printing technique could also be optimized for the large-area production of perovskite solar cells [182]. Regarding sustainability and toxic reduction, reducing the amount of lead content in PSCs is crucial for minimizing environmental impact. There is a development of lead-free alternatives like tin-based perovskites. Ultimately, achieving high efficiency, long-term stability, and reduced environmental impact in PSCs will provide significant societal benefits [183,184].
In perovskite for hydrogen storage, the future trends focus on expanding their applications in hydrogen production, storage, and conversion processes. The key areas that need special focus and further development are perovskites for Hydrogen Evolution Reactions (HER), hydrogen storage materials, and hydrogen sensors. Perovskites are being explored as efficient, low-cost catalysts for hydrogen production through water splitting. In the future, improving the catalytic activity of perovskites by tuning their composition and surface properties is crucial. The next step, is developing new perovskite compositions and structures and exploring the integration of perovskite with other materials to optimize hydrogen storage capacities [185].

7. Conclusion Remarks

In this review, we have presented not only the applications of perovskite solar cells but also the hydrogen storage capacities of perovskite-type materials. Due to their excellent structural properties and cost-effectiveness, perovskite materials offer substantial research potential in the field of hydrogen storage.
The main future advancements for perovskite solar cells include the development of tandem solar cells, flexible solar cells, and space solar cells. Among perovskite-type hydrides, LiSiH3 exhibits the highest gravimetric hydrogen storage capacity at 7.946%, highlighting the significant potential of this compound series for hydrogen storage. Additionally, perovskite-type oxides demonstrate excellent hydrogen adsorption and storage capabilities in electrode applications, leading to increased electrode capacity and robust cycling stability. The low application cost of these materials in energy storage systems further emphasizes the development potential of perovskite-type oxides for hydrogen storage. These findings collectively underscore the promising prospects of perovskite-type hydrides and oxides in the field of hydrogen storage. Overall, the present results are highly significant for future research on perovskite-type hydrogen storage compounds and the development of perovskite solar cells.

Author Contributions

M.-H.K., N.N. and I.S.—conceptualization; M.-H.K.—writing original draft; N.N. and I.S.—the finalization of the manuscript and editing; I.S.—supervision; N.N. and I.S. funding acquiring. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Czech Science Foundation GA ČR, grant no. 23-06543S and by Czech Ministry of Education, Youth and Sports grant no. CZ.02.01.01/00/22_008/0004617—“Energy conversion and storage”.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ipsakis, D.; Voutetakis, S.; Seferlis, P.; Stergiopoulos, F.; Elmasides, C. Power Management Strategies for a Stand-Alone Power System Using Renewable Energy Sources and Hydrogen Storage. Int. J. Hydrogen Energy 2009, 34, 7081–7095. [Google Scholar] [CrossRef]
  2. Kane, M. Small Hybrid Solar Power System. Energy 2003, 28, 1427–1443. [Google Scholar] [CrossRef]
  3. Sakintuna, B.; Lamaridarkrim, F.; Hirscher, M. Metal Hydride Materials for Solid Hydrogen Storage: A Review☆. Int. J. Hydrogen Energy 2007, 32, 1121–1140. [Google Scholar] [CrossRef]
  4. Endo, N.; Goshome, K.; Tetsuhiko, M.; Segawa, Y.; Shimoda, E.; Nozu, T. Thermal Management and Power Saving Operations for Improved Energy Efficiency within a Renewable Hydrogen Energy System Utilizing Metal Hydride Hydrogen Storage. Int. J. Hydrogen Energy 2021, 46, 262–271. [Google Scholar] [CrossRef]
  5. Muellerlanger, F.; Tzimas, E.; Kaltschmitt, M.; Peteves, S. Techno-Economic Assessment of Hydrogen Production Processes for the Hydrogen Economy for the Short and Medium Term. Int. J. Hydrogen Energy 2007, 32, 3797–3810. [Google Scholar] [CrossRef]
  6. Darmawi; Sipahutar, R.; Bernas, S.M.; Imanuddin, M.S. Renewable Energy and Hydropower Utilization Tendency Worldwide. Renew. Sustain. Energy Rev. 2013, 17, 213–215. [Google Scholar] [CrossRef]
  7. Maradin, D. Advantages and Disadvantages of Renewable Energy Sources Utilization. Int. J. Energy Econ. Policy 2021, 11, 176–183. [Google Scholar] [CrossRef]
  8. Shetty, C.; Priyam, A. A Review on Tidal Energy Technologies. Mater. Today Proc. 2022, 56, 2774–2779. [Google Scholar] [CrossRef]
  9. Marques Lameirinhas, R.A.; Torres, J.P.N.; De Melo Cunha, J.P. A Photovoltaic Technology Review: History, Fundamentals and Applications. Energies 2022, 15, 1823. [Google Scholar] [CrossRef]
  10. Ajayan, J.; Nirmal, D.; Mohankumar, P.; Saravanan, M.; Jagadesh, M.; Arivazhagan, L. A Review of Photovoltaic Performance of Organic/Inorganic Solar Cells for Future Renewable and Sustainable Energy Technologies. Superlattices Microstruct. 2020, 143, 106549. [Google Scholar] [CrossRef]
  11. Mahmood, Q.; Nazir, G.; Bouzgarrou, S.; Aljameel, A.I.; Rehman, A.; Albalawi, H.; Haq, B.U.; Ghrib, T.; Mera, A. Study of New Lead-Free Double Perovskites Halides Tl2TiX6 (X = Cl, Br, I) for Solar Cells and Renewable Energy Devices. J. Solid State Chem. 2022, 308, 122887. [Google Scholar] [CrossRef]
  12. Chapin, D.M.; Fuller, C.S.; Pearson, G.L. A New Silicon P-n Junction Photocell for Converting Solar Radiation into Electrical Power. J. Appl. Phys. 1954, 25, 676–677. [Google Scholar] [CrossRef]
  13. Dannenberg, T.; Vollmer, J.; Passig, M.; Scheiwe, C.; Brunner, D.; Pediaditakis, A.; Jäger, U.; Wang, I.; Xie, W.; Xu, S.; et al. Past, Present, and Future Outlook for Edge Isolation Processes in Highly Efficient Silicon Solar Cell Manufacturing. Sol. RRL 2023, 7, 2200594. [Google Scholar] [CrossRef]
  14. Fraas, L.M.; O’Neill, M.J. History of Solar Cell Development. In Low-Cost Solar Electric Power; Springer International Publishing: Cham, Switzerland, 2023; pp. 1–12. ISBN 978-3-031-30811-6. [Google Scholar]
  15. Saga, T. Advances in Crystalline Silicon Solar Cell Technology for Industrial Mass Production. NPG Asia Mater. 2010, 2, 96–102. [Google Scholar] [CrossRef]
  16. Mercaldo, L.V.; Delli Veneri, P. Silicon Solar Cells: Materials, Technologies, Architectures. In Solar Cells and Light Management; Elsevier: Amsterdam, The Netherlands, 2020; pp. 35–57. ISBN 978-0-08-102762-2. [Google Scholar]
  17. Andreani, L.C.; Bozzola, A.; Kowalczewski, P.; Liscidini, M.; Redorici, L. Silicon Solar Cells: Toward the Efficiency Limits. Adv. Phys. X 2019, 4, 1548305. [Google Scholar] [CrossRef]
  18. Yoshikawa, K.; Kawasaki, H.; Yoshida, W.; Irie, T.; Konishi, K.; Nakano, K.; Uto, T.; Adachi, D.; Kanematsu, M.; Uzu, H.; et al. Silicon Heterojunction Solar Cell with Interdigitated Back Contacts for a Photoconversion Efficiency over 26%. Nat. Energy 2017, 2, 17032. [Google Scholar] [CrossRef]
  19. Kim, S.; Quy, H.V.; Bark, C.W. Photovoltaic Technologies for Flexible Solar Cells: Beyond Silicon. Mater. Today Energy 2021, 19, 100583. [Google Scholar] [CrossRef]
  20. Kokkonen, M.; Talebi, P.; Zhou, J.; Asgari, S.; Soomro, S.A.; Elsehrawy, F.; Halme, J.; Ahmad, S.; Hagfeldt, A.; Hashmi, S.G. Advanced Research Trends in Dye-Sensitized Solar Cells. J. Mater. Chem. A 2021, 9, 10527–10545. [Google Scholar] [CrossRef]
  21. Schmidt, J.; Lim, B.; Walter, D.; Bothe, K.; Gatz, S.; Dullweber, T.; Altermatt, P.P. Impurity-Related Limitations of next-Generation Industrial Silicon Solar Cells. In Proceedings of the 2012 IEEE 38th Photovoltaic Specialists Conference (PVSC) PART 2, Austin, TX, USA, 3–8 June 2012; IEEE: Piscataway, NJ, USA, 2012; pp. 1–5. [Google Scholar]
  22. He, D.; Zeng, M.; Zhang, Z.; Bai, Y.; Xing, G.; Cheng, H.; Lin, Y. Exciton Diffusion and Dissociation in Organic and Quantum-dot Solar Cells. SmartMat 2023, 4, e1176. [Google Scholar] [CrossRef]
  23. Yi, J.; Zhang, G.; Yu, H.; Yan, H. Advantages, Challenges and Molecular Design of Different Material Types Used in Organic Solar Cells. Nat. Rev. Mater. 2023, 9, 46–62. [Google Scholar] [CrossRef]
  24. Prajapat, K.; Dhonde, M.; Sahu, K.; Bhojane, P.; Murty, V.; Shirage, P.M. The Evolution of Organic Materials for Efficient Dye-Sensitized Solar Cells. J. Photochem. Photobiol. C Photochem. Rev. 2023, 55, 100586. [Google Scholar] [CrossRef]
  25. Koné, K.E.; Bouich, A.; Soucase, B.M.; Soro, D. Manufacture of Different Oxides with High Uniformity for Copper Zinc Tin Sulfide (CZTS) Based Solar Cells. J. Mol. Graph. Model. 2023, 121, 108448. [Google Scholar] [CrossRef] [PubMed]
  26. Aftab, S.; Hussain, S.; Kabir, F.; Aslam, M.; Rajpar, A.H.; Al-Sehemi, A.G. Advances in Flexible Perovskite Solar Cells: A Comprehensive Review. Nano Energy 2024, 120, 109112. [Google Scholar] [CrossRef]
  27. Manser, J.S.; Christians, J.A.; Kamat, P.V. Intriguing Optoelectronic Properties of Metal Halide Perovskites. Chem. Rev. 2016, 116, 12956–13008. [Google Scholar] [CrossRef] [PubMed]
  28. Kovalenko, M.V.; Protesescu, L.; Bodnarchuk, M.I. Properties and Potential Optoelectronic Applications of Lead Halide Perovskite Nanocrystals. Science 2017, 358, 745–750. [Google Scholar] [CrossRef]
  29. Stranks, S.D.; Eperon, G.E.; Grancini, G.; Menelaou, C.; Alcocer, M.J.P.; Leijtens, T.; Herz, L.M.; Petrozza, A.; Snaith, H.J. Electron-Hole Diffusion Lengths Exceeding 1 Micrometer in an Organometal Trihalide Perovskite Absorber. Science 2013, 342, 341–344. [Google Scholar] [CrossRef]
  30. Shrestha, S.; Li, X.; Tsai, H.; Hou, C.-H.; Huang, H.-H.; Ghosh, D.; Shyue, J.-J.; Wang, L.; Tretiak, S.; Ma, X.; et al. Long Carrier Diffusion Length in Two-Dimensional Lead Halide Perovskite Single Crystals. Chem 2022, 8, 1107–1120. [Google Scholar] [CrossRef]
  31. Miyata, A.; Mitioglu, A.; Plochocka, P.; Portugall, O.; Wang, J.T.-W.; Stranks, S.D.; Snaith, H.J.; Nicholas, R.J. Direct Measurement of the Exciton Binding Energy and Effective Masses for Charge Carriers in Organic–Inorganic Tri-Halide Perovskites. Nat. Phys. 2015, 11, 582–587. [Google Scholar] [CrossRef]
  32. Hansen, K.R.; McClure, C.E.; Powell, D.; Hsieh, H.; Flannery, L.; Garden, K.; Miller, E.J.; King, D.J.; Sainio, S.; Nordlund, D.; et al. Low Exciton Binding Energies and Localized Exciton–Polaron States in 2D Tin Halide Perovskites. Adv. Opt. Mater. 2022, 10, 2102698. [Google Scholar] [CrossRef]
  33. Li, Y.; Lu, Y.; Huo, X.; Wei, D.; Meng, J.; Dong, J.; Qiao, B.; Zhao, S.; Xu, Z.; Song, D. Bandgap Tuning Strategy by Cations and Halide Ions of Lead Halide Perovskites Learned from Machine Learning. RSC Adv. 2021, 11, 15688–15694. [Google Scholar] [CrossRef]
  34. Tan, Q.; Li, Z.; Luo, G.; Zhang, X.; Che, B.; Chen, G.; Gao, H.; He, D.; Ma, G.; Wang, J.; et al. Inverted Perovskite Solar Cells Using Dimethylacridine-Based Dopants. Nature 2023, 620, 545–551. [Google Scholar] [CrossRef] [PubMed]
  35. Huang, Z.; Bai, Y.; Huang, X.; Li, J.; Wu, Y.; Chen, Y.; Li, K.; Niu, X.; Li, N.; Liu, G.; et al. Anion–π Interactions Suppress Phase Impurities in FAPbI3 Solar Cells. Nature 2023, 623, 531–537. [Google Scholar] [CrossRef] [PubMed]
  36. Roy, P.; Kumar Sinha, N.; Tiwari, S.; Khare, A. A Review on Perovskite Solar Cells: Evolution of Architecture, Fabrication Techniques, Commercialization Issues and Status. Sol. Energy 2020, 198, 665–688. [Google Scholar] [CrossRef]
  37. Meng, L.; You, J.; Yang, Y. Addressing the Stability Issue of Perovskite Solar Cells for Commercial Applications. Nat. Commun. 2018, 9, 5265. [Google Scholar] [CrossRef] [PubMed]
  38. Yang, D.; Yang, R.; Priya, S.; Liu, S. (Frank) Recent Advances in Flexible Perovskite Solar Cells: Fabrication and Applications. Angew. Chem. Int. Ed. 2019, 58, 4466–4483. [Google Scholar] [CrossRef]
  39. Liang, J.; Wang, C.; Wang, Y.; Xu, Z.; Lu, Z.; Ma, Y.; Zhu, H.; Hu, Y.; Xiao, C.; Yi, X.; et al. All-Inorganic Perovskite Solar Cells. J. Am. Chem. Soc. 2016, 138, 15829–15832. [Google Scholar] [CrossRef]
  40. Li, Y.; Yang, C.; Guo, W.; Duan, T.; Zhou, Z.; Zhou, Y. All-Inorganic Perovskite Solar Cells Featuring Mixed Group IVA Cations. Nanoscale 2023, 15, 7249–7260. [Google Scholar] [CrossRef]
  41. Jeon, N.J.; Noh, J.H.; Kim, Y.C.; Yang, W.S.; Ryu, S.; Seok, S.I. Solvent Engineering for High-Performance Inorganic–Organic Hybrid Perovskite Solar Cells. Nat. Mater. 2014, 13, 897–903. [Google Scholar] [CrossRef]
  42. Zhang, J.; Tang, S.; Zhu, M.; Li, Z.; Cheng, Z.; Xiang, S.; Zhang, Z. The Role of Grain Boundaries in Organic–Inorganic Hybrid Perovskite Solar Cells and Its Current Enhancement Strategies: A Review. Energy Environ. Mater. 2024, 7, e12696. [Google Scholar] [CrossRef]
  43. Rao, C.N.R. Perovskites. In Encyclopedia of Physical Science and Technology; Elsevier: Amsterdam, The Netherlands, 2003; pp. 707–714. ISBN 978-0-12-227410-7. [Google Scholar]
  44. King-Smith, R.D.; Vanderbilt, D. First-Principles Investigation of Ferroelectricity in Perovskite Compounds. Phys. Rev. B 1994, 49, 5828–5844. [Google Scholar] [CrossRef]
  45. Zhang, H.; Li, N.; Li, K.; Xue, D. Structural Stability and Formability of ABO3-Type Perovskite Compounds. Acta Crystallogr. B 2007, 63, 812–818. [Google Scholar] [CrossRef] [PubMed]
  46. Bartel, C.J.; Sutton, C.; Goldsmith, B.R.; Ouyang, R.; Musgrave, C.B.; Ghiringhelli, L.M.; Scheffler, M. New Tolerance Factor to Predict the Stability of Perovskite Oxides and Halides. Sci. Adv. 2019, 5, eaav0693. [Google Scholar] [CrossRef] [PubMed]
  47. Surucu, G.; Gencer, A.; Candan, A.; Gullu, H.H.; Isik, M. CaXH3 (X = Mn, Fe, Co) Perovskite-Type Hydrides for Hydrogen Storage Applications. Int. J. Energy Res. 2020, 44, 2345–2354. [Google Scholar] [CrossRef]
  48. Schouwink, P.; Ley, M.B.; Tissot, A.; Hagemann, H.; Jensen, T.R.; Smrčok, Ľ.; Černý, R. Structure and Properties of Complex Hydride Perovskite Materials. Nat. Commun. 2014, 5, 5706. [Google Scholar] [CrossRef] [PubMed]
  49. Peña, M.A.; Fierro, J.L.G. Chemical Structures and Performance of Perovskite Oxides. Chem. Rev. 2001, 101, 1981–2018. [Google Scholar] [CrossRef] [PubMed]
  50. Jena, A.K.; Kulkarni, A.; Miyasaka, T. Halide Perovskite Photovoltaics: Background, Status, and Future Prospects. Chem. Rev. 2019, 119, 3036–3103. [Google Scholar] [CrossRef]
  51. Bannikov, V.V.; Shein, I.R.; Ivanovskii, A.L. Electronic Structure, Chemical Bonding and Elastic Properties of the First Thorium-Containing Nitride Perovskite TaThN3. Phys. Status Solidi RRL Rapid Res. Lett. 2007, 1, 89–91. [Google Scholar] [CrossRef]
  52. Guo, P.; Ye, Q.; Yang, X.; Zhang, J.; Xu, F.; Shchukin, D.; Wei, B.; Wang, H. Surface & Grain Boundary Co-Passivation by Fluorocarbon Based Bifunctional Molecules for Perovskite Solar Cells with Efficiency over 21%. J. Mater. Chem. A 2019, 7, 2497–2506. [Google Scholar] [CrossRef]
  53. Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal Halide Perovskites as Visible-Light Sensitizers for Photovoltaic Cells. J. Am. Chem. Soc. 2009, 131, 6050–6051. [Google Scholar] [CrossRef]
  54. Yoo, J.J.; Seo, G.; Chua, M.R.; Park, T.G.; Lu, Y.; Rotermund, F.; Kim, Y.-K.; Moon, C.S.; Jeon, N.J.; Correa-Baena, J.-P.; et al. Efficient Perovskite Solar Cells via Improved Carrier Management. Nature 2021, 590, 587–593. [Google Scholar] [CrossRef]
  55. Szabó, G.; Park, N.-G.; De Angelis, F.; Kamat, P.V. Are Perovskite Solar Cells Reaching the Efficiency and Voltage Limits? ACS Energy Lett. 2023, 8, 3829–3831. [Google Scholar] [CrossRef]
  56. Paik, M.J.; Kim, Y.Y.; Kim, J.; Park, J.; Seok, S.I. Ultrafine SnO2 Colloids with Enhanced Interface Quality for High-Efficiency Perovskite Solar Cells. Joule 2024, 8, 2073–2086. [Google Scholar] [CrossRef]
  57. Li, X.; Yu, H.; Liu, Z.; Huang, J.; Ma, X.; Liu, Y.; Sun, Q.; Dai, L.; Ahmad, S.; Shen, Y.; et al. Progress and Challenges Toward Effective Flexible Perovskite Solar Cells. Nano-Micro Lett. 2023, 15, 206. [Google Scholar] [CrossRef] [PubMed]
  58. Kim, B.J.; Kim, D.H.; Lee, Y.-Y.; Shin, H.-W.; Han, G.S.; Hong, J.S.; Mahmood, K.; Ahn, T.K.; Joo, Y.-C.; Hong, K.S.; et al. Highly Efficient and Bending Durable Perovskite Solar Cells: Toward a Wearable Power Source. Energy Environ. Sci. 2015, 8, 916–921. [Google Scholar] [CrossRef]
  59. Chung, J.; Shin, S.S.; Hwang, K.; Kim, G.; Kim, K.W.; Lee, D.S.; Kim, W.; Ma, B.S.; Kim, Y.-K.; Kim, T.-S.; et al. Record-Efficiency Flexible Perovskite Solar Cell and Module Enabled by a Porous-Planar Structure as an Electron Transport Layer. Energy Environ. Sci. 2020, 13, 4854–4861. [Google Scholar] [CrossRef]
  60. Paik, M.J.; Yoo, J.W.; Park, J.; Noh, E.; Kim, H.; Ji, S.-G.; Kim, Y.Y.; Seok, S.I. SnO2—TiO2 Hybrid Electron Transport Layer for Efficient and Flexible Perovskite Solar Cells. ACS Energy Lett. 2022, 7, 1864–1870. [Google Scholar] [CrossRef]
  61. Gencer, A.; Surucu, G.; Al, S. MgTiO3Hx and CaTiO3Hx Perovskite Compounds for Hydrogen Storage Applications. Int. J. Hydrogen Energy 2019, 44, 11930–11938. [Google Scholar] [CrossRef]
  62. Ikeda, K.; Kogure, Y.; Nakamori, Y.; Orimo, S. Formation Region and Hydrogen Storage Abilities of Perovskite-Type Hydrides. Prog. Solid State Chem. 2007, 35, 329–337. [Google Scholar] [CrossRef]
  63. Schmidt, J.; Pettersson, L.; Verdozzi, C.; Botti, S.; Marques, M.A.L. Crystal Graph Attention Networks for the Prediction of Stable Materials. Sci. Adv. 2021, 7, eabi7948. [Google Scholar] [CrossRef]
  64. Jain, I.P. Hydrogen the Fuel for 21st Century. Int. J. Hydrogen Energy 2009, 34, 7368–7378. [Google Scholar] [CrossRef]
  65. Veziroğlu, T.N.; Şahi˙n, S. 21st Century’s Energy: Hydrogen Energy System. Energy Convers. Manag. 2008, 49, 1820–1831. [Google Scholar] [CrossRef]
  66. Ball, M.; Wietschel, M. The Future of Hydrogen—Opportunities and Challenges. Int. J. Hydrogen Energy 2009, 34, 615–627. [Google Scholar] [CrossRef]
  67. Subramanian, B.; Ismail, S. Production and Use of HHO Gas in IC Engines. Int. J. Hydrogen Energy 2018, 43, 7140–7154. [Google Scholar] [CrossRef]
  68. Sopena, C.; Diéguez, P.M.; Sáinz, D.; Urroz, J.C.; Guelbenzu, E.; Gandía, L.M. Conversion of a Commercial Spark Ignition Engine to Run on Hydrogen: Performance Comparison Using Hydrogen and Gasoline. Int. J. Hydrogen Energy 2010, 35, 1420–1429. [Google Scholar] [CrossRef]
  69. Zhang, Y.; Campana, P.E.; Lundblad, A.; Yan, J. Comparative Study of Hydrogen Storage and Battery Storage in Grid Connected Photovoltaic System: Storage Sizing and Rule-Based Operation. Appl. Energy 2017, 201, 397–411. [Google Scholar] [CrossRef]
  70. Yang, J.; Sudik, A.; Wolverton, C.; Siegel, D.J. High Capacity Hydrogenstorage Materials: Attributes for Automotive Applications and Techniques for Materials Discovery. Chem. Soc. Rev. 2010, 39, 656–675. [Google Scholar] [CrossRef]
  71. Murray, L.J.; Dincă, M.; Long, J.R. Hydrogen Storage in Metal–Organic Frameworks. Chem. Soc. Rev. 2009, 38, 1294. [Google Scholar] [CrossRef]
  72. Liu, P.; Liang, J.; Xue, R.; Du, Q.; Jiang, M. Ruthenium Decorated Boron-Doped Carbon Nanotube for Hydrogen Storage: A First-Principle Study. Int. J. Hydrogen Energy 2019, 44, 27853–27861. [Google Scholar] [CrossRef]
  73. Raza, H.H.; Murtaza, G.; Umm-e-Hani; Muhammad, N.; Ramay, S.M. First-Principle Investigation of XSrH3 (X = K and Rb) Perovskite-Type Hydrides for Hydrogen Storage. Int. J. Quantum Chem. 2020, 120, e26419. [Google Scholar] [CrossRef]
  74. Bouhadda, Y.; Boudouma, Y.; Fennineche, N.; Bentabet, A. Ab Initio Calculations Study of the Electronic, Optical and Thermodynamic Properties of NaMgH3, for Hydrogen Storage. J. Phys. Chem. Solids 2010, 71, 1264–1268. [Google Scholar] [CrossRef]
  75. Xu, Y.; Zhou, Y.; Li, Y.; Ding, Z. Research Progress and Application Prospects of Solid-State Hydrogen Storage Technology. Molecules 2024, 29, 1767. [Google Scholar] [CrossRef] [PubMed]
  76. Wei, T.Y.; Lim, K.L.; Tseng, Y.S.; Chan, S.L.I. A Review on the Characterization of Hydrogen in Hydrogen Storage Materials. Renew. Sustain. Energy Rev. 2017, 79, 1122–1133. [Google Scholar] [CrossRef]
  77. Xu, N.; Chen, Y.; Chen, S.; Li, S.; Zhang, W. First-Principles Investigation for the Hydrogen Storage Properties of XTiH3 (X = K, Rb, Cs) Perovskite Type Hydrides. Int. J. Hydrogen Energy 2024, 50, 114–122. [Google Scholar] [CrossRef]
  78. Kojima, Y. Hydrogen Storage Materials for Hydrogen and Energy Carriers. Int. J. Hydrogen Energy 2019, 44, 18179–18192. [Google Scholar] [CrossRef]
  79. Züttel, A. Hydrogen Storage Methods. Naturwissenschaften 2004, 91, 157–172. [Google Scholar] [CrossRef]
  80. Pukazhselvan, D.; Kumar, V.; Singh, S.K. High Capacity Hydrogen Storage: Basic Aspects, New Developments and Milestones. Nano Energy 2012, 1, 566–589. [Google Scholar] [CrossRef]
  81. Schlapbach, L.; Züttel, A. Hydrogen-Storage Materials for Mobile Applications. Nature 2001, 414, 353–358. [Google Scholar] [CrossRef]
  82. Broom, D.P. Hydrogen Storage Materials: The Characterisation of Their Storage Properties; Green Energy and Technology; Springer London: London, UK, 2011; ISBN 978-0-85729-220-9. [Google Scholar]
  83. Ikeda, K.; Sato, T.; Orimo, S. Perovskite-Type Hydrides—Synthesis, Structures and Properties. Int. J. Mater. Res. 2008, 99, 471–479. [Google Scholar] [CrossRef]
  84. Young, K.; Ng, K.; Bendersky, L. A Technical Report of the Robust Affordable Next Generation Energy Storage System-BASF Program. Batteries 2016, 2, 2. [Google Scholar] [CrossRef]
  85. Bertuol, D.A.; Bernardes, A.M.; Tenório, J.A.S. Spent NiMH Batteries: Characterization and Metal Recovery through Mechanical Processing. J. Power Sources 2006, 160, 1465–1470. [Google Scholar] [CrossRef]
  86. Al-Thyabat, S.; Nakamura, T.; Shibata, E.; Iizuka, A. Adaptation of Minerals Processing Operations for Lithium-Ion (LiBs) and Nickel Metal Hydride (NiMH) Batteries Recycling: Critical Review. Miner. Eng. 2013, 45, 4–17. [Google Scholar] [CrossRef]
  87. Mohtadi, R.; Orimo, S. The Renaissance of Hydrides as Energy Materials. Nat. Rev. Mater. 2016, 2, 16091. [Google Scholar] [CrossRef]
  88. Nie, T.; Fang, Z.; Ren, X.; Duan, Y.; Liu, S. Recent Advances in Wide-Bandgap Organic–Inorganic Halide Perovskite Solar Cells and Tandem Application. Nano-Micro Lett. 2023, 15, 70. [Google Scholar] [CrossRef] [PubMed]
  89. Mali, S.S.; Patil, J.V.; Shao, J.-Y.; Zhong, Y.-W.; Rondiya, S.R.; Dzade, N.Y.; Hong, C.K. Phase-Heterojunction All-Inorganic Perovskite Solar Cells Surpassing 21.5% Efficiency. Nat. Energy 2023, 8, 989–1001. [Google Scholar] [CrossRef]
  90. Tong, J.; Jiang, Q.; Zhang, F.; Kang, S.B.; Kim, D.H.; Zhu, K. Wide-Bandgap Metal Halide Perovskites for Tandem Solar Cells. ACS Energy Lett. 2021, 6, 232–248. [Google Scholar] [CrossRef] [PubMed]
  91. Eperon, G.E.; Hörantner, M.T.; Snaith, H.J. Metal Halide Perovskite Tandem and Multiple-Junction Photovoltaics. Nat. Rev. Chem. 2017, 1, 0095. [Google Scholar] [CrossRef]
  92. Ou, Q.; Bao, X.; Zhang, Y.; Shao, H.; Xing, G.; Li, X.; Shao, L.; Bao, Q. Band Structure Engineering in Metal Halide Perovskite Nanostructures for Optoelectronic Applications. Nano Mater. Sci. 2019, 1, 268–287. [Google Scholar] [CrossRef]
  93. Liu, X.; Luo, D.; Lu, Z.-H.; Yun, J.S.; Saliba, M.; Seok, S.I.; Zhang, W. Stabilization of Photoactive Phases for Perovskite Photovoltaics. Nat. Rev. Chem. 2023, 7, 462–479. [Google Scholar] [CrossRef]
  94. Simenas, M.; Gagor, A.; Banys, J.; Maczka, M. Phase Transitions and Dynamics in Mixed Three- and Low-Dimensional Lead Halide Perovskites. Chem. Rev. 2024, 124, 2281–2326. [Google Scholar] [CrossRef]
  95. Saski, M.; Sobczak, S.; Ratajczyk, P.; Terlecki, M.; Marynowski, W.; Borkenhagen, A.; Justyniak, I.; Katrusiak, A.; Lewiński, J. Unprecedented Richness of Temperature- and Pressure-Induced Polymorphism in 1D Lead Iodide Perovskite. Small 2024, 2403685. [Google Scholar] [CrossRef]
  96. Noh, J.H.; Im, S.H.; Heo, J.H.; Mandal, T.N.; Seok, S.I. Chemical Management for Colorful, Efficient, and Stable Inorganic–Organic Hybrid Nanostructured Solar Cells. Nano Lett. 2013, 13, 1764–1769. [Google Scholar] [CrossRef] [PubMed]
  97. Chiang, Y.-H.; Anaya, M.; Stranks, S.D. Multisource Vacuum Deposition of Methylammonium-Free Perovskite Solar Cells. ACS Energy Lett. 2020, 5, 2498–2504. [Google Scholar] [CrossRef] [PubMed]
  98. Wang, S.; Yousefi Amin, A.A.; Wu, L.; Cao, M.; Zhang, Q.; Ameri, T. Perovskite Nanocrystals: Synthesis, Stability, and Optoelectronic Applications. Small Struct. 2021, 2, 2000124. [Google Scholar] [CrossRef]
  99. Prochowicz, D.; Saski, M.; Yadav, P.; Grätzel, M.; Lewiński, J. Mechanoperovskites for Photovoltaic Applications: Preparation, Characterization, and Device Fabrication. Acc. Chem. Res. 2019, 52, 3233–3243. [Google Scholar] [CrossRef] [PubMed]
  100. Diodati, S.; Nodari, L.; Natile, M.M.; Caneschi, A.; De Julián Fernández, C.; Hoffmann, C.; Kaskel, S.; Lieb, A.; Di Noto, V.; Mascotto, S.; et al. Coprecipitation of Oxalates: An Easy and Reproducible Wet-Chemistry Synthesis Route for Transition-Metal Ferrites. Eur. J. Inorg. Chem. 2014, 2014, 875–887. [Google Scholar] [CrossRef]
  101. Yu, J.; Ran, R.; Zhong, Y.; Zhou, W.; Ni, M.; Shao, Z. Advances in Porous Perovskites: Synthesis and Electrocatalytic Performance in Fuel Cells and Metal–Air Batteries. Energy Environ. Mater. 2020, 3, 121–145. [Google Scholar] [CrossRef]
  102. Assirey, E.A.R. Perovskite Synthesis, Properties and Their Related Biochemical and Industrial Application. Saudi Pharm. J. 2019, 27, 817–829. [Google Scholar] [CrossRef]
  103. Prochowicz, D.; Yadav, P.; Saliba, M.; Kubicki, D.J.; Tavakoli, M.M.; Zakeeruddin, S.M.; Lewiński, J.; Emsley, L.; Grätzel, M. One-Step Mechanochemical Incorporation of an Insoluble Cesium Additive for High Performance Planar Heterojunction Solar Cells. Nano Energy 2018, 49, 523–528. [Google Scholar] [CrossRef]
  104. Al Amin, N.R.; Lee, C.-C.; Huang, Y.-C.; Shih, C.-J.; Estrada, R.; Biring, S.; Kuo, M.-H.; Li, C.-F.; Huang, Y.-C.; Liu, S.-W. Achieving a Highly Stable Perovskite Photodetector with a Long Lifetime Fabricated via an All-Vacuum Deposition Process. ACS Appl. Mater. Interfaces 2023, 15, 21284–21295. [Google Scholar] [CrossRef]
  105. Yin, W.-J.; Yang, J.-H.; Kang, J.; Yan, Y.; Wei, S.-H. Halide Perovskite Materials for Solar Cells: A Theoretical Review. J. Mater. Chem. A 2015, 3, 8926–8942. [Google Scholar] [CrossRef]
  106. Afre, R.A.; Pugliese, D. Perovskite Solar Cells: A Review of the Latest Advances in Materials, Fabrication Techniques, and Stability Enhancement Strategies. Micromachines 2024, 15, 192. [Google Scholar] [CrossRef] [PubMed]
  107. Frost, J.M.; Butler, K.T.; Brivio, F.; Hendon, C.H.; Van Schilfgaarde, M.; Walsh, A. Atomistic Origins of High-Performance in Hybrid Halide Perovskite Solar Cells. Nano Lett. 2014, 14, 2584–2590. [Google Scholar] [CrossRef] [PubMed]
  108. Berhe, T.A.; Su, W.-N.; Chen, C.-H.; Pan, C.-J.; Cheng, J.-H.; Chen, H.-M.; Tsai, M.-C.; Chen, L.-Y.; Dubale, A.A.; Hwang, B.-J. Organometal Halide Perovskite Solar Cells: Degradation and Stability. Energy Environ. Sci. 2016, 9, 323–356. [Google Scholar] [CrossRef]
  109. Marinova, N.; Valero, S.; Delgado, J.L. Organic and Perovskite Solar Cells: Working Principles, Materials and Interfaces. J. Colloid Interface Sci. 2017, 488, 373–389. [Google Scholar] [CrossRef] [PubMed]
  110. Desoky, M.M.H.; Bonomo, M.; Buscaino, R.; Fin, A.; Viscardi, G.; Barolo, C.; Quagliotto, P. Dopant-Free All-Organic Small-Molecule HTMs for Perovskite Solar Cells: Concepts and Structure–Property Relationships. Energies 2021, 14, 2279. [Google Scholar] [CrossRef]
  111. Marchioro, A.; Teuscher, J.; Friedrich, D.; Kunst, M.; Van De Krol, R.; Moehl, T.; Grätzel, M.; Moser, J.-E. Unravelling the Mechanism of Photoinduced Charge Transfer Processes in Lead Iodide Perovskite Solar Cells. Nat. Photonics 2014, 8, 250–255. [Google Scholar] [CrossRef]
  112. Rong, Y.; Hu, Y.; Mei, A.; Tan, H.; Saidaminov, M.I.; Seok, S.I.; McGehee, M.D.; Sargent, E.H.; Han, H. Challenges for Commercializing Perovskite Solar Cells. Science 2018, 361, eaat8235. [Google Scholar] [CrossRef]
  113. Park, J.; Kim, J.; Yun, H.-S.; Paik, M.J.; Noh, E.; Mun, H.J.; Kim, M.G.; Shin, T.J.; Seok, S.I. Controlled Growth of Perovskite Layers with Volatile Alkylammonium Chlorides. Nature 2023, 616, 724–730. [Google Scholar] [CrossRef]
  114. Chavan, R.D.; Wolska-Pietkiewicz, M.; Prochowicz, D.; Jędrzejewska, M.; Tavakoli, M.M.; Yadav, P.; Hong, C.K.; Lewiński, J. Organic Ligand-Free ZnO Quantum Dots for Efficient and Stable Perovskite Solar Cells. Adv. Funct. Mater. 2022, 32, 2205909. [Google Scholar] [CrossRef]
  115. Sani, S.; Usman, A.; Bhatranand, A.; Jiraraksopakun, Y.; Muhammad, K.S.; Yahaya, U. A Study on Defect, Doping, and Performance of ETLs (ZnO, TiO2, and IGZO) for the Lead-Free CsSnCl3 Perovskite Solar Cell by SCAPS-1D Framework. Mater. Today Commun. 2024, 38, 107575. [Google Scholar] [CrossRef]
  116. Min, H.; Lee, D.Y.; Kim, J.; Kim, G.; Lee, K.S.; Kim, J.; Paik, M.J.; Kim, Y.K.; Kim, K.S.; Kim, M.G.; et al. Perovskite Solar Cells with Atomically Coherent Interlayers on SnO2 Electrodes. Nature 2021, 598, 444–450. [Google Scholar] [CrossRef] [PubMed]
  117. Anaraki, E.H.; Kermanpur, A.; Steier, L.; Domanski, K.; Matsui, T.; Tress, W.; Saliba, M.; Abate, A.; Grätzel, M.; Hagfeldt, A.; et al. Highly Efficient and Stable Planar Perovskite Solar Cells by Solution-Processed Tin Oxide. Energy Environ. Sci. 2016, 9, 3128–3134. [Google Scholar] [CrossRef]
  118. Dong, Q.; Li, J.; Shi, Y.; Chen, M.; Ono, L.K.; Zhou, K.; Zhang, C.; Qi, Y.; Zhou, Y.; Padture, N.P.; et al. Improved SnO2 Electron Transport Layers Solution-Deposited at Near Room Temperature for Rigid or Flexible Perovskite Solar Cells with High Efficiencies. Adv. Energy Mater. 2019, 9, 1900834. [Google Scholar] [CrossRef]
  119. Rombach, F.M.; Haque, S.A.; Macdonald, T.J. Lessons Learned from Spiro-OMeTAD and PTAA in Perovskite Solar Cells. Energy Environ. Sci. 2021, 14, 5161–5190. [Google Scholar] [CrossRef]
  120. Yekani, R.; Wang, H.; Ghamari, P.; Bessette, S.; Sharir-Smith, J.; Gauvin, R.; Demopoulos, G.P. Impact of Photo-Induced Doping of Spiro-OMeTAD as HTL on Perovskite Solar Cell Hysteresis Dynamics. J. Phys. Chem. C 2024, 128, 710–722. [Google Scholar] [CrossRef]
  121. Bi, H.; Fujiwara, Y.; Kapil, G.; Tavgeniene, D.; Zhang, Z.; Wang, L.; Ding, C.; Sahamir, S.R.; Baranwal, A.K.; Sanehira, Y.; et al. Perovskite Solar Cells Consisting of PTAA Modified with Monomolecular Layer and Application to All-Perovskite Tandem Solar Cells with Efficiency over 25%. Adv. Funct. Mater. 2023, 33, 2300089. [Google Scholar] [CrossRef]
  122. Li, W.; Martínez-Ferrero, E.; Palomares, E. Self-Assembled Molecules as Selective Contacts for Efficient and Stable Perovskite Solar Cells. Mater. Chem. Front. 2024, 8, 681–699. [Google Scholar] [CrossRef]
  123. Yi, Z.; Li, X.; Xiong, Y.; Shen, G.; Zhang, W.; Huang, Y.; Jiang, Q.; Ng, X.R.; Luo, Y.; Zheng, J.; et al. Self-assembled Monolayers (SAMs) in Inverted Perovskite Solar Cells and Their Tandem Photovoltaics Application. Interdiscip. Mater. 2024, 3, 203–244. [Google Scholar] [CrossRef]
  124. Chiang, C.-H.; Wu, C.-G. Bulk Heterojunction Perovskite–PCBM Solar Cells with High Fill Factor. Nat. Photonics 2016, 10, 196–200. [Google Scholar] [CrossRef]
  125. Gong, C.; Li, H.; Wang, H.; Zhang, C.; Zhuang, Q.; Wang, A.; Xu, Z.; Cai, W.; Li, R.; Li, X.; et al. Silver Coordination-Induced n-Doping of PCBM for Stable and Efficient Inverted Perovskite Solar Cells. Nat. Commun. 2024, 15, 4922. [Google Scholar] [CrossRef]
  126. Wojciechowski, K.; Leijtens, T.; Siprova, S.; Schlueter, C.; Hörantner, M.T.; Wang, J.T.-W.; Li, C.-Z.; Jen, A.K.-Y.; Lee, T.-L.; Snaith, H.J. C60 as an Efficient N-Type Compact Layer in Perovskite Solar Cells. J. Phys. Chem. Lett. 2015, 6, 2399–2405. [Google Scholar] [CrossRef] [PubMed]
  127. Said, A.A.; Aydin, E.; Ugur, E.; Xu, Z.; Deger, C.; Vishal, B.; Vlk, A.; Dally, P.; Yildirim, B.K.; Azmi, R.; et al. Sublimed C60 for Efficient and Repeatable Perovskite-Based Solar Cells. Nat. Commun. 2024, 15, 708. [Google Scholar] [CrossRef] [PubMed]
  128. Liu, S.; Biju, V.P.; Qi, Y.; Chen, W.; Liu, Z. Recent Progress in the Development of High-Efficiency Inverted Perovskite Solar Cells. NPG Asia Mater. 2023, 15, 27. [Google Scholar] [CrossRef]
  129. Tang, H.; Shen, Z.; Shen, Y.; Yan, G.; Wang, Y.; Han, Q.; Han, L. Reinforcing Self-Assembly of Hole Transport Molecules for Stable Inverted Perovskite Solar Cells. Science 2024, 383, 1236–1240. [Google Scholar] [CrossRef] [PubMed]
  130. Li, L.; Wei, M.; Carnevali, V.; Zeng, H.; Zeng, M.; Liu, R.; Lempesis, N.; Eickemeyer, F.T.; Luo, L.; Agosta, L.; et al. Buried-Interface Engineering Enables Efficient and 1960-Hour ISOS-L-2I Stable Inverted Perovskite Solar Cells. Adv. Mater. 2024, 36, 2303869. [Google Scholar] [CrossRef]
  131. Caprioglio, P.; Smith, J.A.; Oliver, R.D.J.; Dasgupta, A.; Choudhary, S.; Farrar, M.D.; Ramadan, A.J.; Lin, Y.-H.; Christoforo, M.G.; Ball, J.M.; et al. Open-Circuit and Short-Circuit Loss Management in Wide-Gap Perovskite p-i-n Solar Cells. Nat. Commun. 2023, 14, 932. [Google Scholar] [CrossRef]
  132. Imran, T.; Raza, H.; Aziz, L.; Chen, R.; Liu, S.; Jiang, Z.; Gao, Y.; Wang, J.; Younis, M.; Rauf, S.; et al. High Performance Inverted RbCsFAPbI3 Perovskite Solar Cells Based on Interface Engineering and Defects Passivation. Small 2023, 19, 2207950. [Google Scholar] [CrossRef]
  133. Cao, F.; Bian, L.; Li, L. Perovskite Solar Cells with High-Efficiency Exceeding 25%: A Review. Energy Mater. Devices 2024, 2, 9370018. [Google Scholar] [CrossRef]
  134. Shockley, W.; Queisser, H. Detailed Balance Limit of Efficiency of p–n Junction Solar Cells. Renew. Energy 2018, 2, 35–54. [Google Scholar]
  135. Jošt, M.; Kegelmann, L.; Korte, L.; Albrecht, S. Monolithic Perovskite Tandem Solar Cells: A Review of the Present Status and Advanced Characterization Methods Toward 30% Efficiency. Adv. Energy Mater. 2020, 10, 1904102. [Google Scholar] [CrossRef]
  136. Bryant, D.; Greenwood, P.; Troughton, J.; Wijdekop, M.; Carnie, M.; Davies, M.; Wojciechowski, K.; Snaith, H.J.; Watson, T.; Worsley, D. A Transparent Conductive Adhesive Laminate Electrode for High-Efficiency Organic-Inorganic Lead Halide Perovskite Solar Cells. Adv. Mater. 2014, 26, 7499–7504. [Google Scholar] [CrossRef] [PubMed]
  137. Berry, J.; Buonassisi, T.; Egger, D.A.; Hodes, G.; Kronik, L.; Loo, Y.; Lubomirsky, I.; Marder, S.R.; Mastai, Y.; Miller, J.S.; et al. Hybrid Organic–Inorganic Perovskites (HOIPs): Opportunities and Challenges. Adv. Mater. 2015, 27, 5102–5112. [Google Scholar] [CrossRef] [PubMed]
  138. Tockhorn, P.; Sutter, J.; Cruz, A.; Wagner, P.; Jäger, K.; Yoo, D.; Lang, F.; Grischek, M.; Li, B.; Li, J.; et al. Nano-Optical Designs for High-Efficiency Monolithic Perovskite–Silicon Tandem Solar Cells. Nat. Nanotechnol. 2022, 17, 1214–1221. [Google Scholar] [CrossRef] [PubMed]
  139. Liu, Z.; Li, H.; Chu, Z.; Xia, R.; Wen, J.; Mo, Y.; Zhu, H.; Luo, H.; Zheng, X.; Huang, Z.; et al. Reducing Perovskite/C60 Interface Losses via Sequential Interface Engineering for Efficient Perovskite/Silicon Tandem Solar Cell. Adv. Mater. 2024, 36, 2308370. [Google Scholar] [CrossRef]
  140. Zheng, X.; Kong, W.; Wen, J.; Hong, J.; Luo, H.; Xia, R.; Huang, Z.; Luo, X.; Liu, Z.; Li, H.; et al. Solvent Engineering for Scalable Fabrication of Perovskite/Silicon Tandem Solar Cells in Air. Nat. Commun. 2024, 15, 4907. [Google Scholar] [CrossRef]
  141. Priyanka, E.; Muchahary, D. Performance Improvement of Perovskite/CIGS Tandem Solar Cell Using Barium Stannate Charge Transport Layer and Achieving PCE of 39% Numerically. Sol. Energy 2024, 267, 112218. [Google Scholar] [CrossRef]
  142. Yu, D.; Pan, M.; Liu, G.; Jiang, X.; Wen, X.; Li, W.; Chen, S.; Zhou, W.; Wang, H.; Lu, Y.; et al. Electron-Withdrawing Organic Ligand for High-Efficiency All-Perovskite Tandem Solar Cells. Nat. Energy 2024, 9, 298–307. [Google Scholar] [CrossRef]
  143. Sun, H.; Xiao, K.; Gao, H.; Duan, C.; Zhao, S.; Wen, J.; Wang, Y.; Lin, R.; Zheng, X.; Luo, H.; et al. Scalable Solution-Processed Hybrid Electron Transport Layers for Efficient All-Perovskite Tandem Solar Modules. Adv. Mater. 2024, 36, 2308706. [Google Scholar] [CrossRef]
  144. Wei, R.; Breite, D.; Song, C.; Gräsing, D.; Ploss, T.; Hille, P.; Schwerdtfeger, R.; Matysik, J.; Schulze, A.; Zimmermann, W. Biocatalytic Degradation Efficiency of Postconsumer Polyethylene Terephthalate Packaging Determined by Their Polymer Microstructures. Adv. Sci. 2019, 6, 1900491. [Google Scholar] [CrossRef]
  145. Liu, J.; Zhao, Z.; Qian, J.; Liang, Z.; Wu, C.; Wang, K.; Liu, S.; Yang, D. Thermal Radiation Annealing for Overcoming Processing Temperature Limitation of Flexible Perovskite Solar Cells. Adv. Mater. 2024, 2401236. [Google Scholar] [CrossRef]
  146. Yang, D.; Yang, R.; Zhang, C.; Ye, T.; Wang, K.; Hou, Y.; Zheng, L.; Priya, S.; Liu, S. Highest-Efficiency Flexible Perovskite Solar Module by Interface Engineering for Efficient Charge-Transfer. Adv. Mater. 2023, 35, 2302484. [Google Scholar] [CrossRef] [PubMed]
  147. Ju, S.; Choi, S.J.; Sung, H.; Kim, M.; Song, J.W.; Choi, I.W.; Kim, H.-B.; Jo, Y.; Lee, S.; Yoon, S.-Y.; et al. High-Performance and Selective Semi-Transparent Perovskite Solar Cells Using 3D-Structured FTO. Renew. Energy 2024, 222, 119817. [Google Scholar] [CrossRef]
  148. Jin, Y.; Feng, H.; Fang, Z.; Yang, L.; Liu, K.; Deng, B.; Chen, J.; Chen, X.; Zhong, Y.; Yang, J.; et al. Stabilizing Semi-Transparent Perovskite Solar Cells with a Polymer Composite Hole Transport Layer. Nano Res. 2023, 17, 1500–1507. [Google Scholar] [CrossRef]
  149. Yoo, J.J.; Lee, J.-W.; Ju, Y.; Kang, B.J.; Kim, Y.; Kim, B.-S.; Kim, Y.Y.; Shin, S.S.; Shin, T.J.; Jeon, N.J. Pyrophosphate Interlayer Improves Performance of Semi-Transparent Perovskite Solar Cells. J. Mater. Chem. A 2024, 12, 12126–12133. [Google Scholar] [CrossRef]
  150. Tu, Y.; Wu, J.; Xu, G.; Yang, X.; Cai, R.; Gong, Q.; Zhu, R.; Huang, W. Perovskite Solar Cells for Space Applications: Progress and Challenges. Adv. Mater. 2021, 33, 2006545. [Google Scholar] [CrossRef]
  151. Ho-Baillie, A.W.Y.; Sullivan, H.G.J.; Bannerman, T.A.; Talathi, H.P.; Bing, J.; Tang, S.; Xu, A.; Bhattacharyya, D.; Cairns, I.H.; McKenzie, D.R. Deployment Opportunities for Space Photovoltaics and the Prospects for Perovskite Solar Cells. Adv. Mater. Technol. 2022, 7, 2101059. [Google Scholar] [CrossRef]
  152. Huan, Z.; Zheng, Y.; Wang, K.; Shen, Z.; Ni, W.; Zu, J.; Shao, Y. Advancements in Radiation Resistance and Reinforcement Strategies of Perovskite Solar Cells in Space Applications. J. Mater. Chem. A 2024, 12, 1910–1922. [Google Scholar] [CrossRef]
  153. Pérez-del-Rey, D.; Dreessen, C.; Igual-Muñoz, A.M.; Van Den Hengel, L.; Gélvez-Rueda, M.C.; Savenije, T.J.; Grozema, F.C.; Zimmermann, C.; Bolink, H.J. Perovskite Solar Cells: Stable under Space Conditions. Sol. RRL 2020, 4, 2000447. [Google Scholar] [CrossRef]
  154. Kong, L.; Liu, G. Synchrotron-Based Infrared Microspectroscopy under High Pressure: An Introduction. Matter Radiat. Extrem. 2021, 6, 068202. [Google Scholar] [CrossRef]
  155. Contreras, L.; Mayacela, M.; Bustillos, A.; Rentería, L.; Book, D. Hydrogen Sorption and Rehydrogenation Properties of NaMgH3. Metals 2022, 12, 205. [Google Scholar] [CrossRef]
  156. Wu, H.; Zhou, W.; Udovic, T.J.; Rush, J.J.; Yildirim, T. Crystal Chemistry of Perovskite-Type Hydride NaMgH3: Implications for Hydrogen Storage. Chem. Mater. 2008, 20, 2335–2342. [Google Scholar] [CrossRef]
  157. Reshak, A.H.; Shalaginov, M.Y.; Saeed, Y.; Kityk, I.V.; Auluck, S. First-Principles Calculations of Structural, Elastic, Electronic, and Optical Properties of Perovskite-Type KMgH3 Crystals: Novel Hydrogen Storage Material. J. Phys. Chem. B 2011, 115, 2836–2841. [Google Scholar] [CrossRef] [PubMed]
  158. Sato, T.; Noréus, D.; Takeshita, H.; Häussermann, U. Hydrides with the Perovskite Structure: General Bonding and Stability Considerations and the New Representative CaNiH3. J. Solid State Chem. 2005, 178, 3381–3388. [Google Scholar] [CrossRef]
  159. Naskar, A.; Khanal, R.; Choudhury, S. Role of Chemistry and Crystal Structure on the Electronic Defect States in Cs-Based Halide Perovskites. Materials 2021, 14, 1032. [Google Scholar] [CrossRef] [PubMed]
  160. Mbonu, I.J.; Louis, H.; Chukwu, U.G.; Agwamba, E.C.; Ghotekar, S.; Adeyinka, A.S. Effects of Metals (X = Be, Mg, Ca) Encapsulation on the Structural, Electronic, Phonon, and Hydrogen Storage Properties of KXCl3 Halide Perovskites: Perspective from Density Functional Theory. Int. J. Hydrogen Energy 2024, 50, 337–351. [Google Scholar] [CrossRef]
  161. Song, F.; Bai, L.; Moysiadou, A.; Lee, S.; Hu, C.; Liardet, L.; Hu, X. Transition Metal Oxides as Electrocatalysts for the Oxygen Evolution Reaction in Alkaline Solutions: An Application-Inspired Renaissance. J. Am. Chem. Soc. 2018, 140, 7748–7759. [Google Scholar] [CrossRef]
  162. Deng, G.; Chen, Y.; Tao, M.; Wu, C.; Shen, X.; Yang, H.; Liu, M. Electrochemical Properties and Hydrogen Storage Mechanism of Perovskite-Type Oxide LaFeO3 as a Negative Electrode for Ni/MH Batteries. Electrochim. Acta 2010, 55, 1120–1124. [Google Scholar] [CrossRef]
  163. Ostadebrahim, M.; Moradlou, O. Electrochemical Hydrogen Storage in LaMO3 (M = Cr, Mn, Fe, Co, Ni) Nano-Perovskites. J. Energy Storage 2023, 72, 108284. [Google Scholar] [CrossRef]
  164. Mohassel, R.; Soofivand, F.; Dawi, E.A.; Shabani-Nooshabadi, M.; Salavati-Niasari, M. ErMnO3/Er2Mn2O7/ZnO/GO Multi-Component Nanocomposite as a Promising Material for Hydrogen Storage: Facile Synthesis and Comprehensive Investigation of Component Roles. J. Energy Storage 2023, 65, 107285. [Google Scholar] [CrossRef]
  165. Esaka, T.; Sakaguchi, H.; Kobayashi, S. Hydrogen Storage in Proton-Conductive Perovskite-Type Oxides and Their Application to Nickel–Hydrogen Batteries. Solid State Ion. 2004, 166, 351–357. [Google Scholar] [CrossRef]
  166. Sakaguchi, H.; Hatakeyama, K.; Kobayashi, S.; Esaka, T. Hydrogenation Characteristics of the Proton Conducting Oxide–Hydrogen Storage Alloy Composite. Mater. Res. Bull. 2002, 37, 1547–1556. [Google Scholar] [CrossRef]
  167. Henao, J.; Sotelo, O.; Casales-Diaz, M.; Martinez-Gomez, L. Hydrogen Storage in a Rare-Earth Perovskite-Type Oxide La0.6Sr0.4Co0.2Fe0.8O3 for Battery Applications. Rare Met. 2018, 37, 1003–1013. [Google Scholar] [CrossRef]
  168. Rehman, Z.U.; Rehman, M.A.; Rehman, B.; Sikiru, S.; Qureshi, S.; Ali, E.M.; Awais, M.; Amjad, M.; Iqbal, I.; Rafique, A.; et al. Ab Initio Insight into the Physical Properties of MgXH3 (X = Co, Cu, Ni) Lead-Free Perovskite for Hydrogen Storage Application. Environ. Sci. Pollut. Res. 2023, 30, 113889–113902. [Google Scholar] [CrossRef] [PubMed]
  169. Mera, A.; Rehman, M.A. First-Principles Investigation for the Hydrogen Storage Properties of AeSiH3 (Ae = Li, K, Na, Mg) Perovskite-Type Hydrides. Int. J. Hydrogen Energy 2024, 50, 1435–1447. [Google Scholar] [CrossRef]
  170. Siddique, A.; Khalil, A.; Almutairi, B.S.; Tahir, M.B.; Sagir, M.; Ullah, Z.; Hannan, A.; Ali, H.E.; Alrobei, H.; Alzaid, M. Structures and Hydrogen Storage Properties of AeVH3 (Ae = Be, Mg, Ca, Sr) Perovskite Hydrides by DFT Calculations. Int. J. Hydrogen Energy 2023, 48, 24401–24411. [Google Scholar] [CrossRef]
  171. Hayat, S.; Khalil, R.M.A.; Hussain, M.I.; Rana, A.M.; Hussain, F. First-Principles Investigations of the Structural, Optoelectronic, Magnetic and Thermodynamic Properties of Hydride Perovskites XCuH3 (X = Co, Ni, Zn) for Hydrogen Storage Applications. Optik 2021, 228, 166187. [Google Scholar] [CrossRef]
  172. Rafique, A.; Usman, M.; Rehman, J.U.; Nazeer, A.; Ullah, H.; Hussain, A. Investigation of Structural, Electronic, Mechanical, Optical and Hydrogen Storage Properties of Cobalt-Based Hydride-Perovskites XCoH3 (X = In, Mn, Sr, Sn, Cd) for Hydrogen Storage Application. J. Phys. Chem. Solids 2023, 181, 111559. [Google Scholar] [CrossRef]
  173. Song, R.; Chen, Y.; Chen, S.; Xu, N.; Zhang, W. First-Principles to Explore the Hydrogen Storage Properties of XPtH3 (X = Li, Na, K, Rb) Perovskite Type Hydrides. Int. J. Hydrogen Energy 2024, 57, 949–957. [Google Scholar] [CrossRef]
  174. Xu, N.; Song, R.; Zhang, J.; Chen, Y.; Chen, S.; Li, S.; Jiang, Z.; Zhang, W. First-Principles Study on Hydrogen Storage Properties of the New Hydride Perovskite XAlH3 (X = Na, K). Int. J. Hydrogen Energy 2024, 60, 434–440. [Google Scholar] [CrossRef]
  175. Xie, D.; Zhang, M.; Liu, Q.; Lin, Y.; Yu, A.; Tang, Y. Organic-Inorganic Conformal Extending High-Purity Metal Nanosheets for Robust Electrochemical Lithium-Ion Storage. Adv. Funct. Mater. 2023, 33, 2306291. [Google Scholar] [CrossRef]
  176. Shah, S.F.A.; Murtaza, G.; Ismail, K.; Raza, H.H.; Khan, I.J. First Principles Investigation of Transition Metal Hydrides LiXH3 (X = Ti, Mn, and Cu) for Hydrogen Storage. J. Comput. Electron. 2023, 22, 921–929. [Google Scholar] [CrossRef]
  177. Surucu, G.; Candan, A.; Gencer, A.; Isik, M. First-Principle Investigation for the Hydrogen Storage Properties of NaXH3 (X = Mn, Fe, Co) Perovskite Type Hydrides. Int. J. Hydrogen Energy 2019, 44, 30218–30225. [Google Scholar] [CrossRef]
  178. Bhardwaj, A.; Bae, H.; Kim, I.-H.; Mathur, L.; Park, J.-Y.; Song, S.-J. High Capacity, Rate-Capability, and Power Delivery at High-Temperature by an Oxygen-Deficient Perovskite Oxide as Proton Insertion Anodes for Energy Storage Devices. J. Electrochem. Soc. 2021, 168, 070540. [Google Scholar] [CrossRef]
  179. Lahlou Nabil, M.A.; Fenineche, N.; Popa, I.; Sunyol, J.J. Morphological, Structural and Hydrogen Storage Properties of LaCrO3 Perovskite-Type Oxides. Energies 2022, 15, 1463. [Google Scholar] [CrossRef]
  180. Ullah, M.A.; Riaz, K.N.; Rizwan, M. Computational Evaluation of KMgO3-xHx as an Efficient Hydrogen Storage Material. J. Energy Storage 2023, 70, 108030. [Google Scholar] [CrossRef]
  181. Parvazian, E.; Watson, T. The Roll-to-Roll Revolution to Tackle the Industrial Leap for Perovskite Solar Cells. Nat. Commun. 2024, 15, 3983. [Google Scholar] [CrossRef]
  182. Chen, C.; Ran, C.; Yao, Q.; Wang, J.; Guo, C.; Gu, L.; Han, H.; Wang, X.; Chao, L.; Xia, Y.; et al. Screen-Printing Technology for Scale Manufacturing of Perovskite Solar Cells. Adv. Sci. 2023, 10, 2303992. [Google Scholar] [CrossRef]
  183. Khatoon, S.; Kumar Yadav, S.; Chakravorty, V.; Singh, J.; Bahadur Singh, R.; Hasnain, M.S.; Hasnain, S.M.M. Perovskite Solar Cell’s Efficiency, Stability and Scalability: A Review. Mater. Sci. Energy Technol. 2023, 6, 437–459. [Google Scholar] [CrossRef]
  184. Noman, M.; Khan, Z.; Jan, S.T. A Comprehensive Review on the Advancements and Challenges in Perovskite Solar Cell Technology. RSC Adv. 2024, 14, 5085–5131. [Google Scholar] [CrossRef]
  185. Ur Rehman, Z.; Rehman, M.A.; Alomar, S.Y.; Rehman, B.; Awais, M.; Amjad, M.; Sikiru, S.; Ali, E.M.; Hamad, A. Hydrogen Storage Capacity of Lead-Free Perovskite NaMTH3 (MT = Sc, Ti, V): A DFT Study. Int. J. Energy Res. 2024, 2024, 4009198. [Google Scholar] [CrossRef]
Figure 1. Crystal structure of hybrid organic-inorganic perovskite materials.
Figure 1. Crystal structure of hybrid organic-inorganic perovskite materials.
Energies 17 04755 g001
Figure 2. Band diagram and main process and PSC: (1) Absorption of photon and free charges generation; (2) Charge transport; (3) Charge extraction. Here, HTL stands for the hole transport layer, FTO for fluorine-doped tin oxide, and ETL is the electron transport layer. Results are reproduced from [109] with permission from Elsevier.
Figure 2. Band diagram and main process and PSC: (1) Absorption of photon and free charges generation; (2) Charge transport; (3) Charge extraction. Here, HTL stands for the hole transport layer, FTO for fluorine-doped tin oxide, and ETL is the electron transport layer. Results are reproduced from [109] with permission from Elsevier.
Energies 17 04755 g002
Figure 3. Four device configurations of PSCs: mesoscopic structure, planar structure, triple mesoscopic structure, and tandem structure with lower-bandgap subcell. Here, TCO is the transparent conductive oxide, and ETL is the electron transport layer. Present results are reproduced from [112] with permission from AAAS.
Figure 3. Four device configurations of PSCs: mesoscopic structure, planar structure, triple mesoscopic structure, and tandem structure with lower-bandgap subcell. Here, TCO is the transparent conductive oxide, and ETL is the electron transport layer. Present results are reproduced from [112] with permission from AAAS.
Energies 17 04755 g003
Figure 4. The crystal structure of MgXH3 for X = Co, X = Cu, X = Ni.
Figure 4. The crystal structure of MgXH3 for X = Co, X = Cu, X = Ni.
Energies 17 04755 g004
Figure 5. The crystal structures of AeSiH3 for Ae = K, Ae = Li, Ae = Na, Ae = Mg.
Figure 5. The crystal structures of AeSiH3 for Ae = K, Ae = Li, Ae = Na, Ae = Mg.
Energies 17 04755 g005
Table 1. The calculated lattice constant (a = b = c in Å), volume (V in Å3), density (ρ in g/cm3), formation enthalpy (ΔH in kJ/mol·H2), gravimetric hydrogen storage capacity (Cwt% in wt%), and desorption temperature (Tdes in K) for MgXH3, where X = Co, Cu, Ni, respectively. Results are reproduced from [168] with permission from Springer Nature.
Table 1. The calculated lattice constant (a = b = c in Å), volume (V in Å3), density (ρ in g/cm3), formation enthalpy (ΔH in kJ/mol·H2), gravimetric hydrogen storage capacity (Cwt% in wt%), and desorption temperature (Tdes in K) for MgXH3, where X = Co, Cu, Ni, respectively. Results are reproduced from [168] with permission from Springer Nature.
Compounds C w t % a 0 · (Å) V · (Å3) ρ · (g/cm3) Δ H f · (eV/Atom) T d e s · (K)
MgCoH33.643.3236.443.93−70.93542.69
MgCuH33.323.4942.423.56−63.27484.08
MgNiH33.493.3637.973.76−68.54524.40
Table 2. The optimized lattice constants (a0), volumes (V), densities (ρ), formation enthalpies (ΔHf), and gravimetric hydrogen storage capacities (Cwt%) of CaXH3 (X: Mn, Fe, or Co) compounds. Results are reproduced from [47] with permission from Wiley.
Table 2. The optimized lattice constants (a0), volumes (V), densities (ρ), formation enthalpies (ΔHf), and gravimetric hydrogen storage capacities (Cwt%) of CaXH3 (X: Mn, Fe, or Co) compounds. Results are reproduced from [47] with permission from Wiley.
Compounds C w t % a 0 · (Å) V · (Å3) ρ · (g/cm3) Δ H f · (eV/Atom)
CaMnH33.093.6046.583.50−0.25
CaFeH33.063.5042.993.82−0.42
CaCoH32.973.4842.164.03−0.44
Table 3. The optimized lattice constants (a0), volumes (V), densities (ρ), formation energy (ΔHf), and gravimetric hydrogen storage capacities (Cwt%) of AeSiH3 (Ae = Li, K, Na, Mg) compounds. Results are reproduced from [169] with permission from Elsevier.
Table 3. The optimized lattice constants (a0), volumes (V), densities (ρ), formation energy (ΔHf), and gravimetric hydrogen storage capacities (Cwt%) of AeSiH3 (Ae = Li, K, Na, Mg) compounds. Results are reproduced from [169] with permission from Elsevier.
Compounds C w t % a 0 · (Å) V · (Å3) ρ · (g/cm3) Δ H f · (eV/Atom)
LiSiH37.9464.00164.0791.046−17.372
KSiH34.3063.91760.1221.952−15.760
NaSiH35.5883.98663.3631.429−16.134
MgSiH35.4563.97762.9331.476−15.063
Table 4. The optimized lattice constants (a0), volumes (V), formation energy (ΔHf), and gravimetric hydrogen storage capacities (Cwt%) of XSrH3 (X = K and Rb) compounds. Results are reproduced from [73] with permission from Wiley.
Table 4. The optimized lattice constants (a0), volumes (V), formation energy (ΔHf), and gravimetric hydrogen storage capacities (Cwt%) of XSrH3 (X = K and Rb) compounds. Results are reproduced from [73] with permission from Wiley.
Compounds C w t % a 0 · (Å) V · (Å3) Δ H f · (eV/Atom)
KSrH32.334.77108.50−6.60
RbSrH31.714.99124.77−5.67
Table 5. The optimized lattice constants (a0), volumes (V), free energy (F), and gravimetric hydrogen storage capacities (Cwt%) of XCuH3 (X = Co, Ni, Zn) compounds. Results are reproduced from [171] with permission from Elsevier.
Table 5. The optimized lattice constants (a0), volumes (V), free energy (F), and gravimetric hydrogen storage capacities (Cwt%) of XCuH3 (X = Co, Ni, Zn) compounds. Results are reproduced from [171] with permission from Elsevier.
Compounds C w t % a 0 · (Å) V · (Å3) F
CoCuH32.83.328736.882−1895.3
NiCuH33.03.324536.742−5499.0
ZnCuH32.73.612947.160−2512.5
Table 6. The optimized lattice constants (a0), volumes (V), formation energy (ΔHf), and gravimetric hydrogen storage capacities (Cwt%) of XCoH3 (X = In, Mn, Sr, Sn, Cd) compounds. Results are reproduced from [172] with permission from Elsevier.
Table 6. The optimized lattice constants (a0), volumes (V), formation energy (ΔHf), and gravimetric hydrogen storage capacities (Cwt%) of XCoH3 (X = In, Mn, Sr, Sn, Cd) compounds. Results are reproduced from [172] with permission from Elsevier.
Compounds C w t % a 0 · (Å) V · (Å3) Δ H f · (eV/Atom)
CdCoH31.743.4220.00−0.93
InCoH31.713.5243.61−1.09
MnCoH32.593.6046.66−0.87
SnCoH31.683.5946.27−1.31
SrCoH32.033.6649.03−0.78
Table 7. The calculated lattice constant (a0), volume (V), formation energy (ΔHf), gravimetric hydrogen storage capacity (Cwt%), and desorption temperature (Tdes) for XPtH3 (X = Li, Na, K, Rb). Results are reproduced from [173] with permission from Elsevier.
Table 7. The calculated lattice constant (a0), volume (V), formation energy (ΔHf), gravimetric hydrogen storage capacity (Cwt%), and desorption temperature (Tdes) for XPtH3 (X = Li, Na, K, Rb). Results are reproduced from [173] with permission from Elsevier.
Compounds C w t % a 0 · (Å) V · (Å3) Δ H f · (eV/Atom) T d e s · (K)
LiPtH31.453.5444.35−0.32237.77
NaPtH31.353.6347.96−0.31225.39
KPtH31.263.8054.83−0.22162.43
RbPtH31.063.9059.25−0.1180.11
Table 8. The optimized lattice constants (a0), volumes (V), formation energy (ΔHf), and gravimetric hydrogen storage capacities (Cwt%) of XAlH3 (X = Na, K) compounds. Results are reproduced from [174] with permission from Elsevier.
Table 8. The optimized lattice constants (a0), volumes (V), formation energy (ΔHf), and gravimetric hydrogen storage capacities (Cwt%) of XAlH3 (X = Na, K) compounds. Results are reproduced from [174] with permission from Elsevier.
Compounds C w t % a 0 · (Å) V · (Å3) Δ H f · (eV/Atom)
NaAlH35.403.79254.526−0.903
KAlH34.193.93861.070−1.250
Table 9. The optimized lattice constants (a0), bulk modulus B0 (GPa), the derivative of bulk modulus B0’ (GPa), formation energy (ΔHf), and gravimetric hydrogen storage capacities (Cwt%) of XTiH3 (X = K, Rb, Cs) compounds. Results are reproduced from [77] with permission from Elsevier.
Table 9. The optimized lattice constants (a0), bulk modulus B0 (GPa), the derivative of bulk modulus B0’ (GPa), formation energy (ΔHf), and gravimetric hydrogen storage capacities (Cwt%) of XTiH3 (X = K, Rb, Cs) compounds. Results are reproduced from [77] with permission from Elsevier.
Compounds C w t % a 0 · (Å) B 0 · (GPa) B 0 · (GPa) Δ H f · (eV/Atom)
KTiH33.363.99944.9381.767−0.285
RbTiH32.224.10342.0831.887−0.220
CsTuH31.654.23337.1761.704−0.146
Table 10. The calculated lattice constant (a = b = c in Å), volume (V in Å3), bulk modulus B0 (GPa), formation enthalpy ( Δ H f in kJ/mol·H2), and gravimetric hydrogen storage capacity (Cwt% in wt%) for KXH3 (X = Be, Ca, Mg). Results are reproduced from [160] with permission from Elsevier.
Table 10. The calculated lattice constant (a = b = c in Å), volume (V in Å3), bulk modulus B0 (GPa), formation enthalpy ( Δ H f in kJ/mol·H2), and gravimetric hydrogen storage capacity (Cwt% in wt%) for KXH3 (X = Be, Ca, Mg). Results are reproduced from [160] with permission from Elsevier.
Compounds C w t % a 0 · (Å) V · (Å3) B 0 · (GPa) Δ H f · (eV/Atom)
KBeH35.8664.41158.3852.650−1.226
KMgH34.5164.74297.5380.916−2.523
KCaH33.6495.1089.71927.902−1.845
Table 11. BET and BJH data of LaMO3 (M = Cr, Mn, Fe, Co, Ni) nanocrystals. Results are reproduced from [163] with permission from Elsevier.
Table 11. BET and BJH data of LaMO3 (M = Cr, Mn, Fe, Co, Ni) nanocrystals. Results are reproduced from [163] with permission from Elsevier.
Samples S A B E T ( m 2 / g ) V t ( c m 2 / g ) D a v g · (nm) S B J H ( m 2 / g )
LaCrO31.48380.004412.0605.8201
LaMnO38.87780.030213.6166.7078
LaFeO320.9970.120110.72120.698
LaCoO38.20640.069333.80710.929
LaNiO34.85190.056298.99816.537
Table 12. Electrochemical hydrogen storage capacity of LaMO3 (M = Cr, Mn, Fe, Co, Ni) nanocrystals. Results are reproduced from [163] with permission from Elsevier.
Table 12. Electrochemical hydrogen storage capacity of LaMO3 (M = Cr, Mn, Fe, Co, Ni) nanocrystals. Results are reproduced from [163] with permission from Elsevier.
Active
Material
SubstrateElectrolyte
Solution
Reference
Electrode
Counter
Electrode
Discharge
Capacity
LaCrO3----6790 mAh/g
LaMnO3----10,500 mAh/g
LaFeO3Cu electrode6M KOHAg/AGClPt13,500 mAh/g
LaCoO3----8800 mAh/g
LaNiO3----7000 mAh/g
Table 13. Cell parameters, mechanical properties of LaCrO3 and LaCrO3H6 [179].
Table 13. Cell parameters, mechanical properties of LaCrO3 and LaCrO3H6 [179].
CompoundsCalculated Cell Parameters (A) B 0 · (GPa) B 0 · (GPa)
LaCrO33.85128.073.2
LaCrO3H64.4362.2623.48
Table 14. The calculated lattice constant (a = b = c in Å), volume (V in Å3), formation enthalpy (ΔHf in kJ/mol·H2), gravimetric hydrogen storage capacity (Cwt% in wt%), and desorption temperature (Tdes in K) for KMO3−xHx (x = 0, 0.3, 0.6, 0.9, 1.2, 1.5, 1.8, 2.1, 2.4, 2.7, and 3.0). Results are reproduced from [180] with permission from Elsevier.
Table 14. The calculated lattice constant (a = b = c in Å), volume (V in Å3), formation enthalpy (ΔHf in kJ/mol·H2), gravimetric hydrogen storage capacity (Cwt% in wt%), and desorption temperature (Tdes in K) for KMO3−xHx (x = 0, 0.3, 0.6, 0.9, 1.2, 1.5, 1.8, 2.1, 2.4, 2.7, and 3.0). Results are reproduced from [180] with permission from Elsevier.
Compounds C w t %
a
a 0 · (Å)
b

c
V · (Å3) Δ H f · (eV/Atom) T d e s · (K)
KMgO3-4.109--69.37−0.023-
KMg2.7O0.30.284.1024.1024.11269.19−11.122825
KMg2.4O0.60.594.0924.0674.14068.90−10.826803
KMg2.1O0.90.924.0594.0574.16268.54−10.529781
KMg1.8O1.21.284.0344.0384.19668.35−10.223759
KMg1.5O1.51.674.1073.9894.18268.51−9.937737
KMg1.2O1.82.104.5104.6534.43792.90−9.648716
KMg0.9O2.12.584.4503.9754.10872.60−9.347694
KMg0.6O2.43.104.1563.9664.09167.43−8.624640
KMg0.3O2.73.694.0584.0844.00466.35−7.904587
KMgH34.354.069--67.36−7.606564
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kuo, M.-H.; Neykova, N.; Stachiv, I. Overview of the Recent Findings in the Perovskite-Type Structures Used for Solar Cells and Hydrogen Storage. Energies 2024, 17, 4755. https://doi.org/10.3390/en17184755

AMA Style

Kuo M-H, Neykova N, Stachiv I. Overview of the Recent Findings in the Perovskite-Type Structures Used for Solar Cells and Hydrogen Storage. Energies. 2024; 17(18):4755. https://doi.org/10.3390/en17184755

Chicago/Turabian Style

Kuo, Meng-Hsueh, Neda Neykova, and Ivo Stachiv. 2024. "Overview of the Recent Findings in the Perovskite-Type Structures Used for Solar Cells and Hydrogen Storage" Energies 17, no. 18: 4755. https://doi.org/10.3390/en17184755

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop