Next Article in Journal
The Characteristics of Light (TiCrAl0.5NbCu)CxNy High-Entropy Coatings Deposited Using a HiPIMS/DCMS Technique
Previous Article in Journal
Study of the Structural-Phase State of Hydroxyapatite Coatings Obtained by Detonation Spraying at Different O2/C2H2 Ratios
Previous Article in Special Issue
Oxidized Graphitic-C3N4 with an Extended π-System for Enhanced Photoelectrochemical Property and Behavior
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hybrid Perovskites and 2D Materials in Optoelectronic and Photocatalytic Applications

by
Shuo Feng
,
Benxuan Li
,
Bo Xu
and
Zhuo Wang
*
International Collaborative Laboratory of 2D Materials for Optoelectronics Science and Technology of Ministry of Education, Institute of Microscale Optoelectronics, Shenzhen University, Shenzhen 518060, China
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(11), 1566; https://doi.org/10.3390/cryst13111566
Submission received: 29 August 2023 / Revised: 17 September 2023 / Accepted: 22 September 2023 / Published: 2 November 2023

Abstract

:
Metal halide perovskites, emerging innovative and promising semiconductor materials with notable properties, have been a great success in the optoelectronic and photocatalytic fields. At the same time, two-dimensional (2D) materials, including graphene, transition metal dichalcogenides (TMDCs), black phosphorus (BP) and so on, have attracted significant interest due to their remarkable attributes. While substantial advancements have been made in recent decades, there are still hurdles in enhancing the performance of devices made from perovskites or 2D materials and in addressing their stability for reliable use. Recently, heterostructures combining perovskites with cost-effective 2D materials have exhibited significant advancements in both efficiency and stability, attributed to the unique properties at the heterointerface. In this review, we provide a thorough overview of perovskite and 2D material heterostructures, spanning from synthesis to application. We begin by detailing the diverse fabrication techniques, categorizing them into solid-state and solution-processed methods. Subsequently, we delve into the applications of perovskite and 2D material heterostructures, elaborating on their use in photodetectors, solar cells, and photocatalysis. We conclude by spotlighting existing challenges in developing perovskite and 2D material heterostructures and suggesting potential avenues for further advancements in this research area.

1. Introduction

Metal halide perovskites have surged to the forefront of research over the past decade. The cutting-edge optoelectronic devices built on these perovskites, such as photodetectors (PDs), light-emitting diodes (LEDs), perovskite solar cells (PSCs), are setting new benchmarks and leading current technological trends due to the intriguing properties of perovskites [1,2,3,4]. Since the birth of the first perovskite-sensitized solar cell with a power conversion efficiency (PCE) of 3.8% in 2009 [5], the best PSC exhibited 25.6% certified power efficiency, with 80% of that efficiency after 500 h of operation at 85 °C [6]. This exceptional device performance is primarily due to the unique features of perovskites, including superior optical absorption coefficient, high carrier mobility, long carrier diffusion length and slow carrier recombination rate [7,8,9]. Additionally, the solubility of perovskite materials offers a promising avenue for producing cost-effective, large-scale solar cell modules, giving them a competitive edge over conventional silicon-based alternatives. While the impressive physical attributes linked with their structural and functional versatility have propelled perovskite materials to success in optoelectronic fields, there are still hurdles to overcome in enhancing their efficiency and addressing stability concerns for upcoming commercialization [10]. Hybridizing perovskites with other functional materials, e.g., two-dimensional (2D) materials, has proven to be one promising way to achieve device performance and higher stability [11,12].
Two-dimensional materials such as graphene, transition metal dichalcogenides (TMDCs), black phosphorus (BP), and transition-metal carbides and nitrides (MXenes) have garnered extensive attention because of their exceptional qualities derived from distinct van der Waals (vdWs) configurations [13,14,15,16,17,18]. Unlike traditional bulk structures, where 2D materials are composed of distinct layers held together by strong in-plane covalent bonds, atomic-scale 2D materials manifest distinct and remarkable optical, electrical, thermal, chemical, and mechanical characteristics [19]. Moreover, the distinctive 2D nature of these materials allows for easy conversion into consistently thin films with highly aligned microstructures, making them ideal for optoelectronic applications. However, due to their atomically thin nature, 2D materials typically face challenges with limited light absorption [20].
Recently, the fusion of perovskites and 2D materials into heterostructures, blending two leading research areas, has captivated interest in multiple fields [21]. Integrating perovskites with 2D materials not only boosts the light absorption capacity of the extremely thin 2D materials, leveraging the exceptional optical properties of perovskites, but also strengthens the stability of perovskites, especially when 2D nanosheets act as protective encapsulation layers [22]. Additionally, the perovskite and 2D material heterostructures display distinctive characteristics such as ultrafast charge and energy transfer [23], photogating effect [24], strong interlayer coupling [25], and so on, which result in significantly improved optoelectronic and photoelectric properties compared to their components. More importantly, the vast array of perovskites and 2D materials offer excellent opportunities to design and tailor the compositions and functionalities of perovskite and 2D material heterostructures.
In this review, different fabrication techniques for perovskite and 2D material heterostructures are discussed, classifying them into solid-state approaches and solution-based processes. Then, a detailed overview of their applications in photodetectors, solar cells and photocatalysis is provided. Lastly, we discuss the obstacles to creating high-quality, stable, eco-friendly perovskite and 2D material heterostructures for broad-scale use. This review intends to offer researchers a structured update on novel fabrication methods and recent applications of perovskite and 2D material heterostructures, thereby advancing the progress of perovskite and 2D material heterostructures.

2. Fabrication Techniques of Perovskite and 2D Material Heterostructures

Perovskite and 2D material heterostructures can be prepared using either solid-state or solution techniques. Typically, solid-state approaches involve the use of vapor deposition to form these heterostructures, followed by dry-wet transfer onto designated substrates. The solid-state method is suitable for fabricating individual devices, studying the effects of the composition, thickness, etc., of different layers of the device on the device, and producing novel heterostructures for foundational studies. On the other hand, solution techniques, such as spin coating, ex situ hybridization and in situ growth, are acknowledged for their simplicity and cost-effectiveness, making them ideal for the large-scale production of these heterostructures. Moreover, using low-temperature processing in these solution methods can minimize energy use during fabrication and can lead to devices with mechanical flexibility. In this section, we summarize the techniques related to solid-state and solution methods for the fabrication of perovskite and 2D material heterostructures.

2.1. Solid-State Methods

Vapor deposition stands out as a prominent solid-state technique in creating perovskite and 2D material heterostructures. In this method, materials or volatile precursors are initially vaporized at elevated temperatures. Subsequently, this vapor phase is condensed onto a substrate, resulting in the desired layered structure where high-quality single-crystal and continuous thin films can be formed. The typical fabrication process for perovskite and 2D material heterostructures through vapor deposition involves initially laying down consistent and flat 2D material flakes using the chemical vapor deposition (CVD) approach, which is followed by a one- or two-step of vapor deposition of a perovskite layer. Niu and colleagues pioneered the use of the vapor deposition technique to create three distinct heterostructures: perovskite/hexagonal boron nitride (h-BN), perovskite/MoS2, and perovskite/graphene [26]. Initially, they laid down 2D monolayers on SiO2/Si platforms using CVD, followed by layering PbI2 nanosheets onto the 2D monolayer via physical vapor deposition (PVD). In the final stage, they transformed the PbI2 on the 2D monolayer into CH3NH3PbI3 by facilitating a reaction with the vapor of methylammonium iodide. The scanning electron microscope (SEM) images in Figure 1a–c show that the perovskite crystals are located on the 2D material monolayer. The surface roughness ≈ 2.0 nm indicates the high quality of the perovskite and 2D material heterostructure. Intriguingly, the use of photolithography allows us to produce these heterostructures in varying shapes, catering to diverse requirements, and high-quality perovskite and 2D material heterostructures were successfully constructed via vapor deposition, facilitating their application in optoelectronics. Furthermore, heterostructures produced via the vapor deposition technique displayed enhanced photodetection capabilities. A notable case is the work by Lu et al., where they applied focused laser annealing to CVD-generated WSe2 monolayers during the fabrication of CH3NH3PbI3/WSe2 heterostructures [27]. This post-treatment procedure allowed chalcogenide gaps in WSe2 to be filled through oxygen substitution, boosting the conductivity of WSe2. Consequently, the resultant PD demonstrated remarkably high photoresponsivity and quantum efficiency.
The dry/wet transfer method presents an alternative solid-state approach for creating perovskite and 2D material heterostructures, which involves transitioning a sample from one substrate to another, utilizing either a solid or liquid medium. The widely used transfer techniques are based on polymer thin films such as polymethyl methacrylate (PMMA) [29]. Due to the adhesive force between the polymer film and the sample, it is possible to detach the sample and relocate it onto a designated substrate in a specific order. Once transferred, the polymer film can be dissolved using typical organic solvents, placing the sample neatly on the target substrate. Using several cycles of the transfer process, a layered heterostructure can be constructed. In 2016, graphene/perovskite/graphene and graphene/WSe2/perovskite/graphene heterojunctions were prepared for the first time via multiple dry/wet transfers [30]. For graphene/perovskite/graphene devices, the bottom graphene is first exfoliated onto an SiO2/Si substrate, and the PbI2 is then mechanically exfoliated and dry-transferred on top of graphene through a polymethyl methacrylate/polypropylene carbonate (PMMA/PPC). Then, PbI2 is converted to CH3NH3PbI3 after reaction with MAI steam, and the obtained CH3NH3PbI3 perovskite is immediately covered by another graphene flake via PMMA. The fabrication process for graphene/WSe2/perovskite/graphene devices parallels that of graphene/perovskite/graphene devices, with a distinction: an added dry transfer phase wherein WSe2 flakes are layered onto the primary graphene monolayer using a PMMA/PPC stack. Additionally, to bolster protection for the device, an extra boron nitride (BN) layer can be introduced through dry-transfer. These multi-layered heterostructures prepared using solid-state techniques such as vapor deposition and dry-wet transfer are surface-cleaned, positioning them as promising platforms for delving into new physical, optical, and electronic characteristics. Nonetheless, these solid-state approaches come with inherent challenges. The intricacies of the fabrication procedures, which might encompass techniques such as CVD, PVD, mechanical extraction, numerous wet and dry transfer phases, and precise alignment, are considerable. Coupled with their energy-intensive prerequisites, these factors constrain the spectrum of applicable perovskites and 2D materials, and they also curb the mass production of high-grade structures.

2.2. Solution Methods

Spin coating stands as a facile technique favored in fabricating optoelectronics. Notably cost-effective and adaptable, it aligns seamlessly with a broad spectrum of 2D materials when paired with perovskites. By spin-coating perovskite and 2D material precursor solutions onto designated substrates, layered structures of perovskites and 2D materials can be constructed. The spin-coating approach paves the way for mass-producing these heterostructures economically. Three-dimensional perovskites on many two-dimensional material substrates, such as mechanically exfoliated MoS2, mechanically exfoliated black phosphorus (BP), CVD graphene, CVD WS2, and chemically exfoliated MoS2, are predominantly realized via one-step spin-coating using perovskite precursor solutions. Beyond just layering structures, a one-step spin-coating method can be employed to introduce 2D materials like graphene derivatives, MXene, and BP directly into perovskite films. Studies have shown that integrating 2D materials with perovskites leads to larger grain sizes in the films as a result of a slower crystallization process [31]. Additionally, these 2D materials can help in reducing surface defects and enhance the conductivity of the perovskite films, thereby improving their performance in optoelectronic applications [32].
Another promising approach for the large-scale production of perovskite and 2D material heterostructures is through direct solution-based synthesis. By mixing pre-fabricated perovskite with 2D materials in solution, the two components can be combined through either chemical bonds or non-covalent interactions, resulting in heterogeneous micro or nano-structures. In 2018, Wu et al. prepared MAPbI3/Reduced graphene oxide (RGO) micro-composites by blending MAPbI3 and GO in a strongly acidic water-based solution [33]. Upon being subjected to visible light for a few hours, the mixture of MAPbI3 and GO naturally transitioned into MAPbI3/RGO micro-assemblies by the reduction process of GO to RGO. Given the rich presence of functional groups like hydroxyl, carbonyl, and carboxyl on the RGO surface and edges, a formation of Pb-O-C linkages are established between MAPbI3 and RGO. In addition to chemical bonds induced by surface functional groups, non-covalent interactions also offer a strategy to construct the perovskite and 2D material heterostructures. However, producing perovskite crystals on the surface of TMDCs is intricate, primarily due to the pronounced difference in the surface chemistry between perovskites and TMDCs. Hassan and his team introduced a method that can be processed in a solution to produce MoSe2-CsPbBr3 nanocomposites, using 4-aminothiophenol (4-ATP) as the bridging agent between MoSe2 and CsPbBr3, as depicted in Figure 1d [28]. Initially, 4-ATP was added to MoSe2 nanosheets rich in defects, enabling the passivation of free Mo atoms via chemical bonding. Subsequently, pre-fabricated CsPbBr3 nanocrystals were integrated into the solution containing MoSe2 nanosheets functionalized with 4-ATP. Note that the 4-ATP not only serves as a binder between TMDCs and CsPbBr3 nanocrystals through electrostatic attraction but also ensures that MoSe2 and CsPbBr3 are closely in contact, promoting efficient charge transfer across the MoSe2-CsPbBr3 boundary.
Thermal injection offers a straightforward and rapid technique for producing perovskite nanocrystals, typically operating at elevated temperatures (around 120–200 °C) under inert gas conditions, boasting high in situ growth efficiency [34,35]. In the improved thermal injection technique for constructing heterostructures of perovskites and 2D materials, 2D nanosheets are typically introduced to the Pb precursor solution beforehand, resulting in a well-mixed solution. Following the addition of cesium oleate, the desired heterostructure of perovskites and 2D materials forms easily in the subsequent chemical reaction. Notwithstanding the advancements in this method, the synthesis demands high temperatures, often ranging from 120 to 200 °C, and the protection of an inert gas like nitrogen. These requirements contribute to significant costs and increased energy usage. As an alternative, the ligand-assisted reprecipitation (LARP) method offers a more efficient approach to synthesize perovskites in ambient conditions [36]. Typically, in LARP synthesis, a mixture of perovskite precursors, including lead halides, alkylammonium halides, and surfactants, is gradually introduced to a non-polar solvent, such as toluene, under agitation. This change in solvent polarity triggers the precipitation of perovskite nanocrystals. For the fabrication of heterostructures involving perovskites and 2D materials, 2D nanosheets can either be incorporated into the perovskite precursor mix or directly into the non-polar solvents. Later, the perovskites and 2D materials heterostructure can be formed when the perovskite precursor solution is merged with the non-polar solvent under stirring.
Achieving atomically precise epitaxial interfaces is crucial for enhancing device performance. Yet, the epitaxial growth of halide perovskites with such sharp interfaces remains unachieved. The inherent high ion mobility in these materials causes issues like interdiffusion, which broadens junction widths. Additionally, their chemical instability often results in the decomposition of layers during subsequent layer preparations [37,38]. Shi et al. report a method to considerably curb in-plane ion diffusion in 2D halide perovskites through the inclusion of rigid π-conjugated organic ligands [39]. They constructed two distinct lateral 2D halide perovskite heterostructures using varied organic ligands: one employed a conjugated ligand based on bithiophenylethylammonium (2T+), while the other utilized a more common alkyl ligand, butylammonium (BA+). Notably, the Br–I heterostructure in 3D perovskites lacks thermal stability owing to the high diffusivity of Br and I anions, and even a mild heating can initiate and speed up the interdiffusion of these halide anions across the heterojunction [40]. However, with the introduction of conjugated 2T ligands, the authors observed a marked reduction in interdiffusion between Br and I within the (2T)2PbI4–(2T)2PbBr4 heterostructures. Employing these rigid conjugated ligands in 2D perovskite frameworks has been shown to dramatically suppress ionic interdiffusion across 2D halide perovskite junctions. By inhibiting ion migration, they were able to produce stable interfaces with near-atomic precision through continuous epitaxial growth.
This section provides a brief overview of the various techniques used to fabricate heterostructures combining perovskites and 2D materials. These methods can be broadly classified into solid-state and solution-based methods. The solid-state method includes vapor deposition and the dry/wet transfer method. While these processes can yield high-quality heterostructures with precision, their prohibitive costs and complex procedures make them less suited for large-scale manufacturing and broader applications. On the other hand, the spin-coating technique offers a more practical alternative, given its simplicity, low-temperature requirements, and energy efficiency. This makes it an attractive method for producing heterostructures of perovskites and 2D materials. For those seeking to synthesize these heterostructures directly in solution, there are two main strategies available: ex situ hybridization and in situ growth. Ex situ hybridization is a direct and uncomplicated method that combines perovskite crystals with 2D materials either through chemical bonding or non-covalent interactions. However, this sometimes results in weaker contacts between the two components. Meanwhile, the in situ growth method allows for the reprecipitation of perovskite crystals onto 2D nanosheets in the solution, leading to a more integrated and optimized heterointerface. This method is especially promising for achieving high-quality heterostructures at an affordable cost and with a high yield. Table 1 compares the advantages and disadvantages of different fabrication techniques.

3. Perovskite and 2D Material Heterostructure Photodetectors

In recent years, with the development of the optoelectronic information industry, there has been a heightened demand for enhanced performance of optoelectronic devices to cater to diverse applications [41]. Photodetection operates on the principle of translating absorbed light signals into electron–hole pairs. This conversion is facilitated by semiconductor materials possessing appropriate band gaps. Under the influence of either an inherent or externally applied electric field, these electron–hole pairs are collected via metal electrodes, resulting in the generation of electrical signals. The main indicators for evaluating photodetection are responsivity (R), detectivity (D*), on/off ratio, external quantum efficiency (EQE), photoconductive gain (G), noise equivalent power (NEP), and response speed (τrise, τdecay). R is the ratio of the photocurrent to the incident light power, expressed as R = Iph/Pin, where Iph is the photocurrent and Pin is the incident light power. Its value depends on the wavelength of the light, as the absorption of photons is related to the band gap of the semiconductor. EQE is the number of electron–hole pairs collected divided by the number of incident photons, that is, EQE = (Iph/e)/(Pin/Eph), where e is the charge of the electron and Eph is the energy of the photon. Photoconductive gain evaluates the ability of a photodetector (PD) to generate charge carriers in response to a single photon. The response speed is usually characterized by the light pulse’s rise and decay/fall times on and off, i.e., the time interval required for the output signal to increase from 10% to 90% and decrease from 90% to 10% of its on-state maximum value, respectively. D is defined as the inverse of the noise equivalent power, and its normalized result to the active detection area and detection bandwidth, called D*, has been widely used to quantify the sensitivity of PDs to noise signals. D* = R/(2eJd)1/2, where Jd is the dark current density in Jones or cm Hz1/2 W−1.
Perovskites have developed tremendously in the past time, and the performance of PDs made of a single material has not been greatly improved. Two-dimensional materials have the advantages of high carrier mobility, a rich and tunable band gap, fast photoresponse speed, and wide spectral response range [41]. Therefore, researchers use perovskite and two-dimensional material heterostructures to construct PDs, and the performance of PDs has been greatly improved. In this section, we will introduce perovskite heterostructure PDs from the discussion of perovskite/graphene, perovskite/TMDCs, perovskite/BP, and perovskite/MXene.

3.1. Perovskite/Graphene Heterostructures

Perovskites offer significant advantages such as a high light absorption coefficient, extended carrier diffusion length, and superior carrier mobility, positioning them as prime candidates for light harvesting in PDs [42]. However, a primary limitation of these PDs is the rapid recombination of photo-induced electron–hole pairs within few picoseconds in perovskites, preventing their collection by the electrodes. To address this, introducing graphene, known for its quick response time and exceptional carrier mobility, can serve as an efficient carrier transport material in PDs. Graphene has attracted enormous attention since its discovery. The carrier mobility of graphene is as high as 15,000 cm2 v−1s−1, and due to its zero-bandgap nature, graphene has an ultrabroad absorption spectrum from ultraviolet to the terahertz region [43]. However, its relatively low absorption cross-section can restrict the broader use of pure graphene in practical optoelectronic applications [44]. To improve the performance of perovskite-based PDs, Lee et al. pioneered the concept of a perovskite/graphene heterostructure PD, paving the way for the development of high-performance devices [45]. The device boasts impressive photosensitive characteristics across the entire visible light spectrum and the n-octadecyltrimethoxysilane layer serves as a surface modifier, minimizing surface charge traps on the SiO2. There is a notable decrease in the PL intensity of the CH3NH3PbI3 perovskite/graphene heterojunction, ensuring efficient electron and hole separation. Due to graphene’s ultra-high electron mobility, holes are transported multiple times within the channel before recombining with trapped electrons. Consequently, an impressive photoresponsivity of 180 A W−1 and effective quantum efficiency 5 × 104% are attained under a substantial illumination power of 1 µW.
While integrating a high-efficiency carrier transport layer enhances performance, it also inadvertently creates a leakage pathway, thereby increasing the device dark current. To address the rise in dark current, improve the on–off current ratio, and maintain efficient carrier transport, Tan et al. proposed a graphene/perovskite/graphene structure perovskite PD, which has its source and drain graphene electrodes covered, showcases a minimal dark current (≈10−10 A) and high on/off current ratio (≈103) [46]. Due to the smaller active area, its photocurrent is comparably lower (~100 nA). To further optimize the dark current and enhance the PD performance, Chen et al. introduced a specific channel structure on graphene. This design ensures an effective carrier separation within perovskites by curtailing the leakage current [47]. Due to the different work functions of perovskite and graphene, the difference in the Fermi level makes the perovskite band at the interface bend upward. Since a Schottky junction is formed at the interface of perovskite/graphene, and a large built-in electric field is introduced, which causes electrons from graphene to migrate into the valence band of perovskite, occupying its vacant states and effectively suppress the recombination of electron–hole pairs within the perovskite. Schottky junctions are not only formed in the vertical alignment of the graphene/perovskite interface but also emerge on either side of the channel. Consequently, PDs structured with the graphene/perovskite/graphene configuration display an impressive on/off switch ratio of 2.64 × 103 and responsivity of 22 mA W−1 when illuminated with 452 nm light. To elucidate the effect of horizontally structured Schottky junctions on photocurrent, one, five, and nine-channel graphene devices are discussed. Despite the formation of additional Schottky junctions, the photocurrent diminishes with an increase in channel numbers due to constrained carrier transfer in multi-channel devices. Additionally, a rise in the recombination and quenching of carriers further reduces the photocurrent. Notably, introducing a single graphene channel yields the most optimal detector performance.
The heterojunction can be structured both horizontally and vertically to enhance device performance. For instance, Chang et al. fabricated graphene-MAPbI3 perovskite hybrid phototransistors by vapor deposition [48]. This device exhibited an astounding responsivity of 1.73 × 107 A W−1 and a superb detectivity of 2 × 1015 Jones, with an effective quantum efficiency approximating 108% across the visible light spectrum. Zou et al. proposed a vertical-structure PD with perovskite single-crystal MAPbBr3 as the light absorption layer and graphene as the transport layer [49]. Due to the mismatch in the graphene and perovskite work functions and the formed interfacial electric field, holes are transferred to the graphene, while electrons are bound to the perovskite, acting as additional phototunable gates. The electric field created by these electrons induces hole generation in the graphene. This, in turn, depresses the Fermi level of the graphene, further amplifying the photocurrent. Under the conditions of a 3 V bias and 532 nm laser exposure, the device exhibits remarkable photoresponse performance at a low light intensity (0.66 mW cm−1), displaying values ≈ 1017.1 A W−1 for photoresponse, approximately 2.02 × 1013 Jones for photodetection and an ultra-high photoconductive gain of ≈2.37 × 103. In comparison to the individual perovskite device, the performance is greatly improved.
By combining perovskites with high-mobility 2D materials like graphene and localized surface plasmon resonance (LSPR) metal nanostructures to create heterojunctions, one can bolster carrier transport and photoresponsivity, subsequently enhancing device performance [50]. Notably, perovskites produce a substantial volume of photo-induced electron–hole pairs, while gold nanostructures possess light-trapping capabilities and amplify electromagnetic fields. Sun et al. found that, upon incorporating plasmonic gold nanoparticles (NPs) (approximately 40 nm in diameter) just beneath the graphene in a perovskite/graphene heterojunction PD, there was a near two-fold increase in responsivity by a factor enhancement of 1.96 [51] (Figure 2a). As shown in Figure 2b, a significantly larger photoresponse can be observed in perovskite/graphene/Au NPs devices compared to perovskite/graphene PDs. Figure 2c describes the generation, diffusion, and transfer of photo-induced carriers in the perovskite layer, comparing scenarios with and without the presence of gold NPs. The majority of photo-induced carriers in the perovskite/graphene/Au NPs PD are produced near the interface of perovskite and graphene, attributed to the near-field enhancement caused by Au NPs. Since these carriers originate close to the perovskite/graphene interface, they have a reduced possibility of undergoing recombination during their transport to graphene compared to carriers that are generated at a distance from the interface. This phenomenon is associated with the gradual reduction in responsivity observed with increased light intensity in perovskite/graphene/Au NPs devices, as shown in Figure 2d.Wang et al. achieved a 59% improvement in the photoresponsivity of vertical perovskite PDs by incorporating gold nanorods into the perovskite layer [52]. Similarly, Liu and colleagues presented devices that combined perovskite, graphene, and Ag nanoparticles, where R was increased by a factor of 7.45 due to the incorporation of Ag nanostructures [53]. Feng et al. fabricated novel broadband quasi 2D perovskite (BA)2(FA)n−1PBnI3n+1 hybrid-structure PD with good stability by combining both monolayer graphene and Au square nanoarrays, as shown in Figure 2e,f [54]. Due to the near-field enhancement of the gold array, most of the photogenerated carriers are generated near the graphene–perovskite interface. Reduced carrier recombination, combined with the efficient extraction of electrons from the perovskite layer via graphene, enhances the photoelectric response of the Graphene-Au-perovskite PDs Therefore, a hybrid system employing both graphene and gold square nanoarrays can markedly boost carrier mobility and light absorption. This synergistic effect, optimizing optical trapping and promoting light-induced carrier extraction, results in a significant surge in photocurrent throughout the visible and near-infrared spectrum. The graphene-Au array-perovskite-based PDs had a low dark current of 10−10 A, large on/off ratio of 104, high responsivity of 18.71 A W−1, and detectivity of 2.21 × 1013 Jones. The responsivity and detectivity were two orders of magnitude higher than those of PDs based only on perovskites, as shown in Figure 2g. Regarding PDs based on perovskite/graphene heterostructures, several limitations remain, including a high dark current, reduced on/off ratio, and detectivity stemming from graphene’s inherent zero bandgap characteristic. In this sense, the electron blocking layer can be added to improve the on/off ratio using poly(3-hexylthiophene) (P3HT) and [6,6]-phenyl-C61-butyric acid methyl ester (PCBM) can be added at the interface, acting as an electron trapping site to enhance the charge separation efficiency and thus the effective gain and responsivity enhancement [55,56]. However, with the incorporation of PCBM, the fall time is observed to be much longer than the rise time, likely because of the slower electron detrapping process from PCBM aggregates. So, effective gain and responsivity enhancement occurs at the price of much lower fall and rise time, which is still not desirable for practical application. Also, the deposition of perovskite on graphene causes the Dirac point of the graphene to shift to a more positive level, leading to a decrease in the mobility of the graphene.

3.2. Perovskite/TMDC Heterostructures

Although perovskite has a significant light absorption coefficient and generates substantial photogenerated carriers after illumination, how to effectively collect them through electrodes is a problem. This issue can be solved using graphene with high electron mobility as a transport layer. Apart from graphene, TMDCs are also a good solution due to their excellent optical properties, electrical properties, and abundant band gaps [57]. Although 2D TMDCs have excellent properties, it is still difficult for them to fully absorb incident light due to the atomic layer thickness and weak light absorption coefficient, limiting the performance of PDs. Hence, developing heterojunction devices with perovskite and TMDCs emerges as a promising strategy to enhance the performance of PDs. Using two materials with suitable energy band structures to form a heterojunction to create a type II energy band arrangement can significantly speed up the separation of carriers, increase the photocurrent, reduce the dark current, and improve the response speed for the device. In addition, the composite structure of different materials can also obtain a wider photoresponse range.
Kang et al. reported the first perovskite/MoS2 heterojunction PD [58]. In this configuration, MAPbI3 serves as the photoactive layer, while multilayer MoS2 functions as the transport layer, as illustrated in Figure 3a. As a laser strikes the perovskite/MoS2 junction, the predominant generation of electron–hole pairs occur in the perovskite absorber layer, given its superior absorption property, and type I heterojunction facilitates the transfer of both photo-induced electrons and holes from the perovskite to MoS2. In the presence of an external electric field, these electrons and holes move in opposite directions within the MoS2 channel, producing a photocurrent. Notably, the MAPbI3/MoS2/APTES device reached a responsivity of 2.11 × 104 A W−1 and a specific detectivity of 1.38 × 1010 Jones, respectively. (Figure 3b,c) The performance of the perovskite/MoS2 PD was further improved by Sun et al., with the device showcasing rapid, stable, and sensitive response features even at minimal operating potentials. At an incident power of 2.2 pW (520 nm) and a bias voltage of 2 V, the responsivity and specific detection rate are 342 A W−1 and 1.14 × 1012 jones. Without the package, the device demonstrated high stability in transient on/off tests, with response and recovery times of 27 ms and 21 ms, respectively [59].
The noble metal disulfide compound PtSe2 stands out as an emerging TMDC material. In recent times, it has garnered substantial research attention, primarily because, compared to traditional TMDCs, it has high electron mobility and a narrower band gap (1.25 eV), with good air stability [61]. It is a good material for photoelectric detection. Zhang et al. developed a self-driven, air-stable, ultrafast PD by combining perovskite with multilayer PtSe2, and a schematic diagram of the device structure is shown in Figure 3d [61]. Zhang et al. fabricated large-scale PtSe2 thin films by selenization of a pre-deposited Pt layer and employed cesium-doped FACsPbI3 perovskite as the light-absorbing medium, which demonstrated a notably higher electron mobility in comparison to the more commonly used MAPbI3 perovskite. Beyond this, it also proved to be more stable when exposed to air. As showcased in Figure 3e, the energy bands of the two materials form a type II band alignment, which facilitates the rapid separation of electrons and holes near the heterojunction interface, driven by an inherent electric field. The efficient separation of photogenerated electrons and holes is further demonstrated by the markedly diminished photoluminescence (PL) intensity of the PtSe2/perovskite hybrid compared to that of the perovskite film alone (Figure 3f). This mechanism is responsible for the impressively swift rise and fall times recorded at 78 and 60 ns, respectively (Figure 3g). At zero bias, the light/dark ratio is 5.7 × 103, the responsivity is 117.7 mA W−1, and the specific detection rate is 2.91 × 1012 Jones. A notable observation from this study was the PD resilience; even after being stored in standard ambient conditions for three weeks, it managed to maintain most of its photoresponsive characteristics. In another study, CsPbBr3/MoS2-based PD was reported to have excellent thermal stability and operational durability in air [62]. By manipulating the characteristic parameters, particularly the solution concentration and the surface ligand content of colloidal CsPbBr3, one can control and modulate the optoelectronic properties of these detectors, which are closely related to the light absorption capability of the perovskite layer and the charge transfer efficiency at the heterojunction interface. Oxygen adsorbents play a significant role in this system. Their typical p-type behavior enhances the reaction rate of the PD. In the air, when the gate voltage is zero, the photoresponsivity is 6.40 × 105 mA W−1, the external quantum efficiency is 1.50 × 105%, and the specific detection rate is 3.38 × 1011 Jones.
Moreover, 2D Ruddlesden–Popper perovskites demonstrate significant promise for optical and optoelectronic applications [63,64]. Their superior environmental stability over 3D perovskites positions them as potential replacements for 3D variants in light-absorbing layers, aiming to enhance the efficiency of monolayer TMDC PDs. However, studies on the optoelectronic properties of TMDC monolayer/2D perovskite vertical heterostructures are still at an early stage. For example, Fang et al. reported PDs based on n-type MoS2 and p-type lead-free 2D Ruddlesden–Popper perovskite (PEA)2SnI4 heterostructures [65]. This combined device can detect light across the full visible to near-infrared spectrum, offering adjustable photoresponse peaks. Incorporating a few layers of graphene flakes as electrical contacts further improves the device performance, recording a response rate of 1100 A W−1 at 3 V bias and ~40 ms without any bias. Wang et al. reported graphene/perovskite/WS2/graphene heterojunction devices [66] (Figure 4a). Under laser illumination, the photogenerated electrons move from the perovskite layer to the WS2 monolayer, whereas the photogenerated holes travel in the reverse direction. This results in an accumulation of extra electrons in the WS2 monolayer, which then combine with the photogenerated electron–hole pairs in WS2, leading to the formation of negatively charged trions, depicted in Figure 4b. Compared with the photoresponse of the WS2 monolayer region, there is a notable surge in photocurrent, as seen in Figure 4c. Figure 4d compares the photoresponsivity of the WS2/PEPI heterostructure with that of the WS2 monolayer, both measured at the same bias voltage Vds of 4 V. It is observed that the photoresponsivity of heterointerface device surpasses that of the WS2 monolayer device. For instance, at a laser power of 50 μW, the heterostructure photoresponsivity achieves 0.13 mA W−1, which is five times greater than the 0.027 mA W−1 responsivity of the WS2 monolayer device. This enhance-ment is attributed to the superior light absorption capacity of the PEPI layer, coupled with the augmented charge carrier extraction efficiency facilitated by the innate potential at the WS2 and PEPI interface. Additionally, the photoresponsivity and EQE of the self-driven PD reach 24.2 μA W−1 and 5.7 × 105, respectively.
Over the past few years, there has been notable advancement in perovskite/TMDCs heterostructures PDs. However, most of the photoelectric responses of 2D perovskite PDs are restricted to the ultraviolet-visible spectrum [68]. Given the substantial intrinsic optical band gap of 2D perovskites, their sensitivity to near-infrared (NIR) wavelengths is practically nonexistent. However, the type II staggered band structure observed in the TMDC-based vdW heterojunction facilitates the inter-layer charge transition between two distinct materials [69]. This means that under light exposure, electrons can shift from the valence band (VB) of one material to the conduction band (CB) of another, creating an energy divergence broader than the inherent bandgaps of both materials individually. This phenomenon broadens the detection capability of heterojunction PDs to encompass longer wavelengths. Zhou et al. realized a high-performance near-infrared PD using 2D perovskite/MoS2 [67]. (Figure 4e) Within the heterostructure region, the interlayer coupling-dependent energy gap is effectively narrowed to less than 0.8 eV, as shown in Figure 4f. Experimental findings reveal that the MoS2/2D perovskite vdW heterostructure responds distinctly to 1550 nm laser exposure, achieving a peak photoresponsivity of 121 A W−1 and a detectivity of 4.3 × 1014 Jones at a wavelength of 860 nm (Figure 4g,h).
In summary, numerous perovskite/TMDC heterostructure PDs have been developed, taking advantage of the exceptional absorption of perovskite and the advanced electronic features of 2D TMDCs. It may be observed that there is a trade-off among the performance metrics for various PDs. Specifically, PDs based on perovskite/TMDC heterostructures showcase exceptionally high responsivity, surpassing that of typical commercial PDs. However, they lag in response speed. In a PD, the gain is intrinsically linked to carrier mobility and carrier lifetime. The remarkable responsivity seen in perovskite/TMDC PDs is often attributed to the significantly high gain due to extended carrier lifetimes, which majorly influences the speed of PDs. However, this prolonged carrier lifetime poses a significant challenge to their market adoption. Therefore, efforts to enhance responsivity should focus on increasing gain through boosting carrier mobility instead of lengthening carrier lifetime.

3.3. Perovskite/BP Heterostructures

In addition to using graphene and TMDC to build heterostructures, black phosphorus is also a good choice due to its unique anisotropic structure, excellent light absorption ability, high carrier mobility, and unique layer-controlled direct bandgap [70]. Combining black phosphorus with perovskites enables the performance improvement of PDs.
Zou et al. introduced an innovative Schottky barrier-controlled phototransistor where the channel current is primarily governed by the Schottky barrier at the source electrode [71]. The device is based on a few-layer BP crystal channel decorated by adding a MAPbI3-xClx perovskite layer. Due to the band alignment of the BP/perovskite heterojunction, holes migrate to the BP channel while electrons amass within the perovskite layer. These gathered electrons induce an added electric field adjacent to the Schottky barrier, which subsequently lowers the barrier height. The pristine BP device without the perovskite layer demonstrates a bipolar nature with an on–off ratio exceeding 104. After coating MAPbI3-xClx, the transfer curves shift towards a higher gate voltage, highlighting the p-type doping effect on the BP channel brought about by the perovskite layer. The peak responsivity reaches 4 × 106 A W−1 at a gate voltage of approximately −30 V. Distinct from traditional phototransistors, this innovative device utilizes a field-assisted mechanism to release photocarriers within its channel. Consequently, the device exhibits high responsivity up to 108 A W−1 and a high specific detection rate up to 9 × 1013 Jones.
Up until now, the majority of 2D PDs reported in the literature operate under forward bias. This mode of operation often resembles a photoconductor with a gating effect, rather than adhering to the intended photovoltaic mechanism. This observed gating effect in photosensitive devices is frequently termed trap-assisted photoconductivity, where photogenerated carriers circulate multiple times within the device channel filled with trapped charges before recombination occurs. The short time necessitates a device channel with high conductivity, which inadvertently leads to a significant dark current. Faced with these challenges, there is an urgent need for a more efficient device structure that is not limited by the above-mentioned operational mechanisms so that the PDs are either fast and insensitive or sensitive and slow. Wang et al. proposed and developed a hybrid device based on perovskite and BP/MoS2 PDs, and the device structure is given in Figure 5a [72]. It is noted in Figure 5b that type I and type II heterojunctions are formed at the perovskite/BP and BP/MoS2 interfaces, respectively. When the perovskite film is exposed to light, photocarriers are spontaneously generated, being attributed to the intrinsic electric field present at the BP/MoS2 junction. These photocarriers subsequently migrate into the BP layer where they undergo separation and collection. As shown in Figure 5c, it is observed that the pure perovskite film will display a pro-nounced PL peak at 775 nm, related to the bandgap transition of MAPbI3. However, nota-ble PL quenching effects are observed in the perovskite/BP, perovskite/MoS2, and perov-skite/BP/MoS2 heterostructures, which is attributed to the efficient exciton dissociation and interlayer charge transfer between layers. Figure 5d presents the I–V curve for a typical p–n PD with BP/MoS2 heterostructure, displaying clear rectifying behavior with a rectifica-tion ratio of 20. When capped with the perovskite layer, the device dark current reduces by roughly two orders of magnitude compared to the original BP/MoS2 device. Distinguishing it from traditional detectors that rely on photovoltaic and gating mechanisms, this hybrid device stands out due to its elevated responsivity and swift response time. When the reverse bias voltage is −2 V, the responsivity of the device can be significantly improved to 11 A W−1, and the response time can be optimized to hundreds of microseconds. The corresponding detective rate is as high as 1.3 × 1012 Jones. Moreover, the combination of perovskites’ efficient light absorption and the substantial electric field created by the BP/MoS2 junction makes this hybrid device exceptionally compatible for self-driven, broadband photodetection. Under 457 nm laser irradiation and zero bias, the device has a high detectivity of 3 × 1011 Jones and an ultra-high light/dark value of 3 × 107.

3.4. Perovskite/MXene Heterostructures

In recent years, two-dimensional Ti3C2Tx (MXene) has attracted much attention due to its distinctive properties such as metallic conductivity, adjustable work function, commendable light transmittance, and the ability to form stackable conductive films, making it well-suited for flexible electronic devices [74,75]. Deng showcased a novel approach by introducing an all-spray processable, wide-area, flexible PD array that combined 2D MXene and 2D CsPbBr3. This was implemented on ordinary paper, serving as electrodes and the primary active materials and the detailed structure of this device can be visualized in Figure 5e [73]. Notably, the authors realized the optical communication function by fabricating a large-area array of 1665 pixels within 72 cm2. Once the illumination energy is larger than the semiconductor bandgap, a large number of photogenerated electron–hole pairs in the CsPbBr3 nanosheets are rapidly separated by the reverse local electric field and have a low recombination rate, resulting in a free carrier concentration increase, as shown in Figure 5f. Conversely, during the charge transfer process at the interface between CsPbBr3 and MXene, charge accumulation and band bending create a depletion layer. This layer plays a crucial role in enhancing the separation efficiency of photogenerated excitons. The light response speed is faster than 18 ms, and the on/off ratio reaches 2.3 × 103. The optoelectronic characteristics of the MXene-based PD at various bending degrees can be observed in Figure 5g. As the bending angle progressively increases, there is a slight decline in the photocurrent, which eventually stabilizes. This behavior underscores the fact that both MXene nanosheets and CsPbBr3 nanosheets possess excellent layered structures, coupled with an impressive ability to relieve stress. This research not only showcases a cost-effective technique for PD fabrication but also hints at promising applications in the realms of wearable electronics and optical communication.
The electromagnetic field near the surface is naturally intensified due to the oscillation of electrons within metal nanoparticles, which can enhance PD performance even when incident light is scarce. However, the enhanced electric field can only be distributed near the surface of metal nanoparticles, limiting the light absorption capacity of the entire photosensitive layer [76]. Two-dimensional materials, with their extensive surface area and minimal thickness, allow for more interaction with the photosensitive layer without adding stress to the film. Among MXenes, Ti3C2Tx nanosheets are noteworthy. They exhibit outstanding features such as high conductivity (104 S cm−1), broad spectral response, and a work function that aligns well with CsPbBr3, and abundant free electrons offer great possibilities to improve the optoelectronic response of perovskite films instead of metal nanostructures. Li et al. proposed a simple method to prepare high-performance all-inorganic halide perovskite CsPbBr3 quantum dot PDs by uniformly mixing two-dimensional Ti3C2Tx nanosheets with different concentrations [77]. Electrons at the conduction band tend to be spontaneously injected into the 2D Ti3C2Tx nanosheets after excitation, and ions can be partially enhanced due to the enhanced carrier transport. Simulations using the finite-difference time-domain (FDTD) method demonstrate that the enhanced near-surface electromagnetic field of Ti3C2Tx nanosheet notably increases the light efficiency of the entire CsPbBr3 QD film. In comparison with unmodified CsPbBr3 QD devices, the presence of the nanosheets hot electrons considerably enhances the photocurrent, leading to an impressive 300% surge in EQE. The photogenerated carriers naturally migrate from the CsPbBr3 QDs to the Ti3C2Tx nanosheets due to the creation of downward band bending, which is also beneficial for enhancing the photocurrent. Under 490 nm illumination, the CsPbBr3 QD/MXene nanosheet PDs have a response rate of 97 μA W−1 at a power of 2.9 mW cm−2. This peak performance was sustained for four months in normal atmospheric conditions, suggesting a promising approach to achieve advancements in the mass production of perovskite optoelectronic devices.
This section mainly introduces the research progress of perovskite and 2D material heterostructures in PDs. PD devices based on perovskite/2D heterostructures generally perform better than single devices due to the ultrafast charge transfer and unique energy band mechanism at the heterointerface. Heterostructures combining perovskites and TMDCs mainly absorb light in the visible to near-IR range, a characteristic stemming from the bandgaps of the base materials and the way their energy bands align. When targeting the crucial mid-to-far IR spectrum, which has specific resonance with many molecules, researchers have turned to materials like graphene and other 2D semiconductors with narrow gaps, such as BP. Nonetheless, these materials present challenges. For instance, graphene-integrated PDs often have significant dark current, while BP-based ones suffer from reduced resilience in ambient conditions. The distinct characteristics of interlayer excitons in 2D TMDCs pave the way for creating intriguing and novel devices [78]. Perovskites have great advantages as light-absorbing layers in heterostructures. Two-dimensional materials have rich and tunable band gaps, extremely high electron mobility, ultra-thin and transparent properties, and can obtain PDs with excellent performance by constructing heterojunctions with perovskites. The PD parameters in this section are summarized in Table 2.

4. Perovskite and 2D Material Heterostructure Solar Cells

Solar cells are essential devices that convert light energy into electrical energy. As the global population grows and the demand for energy increases, ensuring a stable supply of energy becomes even more critical. At present, there have been many problems in the way of obtaining energy by burning fossil fuels, such as insufficient fossil resources and environmental pollution [80]. The use of efficient and clean solar energy to replace fossil energy is a very effective method.
The solar energy conversion efficiency of perovskite solar cells (PSCs) has reached 25.2% in the past decade, which has attracted great interest, and the device structure of PSCs can be divided into the n-i-p structure and p-i-n structure [81]. PSC devices typically consist of a transparent bottom electrode, an electron transport layer (ETL), a perovskite photoactive layer, a hole transport layer (HTL), and a top electrode. The fundamental operation of a solar cell can be described in these steps: First, when exposed to light, electron–hole pairs form in the perovskite layer. Then, driven by the inherent electric field, the generated electrons are captured by the ETL, while the holes move into the HTL. These photo-induced charge carriers are then gathered by the electrodes and converted into electrical energy for output. The key parameters to evaluate the performance of PSCs are open-circuit voltage (Voc), short-circuit current density (Jsc), fill factor (FF), and power conversion efficiency (PCE). The charge transport material dramatically influences the efficiency and stability of PSCs [82]. HTL and ETL materials promote efficient carrier extraction and recently suppress recombination between perovskite films and electrodes. Two-dimensional materials are often used as charge transport layers for PSCs to collect electrons and holes to improve performance efficiently. Note that graphene and other 2D materials can also be used as additives in the photoactive layer inside PSCs; this part of the review is not included, but it can be found in previous review paper [22]. Additionally, perovskite itself as a 2D additive within 3D perovskite and its potential for photovoltaics is also not included here in this part.

4.1. Perovskite/Graphene

For PSCs, the ideal interfacial layers are those that can be processed from solutions, are affordable, highly conductive, and maintain stability either chemically or thermally. Also, these layers should align their energy bands appropriately with the photoactive layer to ensure selective charge extraction [83,84]. Graphene, with its high electron mobility, stands out as an optimal material for extracting and transferring charges to electrodes. For instance, to extract electrons from the perovskite, a layer of graphene quantum dots (GQDs) was strategically positioned between the perovskite photoactive layer and the TiO2, serving as the ETL to extract electrons from the perovskite [85] (Figure 6a). The improvement in efficiency is mainly attributed to the increase in photocurrent since the incident photon-to-current conversion efficiency (IPCE) indicates a significant amplification in the conversion of absorbed photons to current within the visible to near-infrared spectrum due to the incorporation of GQDs. Utilizing ultrafast transient absorption spectroscopy, the researchers observed that the electron extraction time for the PGT film at 90~106 ps, substantially quicker than the PT film (260~307 ps), which can effectively compete with carrier trapping, as shown in Figure 6b. With the addition of GQDs, the PCE of the solar cell was significantly improved from 8.81% to 10.15% (Figure 6c). Agresti et al. further optimized the electron extraction performance of the ETL layer [86]. Li-neutralized graphene oxide was used as the ETL in FTO/cTiO2/G+mTiO2/GO-Li/CH3NH3PbI3/spiro-OMeTAD/Au PSCs, and the PCE is increased from 11.6% to 12.6%.
In addition to being used as ETL, graphene has also been used as the HTL to improve the performance of solar cells. Wang et al. formed strong Pb-Cl and Pb-O bonds between [CH(NH2)2]x[CH3NH3]1-xPb1+yI3 films with Pb-rich surfaces and chlorinated graphene oxide layers [87]. Serving as the HTL, Cl-GO exhibits an efficient extraction of holes. Under a forward scan, the perovskite/Cl-GO cell achieved notable metrics: a PCE of 21.08%, a Jsc of 23.82 mA cm−2, a Voc of 1.12 V, and a fill factor (FF) of 0.79. Furthermore, forming strong chemical bonds on the surface of soft perovskite films stabilizes the perovskite heterostructure, which considerably curtails the deterioration of perovskite components, subsequently minimizing harm to the organic HTLs. As a result, the stability of PSC was dramatically improved, maintaining 21% of the initial value after 1000 h of operation at 60 °C under AM1.5G solar lamp and 1000 h of operation at a maximum power point of 90% of the efficiency.

4.2. Perovskite/TMDCs

Apart from graphene, TMDCs have exhibited notable benefits as charge transport materials in PSCs. For example, Zhao et al. reported the low-temperature processing of 2D SnS2 nanosheets as a novel ETL in PSCs and achieved a PCE of 13.63% for the device, a Jsc of 23.70 mA cm−2, a Voc of 0.95 V, and a FF of 0.61 [88]. Similarly, 2D SnS2 was used as an ETL for flexible PSCs, and the structure is shown in Figure 7a [89]. As the ETL of PSCs, SnS2 can align its conduction band minimum with that of the perovskite and photoanode to ensure smooth carrier transfer. In addition, the authors optimized the SnS2 film thickness. As shown in Figure 7b, with a SnS2 ETL of 60 nm, the PCE of the device reached 13.2% due to the best coverage, electron extraction and larger charge recombination resistance, with a Voc of 1.011 V, a Jsc of 1.011 V, a Jsc of 21.70 mA cm−2, and a FF of 0.60. In addition, since MAPbI3 can destroy ZnO, the introduction of the SnS2 buffer layer also improves the stability of PSCs: The device using ZnO ETL degrades very seriously at 40~50% humidity, resulting in an initial after 30 days. PCE lost about 30%. In contrast, SnS2-based devices exhibited good long-term stability with almost unchanged PCE with only 7% loss in PCE after 30 days of storage (Figure 7c). Zhao et al. used 2D SnS2 ETL prepared via the self-assembly stack deposition method for efficient planar PSCs [90]. The Pb-S intermolecular interaction between perovskite and SnS2 helps in passivating the interface trap states. This interaction reduces charge recombination and facilitates electron extraction at the interface of both the ETL/perovskite and the HTL/perovskite, ensuring a balanced charge transport throughout the device. As a result, a PCE of more than 20.12% was achieved, with an outstanding Voc of 1.161 V, a Jsc of 23.55 mA cm−2, and a high FF of 0.73 s.
Meanwhile, the HTL also plays an essential role in PSCs. The efficiency with which holes are separated directly influences the device performance [92]. Due to their high charge mobility, 2D TMDCs are also suitable candidates for the HTL. Notably, while impressive PCEs exceeding 20% have been realized using Spiro-OMeTAD as HTLs in PSCs, the migration of iodine from the perovskite active layer through Spiro-OMeTAD to the metal electrodes poses a challenge to the efficiency and durability of the devices [93]. To eliminate this effect, Capasso et al. added a few-layer MoS2 flakes as a buffer in PSCs [94]. The MoS2 layer serves a dual purpose. It acts as a protective barrier, preventing any interaction between the perovskite and the electrode. This protective measure avoids efficiency losses that can occur due to the direct contact of the perovskite with the electrode. Simultaneously, the MoS2 sheet functions as an HTL, facilitating the effective transfer of holes from the perovskite to the electrode. The PCE of the as-obtained PSCs was 13.3%. Additionally, the stability of the device is significantly improved: the PCE of PSCs without MoS2 decreased by 14.2% in the 550 h durability test, while that of PSCs with MoS2 added decreased to 12.4% after 550 h. Wang et al. mixed MoS2 nanosheets in a PEDOT:PSS layer to form a hybrid HTL layer embedded in inverted PSCs [91] (Figure 7d). MoS2 nanosheets enhance the charge extraction efficiency of the HTL, mitigating recombination at the interface and substantially decreasing electrode polarization and hysteresis effects. Impedance spectroscopy (IS) analysis reveals a 50% rise in the composite resistance when MoS2 is combined with PEDOT:PSS in the HTL. As shown in Figure 7e, the average Jsc for six samples of the device, enhanced with a 10% MoS2 blend, at 20.7 mA cm−2, outperforming the pristine PEDOT:PSS HTL devices. How-ever, the Voc of the modified HTL-based device is marginally lower than that of the un-modified PEDOT:PSS HTL device. Devices with the MoS2 hybrid PEDOT:PSS HTL witnessed a significant boost in their PCE, achieving approximately 14.7% (Figure 7f). Moreover, these modified HTL-based devices demonstrated impressive durability, maintaining over 95% of their initial PCE even after a duration of 4 weeks.

4.3. Perovskite/BP

Liquid-exfoliated, few-layered 2D BP nanosheets have been explored as HTLs for PSCs by Muduli et al. [95]. Photoelectron spectroscopy (PESA) measurements carried out in ambient conditions verified that the valence band level of BP nanosheets, at −5.2 eV, is optimally suited for the hole injection of CH3NH3PbI3. The electrical properties of the best-performing PSC incorporating BP nanosheets as HTM. When the BP nanosheets are fully utilized as HTM, a PCE of 7.88% is obtained, an improvement over the PCE value of 4% for the HTM-free device. For comparison, the authors mixed BP nanosheets with the commonly used Spiro-OMeTAD as HTL, achieving a PCE of 16.4%. In contrast, using Spiro-OMeTAD exclusively as HTL yielded a PCE of just 13.1%, indicating that BP can be vital in facilitating hole extraction.
Photostability remains a significant concern for the widespread adoption of PSCs. A key degradation pathway, especially under continuous illumination, is the reduction of the Pb2+ ions in the commonly used MAPbI3 perovskite material to elemental metallic Pb0 during continuous illumination. Pb0, as a deep defect state within the perovskite structure, not only degrades the performance of PSC devices but also reduces the long-term stability of the devices [96]. The introduction of 2D BP into MAPbI3 can not only delay hot carrier recombination but also suppress the Pb0 defect formation, improving the stability of PSCs. Wang et al. added BP to MAPbI3 perovskite to fabricate PSCs (Figure 8a) [97]. The PCE of the device is 20.62%, with a Jsc of 23.31 mA cm−2, Voc of 1.082 V and FF of 0.82. Both photostability and photovoltaic performance were significantly enhanced due to the application of BP (Figure 8b,c). The addition of BP has also greatly improved the stability of PSC and provided people with excellent ideas. In particular, in a dry N2 glove box, the initial efficiency of MAPbI3/BP-based PSCs remained at 94% after 1000 h of continuous white LED irradiation, while the initial efficiency of PSCs without BP addition dropped to 30%.

4.4. Perovskite/MXene

MXenes possess outstanding optoelectronic characteristics, superior electrical conductivity, and enhanced carrier mobility, making them prime candidates for fabricating PSC interfacial layers. Utilizing 2D Ti3C2Tx through spin-coating, Yang et al. developed flat and uniform ETL surfaces for ITO/ETL/CH3NH3PbI3/Spiro-OMeTAD/Ag PSCs [100]. A brief UV ozone treatment on metallic Ti3C2Tx enhances its surface bonding with Ti-O without affecting its inherent properties, thereby improving its suitability as an ETL. This process improved electron transfer and suppressed the recombination at the ETL/perovskite interface, elevating the PCE of untreated Ti3C2Tx films from 5.00% to a notable 17.17% after being subjected to UV-ozone for 30 min. Following their initial findings, Yang et al. continued to optimize the performance of PSCs by introducing in situ oxidized Ti3C2Tx (O-Ti3C2Tx) MXene with SnO2, creating a nanoscale heterojunction and using SnO2 as the ETL of MAPbI3 PSCs [98] (Figure 8d). The oxidation process transitioned the Ti3C2Tx from being predominantly metallic to semiconducting, and this change allowed for the tuning of Ti3C2Tx properties to better align with the energy levels of perovskites. Adding O-Ti3C2Tx also enhanced the electron mobility of SnO2 ETL. As a result, the O-Ti3C2Tx/SnO2 heterojunction excelled in electron extraction and minimized the recombination between the ETL and perovskite, resulting in a significant improvement in the performance of PSC, with PCE increasing from 17.68% to 20.09% and enhanced stability in air (Figure 8e). After being stored in ambient air for over 1000 h, the O-Ti3C2Tx/SnO2 heterojunction device maintained 82% of its original PCE, showing great long-term stability, as seen in Figure 8f. Furthermore, MXenes can serve as HTLs within PSCs. Chen et al. introduced 2D Ti3C2-MXene as an interlayer into all-inorganic PSCs as HTL [99] (Figure 8g). Between the CsPbBr3 layer and the carbon electrode, the presence of 2D Ti3C2-MXene improves the energy level alignment, accelerates hole extraction, and achieves efficient carrier transport and low interfacial recombination. Compared with the one without MXene, PCE increased from 8.21% to 9.01% (Figure 8h,i). Moreover, the functional groups within Ti3C2-MXene effectively passivated perovskite grains, diminished trap defects in CsPbBr3 films, and elevated the overall quality of the perovskite films. Impressively, the film showcased excellent hygrothermal stability, lasting for 1900 h and extending beyond 600 h.
This section presents the recent research progress of perovskite and 2D material heterostructures in PSCs. ETL and HTL are the critical parts of PSCs, and whether the material can efficiently transfer charges directly affects the overall performance of PSCs. So far, significant research has been directed towards transition metal oxide ETLs, including titanium dioxide (TiO2), zinc oxide (ZnO), and tin dioxide (SnO2), but they come with multiple limitations that restrict the device performance [101]. For example, TiO2, widely used as an ETL, results in subpar device stability and an elevated density of trap states at the TiO2/perovskite boundary [102]. Conversely, Spiro-MeOTAD, which is the predominant HTL material used in n-i-p PSCs, faces stability issues and requires complex doping processes. Additionally, the material PEDOT:PSS is compromised due to its acidic nature and its vulnerability to moisture [103]. Two-dimensional materials are characterized by their exceptionally high electron mobility and diverse, tunable energy bands. When selected appropriately for interface layers, these materials facilitate ultrafast charge extraction. Furthermore, 2D materials serve as efficient barriers, protecting the perovskite photoactive layer from moisture. This makes them promising candidates for both ETL and HTL in high-performance PSCs. In general, PSCs that incorporate 2D materials as interfacial layers (like HTL and ETL) tend to exhibit superior performance and stability compared to traditional configurations. These 2D layers not only enable rapid charge extraction but also act as effective shields against moisture for the perovskite photoactive layer. Though perovskite and 2D material heterostructures have been incorporated into PSCs to enhance both PCE and stability, the intricacies of the interface dynamics and interactions between perovskites and 2D materials remain largely unexplored. Delving deeper into the interface structure is essential to gain clearer insights into the formation mechanism and the combined benefits of perovskite and 2D material heterostructures. Table 3 compares the key parameters of the PSCs discussed in this section.

5. Perovskite and 2D Material Heterostructure Photocatalysis

The advancement of photocatalysis is viewed as one of the most promising approaches to address the growing energy and environmental problems, during which H2 can be produced using photocatalytic techniques and CO2 can be converted [104]. The photocatalytic process typically involves three stages: (1) under light exposure, the photocatalyst produces free charges, (2) these charges migrate to the photocatalyst’s surface, and (3) a redox reaction takes place with the adsorbed substances [105]. From a thermodynamic perspective, the potential of the conduction band ought to be more negative, while the potential of the valence band should be more positive than the reaction potential of particular species. For example, in order to facilitate photocatalytic H2 evolution, the conduction band potential of an optimal photocatalyst should lie more negative than the H+ reduction potential. When assessing photocatalysis, metrics such as CO2 reduction rate, H2 production rate, and pollutant degradation rate are used, varying based on the specific photocatalysis reaction. Conventional photocatalysis materials such as TiO2, ZnO, and ZnS usually suffer from narrow light absorption, short photoinduced charge lifetime, and poor stability [106]. As a new type of semiconductor material, perovskite has a high absorption coefficient and can fully use light energy; however, the carrier recombination in perovskite seriously affects the performance of the catalyst. Constructing a heterojunction with 2D materials can quickly separate the charges and improve the catalyst’s performance. Next, some developments in perovskites and 2D materials for photocatalysis will be introduced.

5.1. Perovskite and 2D Material Heterostructures for Hydrogen Production

Hydrogen is increasingly viewed as a potential successor to fossil fuels in the upcoming years given its exceptional purity, high energy density per unit mass, and non-polluting combustion [107]. Currently, hydrogen production predominantly relies on the steam reforming of natural gas. This method necessitates the use of non-renewable fossil fuels, operates at elevated temperatures, incurs high production expenses, and contributes to rising greenhouse gas emissions. Using solar energy for photocatalytic hydrogen production offers a compelling alternative to traditional steam reforming. However, the development of photocatalytic hydrogen evolution technology has been curtailed due to the slow reaction rates and instability of semiconductor photocatalysts [108]. Perovskite materials with high absorption coefficients, abundant band gaps, and long charge diffusion lengths are suitable candidates for photocatalytic hydrogen evolution [109]. Nevertheless, the rapid charge recombination observed in perovskites impedes efficient charge transport, consequently restricting H2 production [110]. To enhance the photocatalytic activity of perovskites, noble metals such as platinum are usually used as cocatalysts, but the rarity and elevated cost hinder its widespread use in large-scale photocatalytic H2 production [111]. In recent times, 2D materials have gained attention as cocatalysts for perovskites, given their cost-effectiveness, robust stability, and impressive carrier mobility, thereby enhancing photocatalytic performance.
In 2016, Park et al. first used 3D bulk MAPbI3 powder in conjunction with a saturated HI aqueous solution to stably achieve photocatalytic decomposition of HI, releasing hydrogen in the process [111]. Notably, splitting HX presents certain advantages over direct water splitting as HX reduction engages only two electrons, in contrast to the four-electron process of water splitting. Moreover, perovskites boast a substantial absorption coefficient for visible light, thus fully harnessing light energy. Reduced graphene oxide (rGO) acts as both an electron acceptor and transporter. It enhances the performance of semiconductor photocatalysts by facilitating the transfer of photogenerated electrons within semiconductors and also serves as a framework, augmenting the surface area [112]. To further improve the efficiency, Wu et al. reported a novel MAPbI3/rGO composite structure as a photocatalyst for hydrogen evolution [33]. The MAPbI3/rGO composite structure can be obtained by dispersing GO powder in a high-acid aqueous solution saturated with MAPbI3 and then exposing the solution to visible light (λ ≥ 420 nm) for several hours of photoreaction. Figure 9a illustrates that the MAPbI3/rGO composite exhibits superior catalytic activity, with a hydrogen evolution rate of 93.9 μmol h−1. Under visible light exposure (λ ≥ 420 nm) and intensity below 120 mW cm−2, this rate is an impressive 67 times faster than that of unmodified MAPbI3. Furthermore, the composite demonstrates high stability, with consistent hydrogen evolution results even after 200 h, as depicted in Figure 9b. To investigate the charge transfer properties of MAPbI3 and MAPbI3/rGO, the corresponding photoelectrodes were assembled, and electrochemical impedance spectroscopy (EIS) analysis was conducted in a dichloromethane solution. The Nyquist plot, depicted in Figure 9c, reveals that the charge transfer resistance in MAPbI3/rGO films is markedly reduced compared to MAPbI3. This observation is corroborated by the notably smaller semicircle evident in the plot for MAPbI3/rGO, underscoring the enhanced charge transport ability offered by the incorporated rGO, thus facilitating swifter charge transfer. Consequently, during the photocatalytic activity, photogenerated electrons within MAPbI3 are swiftly relayed to rGO via the Pb-O-C bond, which ensures their segregation from the photogenerated holes. At the rGO location, H+ gets reduced to H2. Simultaneously, the photogenerated holes participate in the oxidation of I to I 3 .
Wang et al. pioneered the utilization of Cs2AgBiB6 (CABB) double perovskite for cracking hydrobromic acid (HBr) when exposed to visible light [115]. RGO, with good electron transfer properties and acid stability, was introduced to enhance the hydrogen production performance further. Using the CABB/RGO composite as a photocatalyst under visible light (λ ≥ 420 nm, via a 300 W Xe lamp), they were able to generate a remarkable 489 μmol g−1 H2 for 10 h. Notably, the exemplary photocatalyst of CABB combined with 2.5% RGO showcased robust stability, persisting through 120 h of ongoing photocatalytic hydrogen evolution. In a saturated solution, CABB perovskite undergoes a dynamic precipitation-dissolution equilibrium process, and RGO can attach CABB particles. As a proficient visible light absorber, CABB produces electrons and holes when irradiated. These generated electrons, in turn, transfer to the conductive RGO via the M-O-C bonds, which reduces H+ to generate H2 at the active site of RGO. Moreover, BP is renowned for its outstanding optoelectronic characteristics, serving as an efficient charge-transfer material. Li et al. synthesized a BP/MAPbI3 heterostructure in a MAPbI3-saturated HI aqueous solution system by an electrostatic coupling method and used it as a photocatalyst for HI splitting for hydrogen production [116]. The type I heterojunction on the BP/MAPbI3 interface captures the photogenerated electrons from MAPbI3, significantly diminishing the recombination of these electrons. As a result, the engineered BP/MAPbI3 photocatalyst showcases an impressive capability for H2 production when exposed to visible light, achieving a photocatalytic rate of 3742 μmol h−1g−1. Remarkably, this rate of hydrogen evolution is a hundred times faster compared to MAPbI3 at 35 μmol h−1g−1. Furthermore, the BP/MAPbI3 photocatalyst maintains its stability in a MAPbI3-saturated HI solution, with its performance remaining largely consistent throughout the photoreaction period.
MoS2, another non-noble metal layered semiconductor, has garnered significant research interest because of its outstanding optoelectronic characteristics [117]. Wang et al. prepared a multilayer MoS2/MAPbI3 composite catalyst using a facile in situ coupling method, combining synthesized MAPbI3 powder with MoS2 nanosheet powder [118]. Compared with pristine MAPbI3, the composite structure has a higher photocatalytic rate of 2061 µmol h−1, a 121-fold increase. The composite maintained stability throughout the 156-h hydrogen evolution reaction. By employing in situ rapid crystallization on a monolayer of MoS2, a new type II heterojunction photocatalyst was introduced, leading to the creation of MoS2/MAPbI3 heterojunction catalysts [119]. The robust built-in electric field in the MoS2/MAPbI3 enhances charge separation, which is crucial for efficient HI splitting and photocatalytic hydrogen generation. A type II heterojunction, along with this strong electric field, is established between the precisely structured MAPbI3 and MoS2 nanosheets. This has been directly validated using Kelvin probe force microscopy (KPFM). As a result, this takes the hydrogen production of perovskite photocatalysts to unprecedented levels, achieving a hydrogen production efficiency of 1.09% and an activity rate of 13,600 µmol g−1h−1 when exposed to visible light. Compared with the single MAPbI3 catalyst, the catalytic efficiency of the heterojunction is 220 times higher. Guan et al. also adopted the MAPbI3/MoS2 composite structure, and the MoS2 nanoflowers with abundant active sites were assembled with MAPbI3 crystallites to form a heterostructure [113]. Remarkably, its rate of hydrogen evolution reaches up to 30,000 μmol g−1h−1, a rate more than 1000 times greater than that of pure MAPbI3 under the visible light. (as depicted in Figure 9d) Importantly, the solar splitting efficiency peaks at 7.35%, positioning it among the highest efficiencies recorded so far. By incorporating MoS2, which has a suitable band configuration and unsaturated entities, it can effectively promote charge separation and provide more active sites for hydrogen production.

5.2. Perovskite and 2D Material Heterostructures for CO2 Reduction

Using solar energy for the photocatalytic reduction of CO2 to transform it into renewable fuels is seen as a potential remedy for the escalating CO2 levels in the atmosphere [120,121]. Unfortunately, most pure perovskites have low photocatalytic efficiency due to their fast charge recombination. Consequently, the creation of composite structures has been pursued to enhance the performance of perovskite-based catalysts. Graphene and its offshoots serve as superior charge transport mediums, facilitating the increase of electron extraction and transportation, thereby boosting catalytic efficiency. Xu et al. constructed a CsPbBr3 QD/graphene oxide (GO) composite as a catalyst for CO2 reduction, as shown in Figure 9e [114]. Compared with single CsPbBr3 QDs, the electron consumption rate was increased from 23.7 μmol g−1h−1 to 29.8 μmol g−1h−1 after the introduction of GO. Photoluminescence and photoelectrochemical impedance, as showcased in Figure 9f, g prove that the boost in photocatalytic activity is attributed to the electrical connectivity offered by conductive GO. This ensures the effective conveyance of photogenerated carriers from the perovskite. Contrary to pure graphene, rGO nanosheets often display defects on their basal planes and edges due to lingering oxygen-related impurities [122]. Using the combination of perovskite with rGO-like defects, Wang et al. obtained an efficient catalyst for CO reduction [123]. Due to the composite structure of Cs4PbBr6/rGO, Cs4PbBr6/rGO exhibited higher selectivity (~94.6%) and 6.2 times higher CO2 conversion (11.4 μmol g−1h−1) than pristine Cs4PbBr6. One of the key reasons is that these surface irregularities in rGO serve as trapping centers, enabling electrons from the excited Cs4PbBr6 to be adeptly channeled to the defect spots within rGO. The other is that oxygen deficiency acts as an active center, assisting CO2 adsorption, activation and reduction of CO. Among these oxygen-based defects, the OH group emerges as the prime site for the photoreduction of CO to CO, due to its minimal energy barrier for CO adsorption, the creation of CO*, and its subsequent release. This is the first study to improve CO2 photoreduction activity and selectivity by introducing OH defects, providing a new way to construct high-performance, low-cost perovskite-type solar energy conversion composites. While the heterostructures of perovskites and 2D materials have risen to prominence as potent materials in the realm of photocatalysis, the issue of lead toxicity continues to hinder their widespread adoption. Wang et al. synthesized a novel tin atom-sharing Cs2SnI6 perovskite nanocrystal/SnS2 nanosheet heterojunction by a facile solution method [124]. A joint study combining DFT calculations and transient absorption and KPFM characterizations shows that holes are transferred from SnS2 to Cs2SnI6, and electrons are transferred from SnS2 to Cs2SnI6. Further investigations through photocatalytic CO2 reduction and photo-electrochemical tests confirmed that the build-up of electrons in SnS2 effectively extends the charge lifetime, enhancing efficiency by 5.4-fold and 10.6-fold. Given its non-toxic nature and ease of fabrication, this catalyst shows immense promise for widespread application.
Two-dimensional black phosphorus (BP) has also been widely used as a metal-free candidate catalyst for photocatalytic purposes and when combined with other semiconductors to form heterojunctions, BP exhibits superior performance in CO2 photoreduction [125]. For the first time, Wang et al. synthesized a CsPbBr3 heterostructure onto 2D BP, showing its efficacy as a photocatalyst for CO2 reduction [126]. In comparison to the single-component perovskite CsPbBr3, the introduced BP greatly enhances the photocatalytic activity, and the conversion rates of CO2 to CO and CH4 are 44.7 and 10.7 μmol g−1 h−1, respectively. CO2 photocatalytic reduction of CH4 and CH4. The photocatalytic CO2 reduction yields for CO and CH4 rose by factors of 2.4 and 4.4, respectively. The CsPbBr3/BP composition, with its structural and interfacial interactions, offers a higher number of active sites that facilitate the adsorption and activation of CO2. Concurrently, the integrated BP serves as a channel for electron transfer, enhancing the effective use of charge carriers that reach the surface. On the other hand, MXene is also an excellent conductive material. Pan et al. synthesized CsPbBr3/Ti3C2Tx nanocomposites by in situ growth of CsPbBr3 on the surface of monolayer Ti3C2Tx nanosheets [127]. The pronounced quenching of fluorescence in CsPbBr3 nanocrystals, alongside time-resolved photoluminescence lifetime and photoconductivity assessments, confirmed the effective charge transfer between Ti3C2Tx and CsPbBr3 nanocrystals. When exposed to simulated sunlight, the CO production rate for CsPbBr3/Ti3C2Tx clocked in at 26.32 μmol g−1 h−1. This is significantly superior to the yield from standalone CsPbBr3 nanocrystals, which was less than 4.4 μmol g−1 h−1.
This section mainly describes the application of perovskite/two-dimensional material conforming structure in photocatalysis, including CO2 emission reduction and H2 production. People are facing huge problems with energy and the environment. Photocatalytic CO2 emission reduction combined with H2 production presents an effective solution to these challenges. While the exploration of perovskites and 2D materials in heterostructures for photocatalysis is still in its early stages, the current study opens promising prospects for visible light-driven photocatalysts based on perovskite and 2D material heterostructures. Table 4 summarizes the relevant parameters of photocatalysts based on perovskite and 2D material heterostructures in this section.

6. Conclusions and Perspectives

This review covers the synthesis methods of perovskite and 2D material heterostructures and their applications in photodetectors, solar cells and photocatalysis. First, the fabrication techniques are introduced for perovskite and 2D material heterostructures, including solid-state and solution methods. These can be broadly classified into solid-state and solution methods. The solid-state techniques are suitable for producing high-quality perovskite and 2D material heterostructures characterized by sharp and clean interfaces. In contrast, solution-processed techniques are more apt for mass-scale manufacturing of these heterostructures. We then partition the exploration of applications into three distinct segments: photodetectors, solar cells, and photocatalysis, detailing the advancements and potentialities in each domain. Due to fast charge transfer and transport at the interface, combining 2D materials with perovskite facilitates better figure-of-merits in hybrid photodetectors than those with a single material. In addition, the addition of 2D materials inside PSCs leads to more stable perovskite thin films and more efficient charge separation and collection when 2D materials are used as ETLs and HTLs additives. In addition, perovskite and 2D material heterostructures in photocatalysis exhibit great potential as high-performance and stable photocatalysts due to ultrafast charge separation and transport in perovskites and 2D materials. Numerous studies in materials and physics still await, with a wealth of potential yet to be unearthed. Some foundational characteristics of perovskite and 2D material heterostructures have already been examined, including PL quenching, enhancement, and shifts attributable to interlayer charge transfers, as well as interface-driven broadband emissions and selective spin injections. Nevertheless, other prospective properties remain unverified in experimental settings. A deeper investigation into the optical properties would be beneficial, especially at lower temperatures. Analyzing these properties in relation to temperature can offer valuable insights into the recombination mechanism and phonon interactions, and such temperature-dependent studies can help identify fine spectral components that cannot be discernible at room temperature.
While perovskite and 2D material heterostructures hold immense promise, several hurdles persist, notably concerning stability, scalability, and material safety. To begin with, the stability of perovskite and 2D material heterostructures is often compromised. The inherent sensitivity of perovskite to external factors such as moisture, oxygen, light, and heat largely contributes to its poor stability [128]. Therefore, it is inevitable to utilize appropriate passivation strategies for fabricating devices with high performance and long-term stability. Chemical passivation primarily addresses intrinsic point defects through chemical bonding, while physical passivation relies on physical processes like strain relaxation or treatments such as temperature adjustments, lasers and adhesive tape [129]. Therefore, enhancing the stability of perovskite and 2D material heterostructures using both chemical and physical passivation techniques could offer a viable path forward. At the same time, novel perovskites have been developed for stability enhancement. Notably, the 2D layered Ruddlesden–Popper perovskites, which incorporate hydrophobic organic chains, have displayed superior photostability when juxtaposed with their traditional 3D counterparts [130]. Similarly, 2D layered Dion−Jacobson perovskites have emerged as a potential material with excellent environmental stability. The unique structure of Dion−Jacobson perovskites features adjacent inorganic layers connected by diammonium ligands. These ligands serve to eradicate the van der Waals (vdWs) gaps and reduce the inter-slab distance. Consequently, they offer enhanced stability when compared with the 2D layered Ruddlesden–Popper perovskites [131]. These novel 2D layered perovskites are promising candidates for fabricating stable perovskite and 2D material heterostructures.
Secondly, there is a trade-off between high-quality perovskite and 2D material heterostructures and high yields. Though solution-processed methods facilitate mass production of perovskite and 2D material heterostructures, the heterostructure quality is adversely affected by the poor contact and large lattice mismatch at the heterointerface. It has been reported that seed engineering is a successful approach to the controlled synthesis of hierarchical heterostructures, which may also be used for the highly controlled epitaxial growth of perovskite and 2D material heterostructures [132]. It also has been demonstrated that the dangling-bond-free surfaces of 2D material nanosheets as synthetic templates may trigger the solution-phase epitaxial growth of other materials [133]. Therefore, the solution-processed epitaxy technique may provide a promising construction of perovskite and 2D material heterostructures.
Lastly, the most extensively studied perovskites in heterostructures contain toxic and dangerous heavy metals, especially Pb, which is concerning both for environmental reasons and human health. This toxicity issue of lead-related perovskite and 2D material heterostructures can be solved by replacing Pb with other elements such as Sn, Mn, and Zn [134]. However, performances of optoelectronics or photocatalysis are still unsatisfactory compared with their Pb-based counterparts. For instance, devices based on Sn underperformed when compared to those using Pb. This inferior performance can be attributed to a high defect density within the perovskite lattice and the significant oxidation of Sn, transitioning from a +2 to a +4 state, which hampers their prolonged stability. Therefore, the development of perovskites based on non-toxic substances is required for future advancement of perovskite and 2D material heterostructures optoelectronics.

Author Contributions

S.F.: writing—original draft, writing—review and editing, formal analysis; B.L.: writing—original draft, writing—review and editing, formal analysis; B.X.: formal analysis; Z.W.: writing—review and editing, formal analysis, visualization, project administration; resources. S.F. and B.L. contributed equally. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

This research was funded by the National Natural Science Foundation of China (62175160).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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] [PubMed]
  2. Lin, K.; Xing, J.; Quan, L.N.; de Arquer, F.P.G.; Gong, X.; Lu, J.; Xie, L.; Zhao, W.; Zhang, D.; Yan, C. Perovskite light-emitting diodes with external quantum efficiency exceeding 20 percent. Nature 2018, 562, 245–248. [Google Scholar] [CrossRef]
  3. Dou, L.; Yang, Y.; You, J.; Hong, Z.; Chang, W.-H.; Li, G.; Yang, Y. Solution-processed hybrid perovskite photodetectors with high detectivity. Nat. Commun. 2014, 5, 5404. [Google Scholar] [CrossRef] [PubMed]
  4. Jeong, M.; Choi, I.W.; Go, E.M.; Cho, Y.; Kim, M.; Lee, B.; Jeong, S.; Jo, Y.; Choi, H.W.; Lee, J. Stable perovskite solar cells with efficiency exceeding 24.8% and 0.3V voltage loss. Science 2020, 369, 1615–1620. [Google Scholar] [CrossRef] [PubMed]
  5. 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]
  6. Zhao, Y.; Ma, F.; Qu, Z.; Yu, S.; Shen, T.; Deng, H.-X.; Chu, X.; Peng, X.; Yuan, Y.; Zhang, X.; et al. Inactive (PbI2)2RbCl stabilizes perovskite films for efficient solar cells. Science 2022, 377, 531–534. [Google Scholar] [CrossRef]
  7. Zhao, Y.; Zhu, K. Organic–inorganic hybrid lead halide perovskites for optoelectronic and electronic applications. Chem. Soc. Rev. 2016, 45, 655–689. [Google Scholar] [CrossRef] [PubMed]
  8. Stranks, S.D.; Eperon, G.E.; Grancini, G.; Menelaou, C.; Alcocer, M.J.; 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]
  9. Xing, G.; Wu, B.; Wu, X.; Li, M.; Du, B.; Wei, Q.; Guo, J.; Yeow, E.K.; Sum, T.C.; Huang, W. Transcending the slow bimolecular recombination in lead-halide perovskites for electroluminescence. Nat. Commun. 2017, 8, 14558. [Google Scholar] [CrossRef]
  10. Cheng, Y.; Ding, L. Pushing commercialization of perovskite solar cells by improving their intrinsic stability. Energy Environ. Sci. 2021, 14, 3233–3255. [Google Scholar] [CrossRef]
  11. Ghosh, J.; Giri, P. Recent advances in perovskite/2D materials based hybrid photodetectors. J. Phys. Mater. 2021, 4, 032008. [Google Scholar] [CrossRef]
  12. Chen, S.; Shi, G. Two-dimensional materials for halide perovskite-based optoelectronic devices. Adv. Mater. 2017, 29, 1605448. [Google Scholar] [CrossRef] [PubMed]
  13. Novoselov, K.; Mishchenko, A.; Carvalho, A.; Castro Neto, A. 2D materials and van der Waals heterostructures. Science 2016, 353, aac9439. [Google Scholar] [CrossRef] [PubMed]
  14. Mas-Balleste, R.; Gomez-Navarro, C.; Gomez-Herrero, J.; Zamora, F. 2D materials: To graphene and beyond. Nanoscale 2011, 3, 20–30. [Google Scholar] [CrossRef] [PubMed]
  15. Geim, A.K.; Grigorieva, I.V. Van der Waals heterostructures. Nature 2013, 499, 419–425. [Google Scholar] [CrossRef]
  16. Rhodes, D.; Chae, S.H.; Ribeiro-Palau, R.; Hone, J. Disorder in van der Waals heterostructures of 2D materials. Nat. Mater. 2019, 18, 541–549. [Google Scholar] [CrossRef]
  17. Doganov, R.A.; O’farrell, E.C.; Koenig, S.P.; Yeo, Y.; Ziletti, A.; Carvalho, A.; Campbell, D.K.; Coker, D.F.; Watanabe, K.; Taniguchi, T. Transport properties of pristine few-layer black phosphorus by van der Waals passivation in an inert atmosphere. Nat. Commun. 2015, 6, 6647. [Google Scholar] [CrossRef]
  18. Hart, J.L.; Hantanasirisakul, K.; Lang, A.C.; Anasori, B.; Pinto, D.; Pivak, Y.; van Omme, J.T.; May, S.J.; Gogotsi, Y.; Taheri, M.L. Control of MXenes’ electronic properties through termination and intercalation. Nat. Commun. 2019, 10, 522. [Google Scholar] [CrossRef]
  19. Gupta, A.; Sakthivel, T.; Seal, S. Recent development in 2D materials beyond graphene. Prog. Mater. Sci. 2015, 73, 44–126. [Google Scholar] [CrossRef]
  20. Cao, L. Two-dimensional transition-metal dichalcogenide materials: Toward an age of atomic-scale photonics. Mrs Bull. 2015, 40, 592–599. [Google Scholar] [CrossRef]
  21. Miao, S.; Liu, T.; Du, Y.; Zhou, X.; Gao, J.; Xie, Y.; Shen, F.; Liu, Y.; Cho, Y. 2D Material and Perovskite Heterostructure for Optoelectronic Applications. Nanomaterials 2022, 12, 2100. [Google Scholar] [CrossRef]
  22. Zhang, Z.; Wang, S.; Liu, X.; Chen, Y.; Su, C.; Tang, Z.; Li, Y.; Xing, G. Metal Halide Perovskite/2D Material Heterostructures: Syntheses and Applications. Small Methods 2021, 5, 2000937. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, Q.; Linardy, E.; Wang, X.; Eda, G. Excitonic energy transfer in heterostructures of quasi-2D perovskite and monolayer WS2. ACS Nano 2020, 14, 11482–11489. [Google Scholar] [CrossRef] [PubMed]
  24. Xie, C.; Liu, C.K.; Loi, H.L.; Yan, F. Perovskite-based phototransistors and hybrid photodetectors. Adv. Funct. Mater. 2020, 30, 1903907. [Google Scholar] [CrossRef]
  25. Fang, C.; Wang, H.; Li, D. Recent progress in two-dimensional Ruddlesden–Popper perovskite based heterostructures. 2D Mater. 2021, 8, 022006. [Google Scholar] [CrossRef]
  26. Niu, L.; Liu, X.; Cong, C.; Wu, C.; Wu, D.; Chang, T.R.; Wang, H.; Zeng, Q.; Zhou, J.; Wang, X.; et al. Controlled Synthesis of Organic/Inorganic van der Waals Solid for Tunable Light-Matter Interactions. Adv. Mater. 2015, 27, 7800–7808. [Google Scholar] [CrossRef]
  27. Lu, J.; Carvalho, A.; Liu, H.; Lim, S.X.; Castro Neto, A.H.; Sow, C.H. Hybrid Bilayer WSe2-CH3NH3PbI3 Organolead Halide Perovskite as a High-Performance Photodetector. Angew. Chem. Int. Ed. Engl. 2016, 55, 11945–11949. [Google Scholar] [CrossRef]
  28. Hassan, M.S.; Basera, P.; Bera, S.; Mittal, M.; Ray, S.K.; Bhattacharya, S.; Sapra, S. Interfaces, Enhanced photocurrent owing to shuttling of charge carriers across 4-aminothiophenol-functionalized MoSe2–CsPbBr3 nanohybrids. ACS Appl. Mater. Interfaces 2020, 12, 7317–7325. [Google Scholar] [CrossRef]
  29. Martins, L.G.; Song, Y.; Zeng, T.; Dresselhaus, M.S.; Kong, J.; Araujo, P.T. Direct transfer of graphene onto flexible substrates. Proc. Natl. Acad. Sci. USA 2013, 110, 17762–17767. [Google Scholar] [CrossRef]
  30. Cheng, H.C.; Wang, G.; Li, D.; He, Q.; Yin, A.; Liu, Y.; Wu, H.; Ding, M.; Huang, Y.; Duan, X. van der Waals Heterojunction Devices Based on Organohalide Perovskites and Two-Dimensional Materials. Nano Lett. 2016, 16, 367–373. [Google Scholar] [CrossRef]
  31. You, P.; Tang, G.; Cao, J.; Shen, D.; Ng, T.-W.; Hawash, Z.; Wang, N.; Liu, C.-K.; Lu, W.; Tai, Q. 2D materials for conducting holes from grain boundaries in perovskite solar cells. Light Sci. Appl. 2021, 10, 68. [Google Scholar] [CrossRef] [PubMed]
  32. Qin, Z.; Chen, Y.; Zhu, K.; Zhao, Y. Two-dimensional materials for perovskite solar cells with enhanced efficiency and stability. ACS Mater. Lett. 2021, 3, 1402–1416. [Google Scholar] [CrossRef]
  33. Wu, Y.; Wang, P.; Zhu, X.; Zhang, Q.; Wang, Z.; Liu, Y.; Zou, G.; Dai, Y.; Whangbo, M.H.; Huang, B. Composite of CH3NH3PbI3 with Reduced Graphene Oxide as a Highly Efficient and Stable Visible-Light Photocatalyst for Hydrogen Evolution in Aqueous HI Solution. Adv. Mater. 2018, 30, 1704342. [Google Scholar] [CrossRef]
  34. Yang, D.; Li, X.; Zhou, W.; Zhang, S.; Meng, C.; Wu, Y.; Wang, Y.; Zeng, H. CsPbBr3 quantum dots 2.0: Benzenesulfonic acid equivalent ligand awakens complete purification. Adv. Mater. 2019, 31, 1900767. [Google Scholar] [CrossRef] [PubMed]
  35. Adhikari, G.C.; Vargas, P.A.; Zhu, H.; Zhu, P. Saponification precipitation method for CsPbBr3 nanocrystals with blue-green tunable emission. J. Phys. Chem. C 2018, 123, 1406–1412. [Google Scholar] [CrossRef]
  36. Zhu, F.; Men, L.; Guo, Y.; Zhu, Q.; Bhattacharjee, U.; Goodwin, P.M.; Petrich, J.W.; Smith, E.A.; Vela, J. Shape evolution and single particle luminescence of organometal halide perovskite nanocrystals. ACS Nano 2015, 9, 2948–2959. [Google Scholar] [CrossRef] [PubMed]
  37. Lee, J.W.; Park, N.G. Chemical approaches for stabilizing perovskite solar cells. Adv. Energy Mater. 2020, 10, 1903249. [Google Scholar] [CrossRef]
  38. Quintela, C.X.; Song, K.; Shao, D.-F.; Xie, L.; Nan, T.; Paudel, T.R.; Campbell, N.; Pan, X.; Tybell, T.; Rzchowski, M.S. Epitaxial antiperovskite/perovskite heterostructures for materials design. Sci. Adv. 2020, 6, eaba4017. [Google Scholar] [CrossRef]
  39. Shi, E.; Yuan, B.; Shiring, S.B.; Gao, Y.; Akriti; Guo, Y.; Su, C.; Lai, M.; Yang, P.; Kong, J.; et al. Two-dimensional halide perovskite lateral epitaxial heterostructures. Nature 2020, 580, 614–620. [Google Scholar] [CrossRef]
  40. Pan, D.; Fu, Y.; Chen, J.; Czech, K.J.; Wright, J.C.; Jin, S.J.N. Visualization and studies of ion-diffusion kinetics in cesium lead bromide perovskite nanowires. Nano Lett. 2018, 18, 1807–1813. [Google Scholar] [CrossRef]
  41. Long, M.; Wang, P.; Fang, H.; Hu, W. Progress, challenges, and opportunities for 2D material based photodetectors. Adv. Funct. Mater. 2019, 29, 1803807. [Google Scholar] [CrossRef]
  42. Wangyang, P.; Gong, C.; Rao, G.; Hu, K.; Wang, X.; Yan, C.; Dai, L.; Wu, C.; Xiong, J.J.A.O.M. Recent advances in halide perovskite photodetectors based on different dimensional materials. Adv. Opt. Mater. 2018, 6, 1701302. [Google Scholar] [CrossRef]
  43. Geim, A.K. Graphene: Status and prospects. Science 2009, 324, 1530–1534. [Google Scholar] [CrossRef] [PubMed]
  44. Mueller, T.; Xia, F.; Avouris, P.J.N. Graphene photodetectors for high-speed optical communications. Nat. Photonics 2010, 4, 297–301. [Google Scholar] [CrossRef]
  45. Lee, Y.; Kwon, J.; Hwang, E.; Ra, C.H.; Yoo, W.J.; Ahn, J.H.; Park, J.H.; Cho, J.H. High-performance perovskite-graphene hybrid photodetector. Adv. Mater. 2015, 27, 41–46. [Google Scholar] [CrossRef] [PubMed]
  46. Tan, Z.; Wu, Y.; Hong, H.; Yin, J.; Zhang, J.; Lin, L.; Wang, M.; Sun, X.; Sun, L.; Huang, Y.; et al. Two-Dimensional (C4H9NH3)2PbBr4 Perovskite Crystals for High-Performance Photodetector. J. Am. Chem. Soc. 2016, 138, 16612–16615. [Google Scholar] [CrossRef]
  47. Chen, Z.; Kang, Z.; Rao, C.; Cheng, Y.; Liu, N.; Zhang, Z.; Li, L.; Gao, Y. Improving Performance of Hybrid Graphene–Perovskite Photodetector by a Scratch Channel. Adv. Electron. Mater. 2019, 5, 1900168. [Google Scholar] [CrossRef]
  48. Chang, P.H.; Liu, S.Y.; Lan, Y.B.; Tsai, Y.C.; You, X.Q.; Li, C.S.; Huang, K.Y.; Chou, A.S.; Cheng, T.C.; Wang, J.K.; et al. Ultrahigh Responsivity and Detectivity Graphene-Perovskite Hybrid Phototransistors by Sequential Vapor Deposition. Sci. Rep. 2017, 7, 46281. [Google Scholar] [CrossRef]
  49. Zou, Y.; Zou, T.; Zhao, C.; Wang, B.; Xing, J.; Yu, Z.; Cheng, J.; Xin, W.; Yang, J.; Yu, W.; et al. A Highly Sensitive Single Crystal Perovskite-Graphene Hybrid Vertical Photodetector. Small 2020, 16, e2000733. [Google Scholar] [CrossRef]
  50. Shao, Y.; Liu, Y.; Chen, X.; Chen, C.; Sarpkaya, I.; Chen, Z.; Fang, Y.; Kong, J.; Watanabe, K.; Taniguchi, T.J.N. Stable graphene-two-dimensional multiphase perovskite heterostructure phototransistors with high gain. Nano Lett. 2017, 17, 7330–7338. [Google Scholar] [CrossRef]
  51. Sun, Z.; Aigouy, L.; Chen, Z.J.N. Plasmonic-enhanced perovskite–graphene hybrid photodetectors. Nanoscale 2016, 8, 7377–7383. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, H.; Lim, J.W.; Quan, L.N.; Chung, K.; Jang, Y.J.; Ma, Y.; Kim, D.H. Perovskite-Gold Nanorod Hybrid Photodetector with High Responsivity and Low Driving Voltage. Adv. Opt. Mater. 2018, 6, 1701397. [Google Scholar] [CrossRef]
  53. Liu, B.; Gutha, R.R.; Kattel, B.; Alamri, M.; Gong, M.; Sadeghi, S.M.; Chan, W.L.; Wu, J.Z. Using Silver Nanoparticles-Embedded Silica Meta films as Substrates to Enhance the Performance of Perovskite Photodetectors. ACS Appl. Mater. Interfaces 2019, 11, 32301–32309. [Google Scholar] [CrossRef] [PubMed]
  54. Feng, F.; Wang, T.; Qiao, J.; Min, C.; Yuan, X.; Somekh, M. Plasmonic and Graphene-Functionalized High-Performance Broadband Quasi-Two-Dimensional Perovskite Hybrid Photodetectors. ACS Appl. Mater. Interfaces 2021, 13, 61496–61505. [Google Scholar] [CrossRef]
  55. Xie, C.; Yan, F. Perovskite/poly (3-hexylthiophene)/graphene multi heterojunction phototransistors with ultrahigh gain in broadband wavelength region. ACS Appl. Mater. Interfaces 2017, 9, 1569–1576. [Google Scholar] [CrossRef]
  56. Qin, L.; Wu, L.; Kattel, B.; Li, C.; Zhang, Y.; Hou, Y.; Wu, J.; Chan, W.L. Using bulk heterojunctions and selective electron trapping to enhance the responsivity of perovskite–graphene photodetectors. Adv. Funct. Mater. 2017, 27, 1704173. [Google Scholar] [CrossRef]
  57. Tian, H.; Chin, M.L.; Najmaei, S.; Guo, Q.; Xia, F.; Wang, H.; Dubey, M. Optoelectronic devices based on two-dimensional transition metal dichalcogenides. Nano Res. 2016, 9, 1543–1560. [Google Scholar] [CrossRef]
  58. Kang, D.H.; Pae, S.R.; Shim, J.; Yoo, G.; Jeon, J.; Leem, J.W.; Yu, J.S.; Lee, S.; Shin, B.; Park, J.H. An Ultrahigh-Performance Photodetector based on a Perovskite-Transition-Metal-Dichalcogenide Hybrid Structure. Adv. Mater. 2016, 28, 7799–7806. [Google Scholar] [CrossRef]
  59. Sun, B.; Xi, S.; Liu, Z.; Liu, X.; Wang, Z.; Tan, X.; Shi, T.; Zhou, J.; Liao, G. Sensitive, fast, and stable photodetector based on perovskite/MoS2 hybrid film. Appl. Surf. Sci. 2019, 493, 389–395. [Google Scholar] [CrossRef]
  60. Zhang, Z.X.; Long-Hui, Z.; Tong, X.W.; Gao, Y.; Xie, C.; Tsang, Y.H.; Luo, L.B.; Wu, Y.C. Ultrafast, Self-Driven, and Air-Stable Photodetectors Based on Multilayer PtSe2/Perovskite Heterojunctions. J. Phys. Chem. Lett. 2018, 9, 1185–1194. [Google Scholar] [CrossRef]
  61. Zhuang, H.L.; Hennig, R.G. Computational search for single-layer transition-metal dichalcogenide photocatalysts. J. Phys. Chem. C 2013, 117, 20440–20445. [Google Scholar] [CrossRef]
  62. Zhang, L.; Shen, S.; Li, M.; Li, L.; Zhang, J.; Fan, L.; Cheng, F.; Li, C.; Zhu, M.; Kang, Z.; et al. Strategies for Air-Stable and Tunable Monolayer MoS2-Based Hybrid Photodetectors with High Performance by Regulating the Fully Inorganic Trihalide Perovskite Nanocrystals. Adv. Opt. Mater. 2019, 7, 1801744. [Google Scholar] [CrossRef]
  63. Tian, X.; Zhang, Y.; Zheng, R.; Wei, D.; Liu, J. Two-dimensional organic–inorganic hybrid Ruddlesden–Popper perovskite materials: Preparation, enhanced stability, and applications in photodetection. Sustain. Energy Fuels 2020, 4, 2087–2113. [Google Scholar] [CrossRef]
  64. Roghabadi, F.A.; Alidaei, M.; Mousavi, S.M.; Ashjari, T.; Tehrani, A.S.; Ahmadi, V.; Sadrameli, S.M. Stability progress of perovskite solar cells dependent on the crystalline structure: From 3D ABX3 to 2D Ruddlesden–Popper perovskite absorbers. J. Mater. Chem. A 2019, 7, 5898–5933. [Google Scholar] [CrossRef]
  65. Fang, C.; Wang, H.; Shen, Z.; Shen, H.; Wang, S.; Ma, J.; Wang, J.; Luo, H.; Li, D. High-Performance Photodetectors Based on Lead-Free 2D Ruddlesden-Popper Perovskite/MoS2 Heterostructures. ACS Appl. Mater. Interfaces 2019, 11, 8419–8427. [Google Scholar] [CrossRef]
  66. Wang, Q.; Zhang, Q.; Luo, X.; Wang, J.; Zhu, R.; Liang, Q.; Zhang, L.; Yong, J.Z.; Yu Wong, C.P.; Eda, G.; et al. Optoelectronic Properties of a van der Waals WS2 Monolayer/2D Perovskite Vertical Heterostructure. ACS Appl. Mater. Interfaces 2020, 12, 45235–45242. [Google Scholar] [CrossRef]
  67. Zhang, K.; Zhang, T.; Cheng, G.; Li, T.; Wang, S.; Wei, W.; Zhou, X.; Yu, W.; Sun, Y.; Wang, P. Interlayer transition and infrared photodetection in atomically thin type-II MoTe2/MoS2 van der Waals heterostructures. ACS Nano 2016, 10, 3852–3858. [Google Scholar] [CrossRef]
  68. Zhou, H.; Lai, H.; Sun, X.; Zhang, N.; Wang, Y.; Liu, P.; Zhou, Y.; Xie, W. Van der Waals MoS2/Two-Dimensional Perovskite Heterostructure for Sensitive and Ultrafast Sub-Band-Gap Photodetection. ACS Appl. Mater. Interfaces 2022, 14, 3356–3362. [Google Scholar] [CrossRef]
  69. Cherusseri, J.; Varma, S.J.; Pradhan, B.; Li, J.; Kumar, J.; Barrios, E.; Amin, M.Z.; Towers, A.; Gesquiere, A.; Thomas, J. Synthesis of air-stable two-dimensional nanoplatelets of Ruddlesden–Popper organic–inorganic hybrid perovskites. Nanoscale 2020, 12, 10072–10081. [Google Scholar] [CrossRef]
  70. Xu, Y.; Shi, Z.; Shi, X.; Zhang, K.; Zhang, H. Recent progress in black phosphorus and black-phosphorus-analogue materials: Properties, synthesis and applications. Nanoscale 2019, 11, 14491–14527. [Google Scholar] [CrossRef]
  71. Zou, X.; Li, Y.; Tang, G.; You, P.; Yan, F. Schottky Barrier-Controlled Black Phosphorus/Perovskite Phototransistors with Ultrahigh Sensitivity and Fast Response. Small 2019, 15, e1901004. [Google Scholar] [CrossRef]
  72. Wang, L.; Zou, X.; Lin, J.; Jiang, J.; Liu, Y.; Liu, X.; Zhao, X.; Liu, Y.F.; Ho, J.C.; Liao, L. Perovskite/Black Phosphorus/MoS2 Photogate Reversed Photodiodes with Ultrahigh Light On/Off Ratio and Fast Response. ACS Nano 2019, 13, 4804–4813. [Google Scholar] [CrossRef] [PubMed]
  73. Deng, W.; Huang, H.; Jin, H.; Li, W.; Chu, X.; Xiong, D.; Yan, W.; Chun, F.; Xie, M.; Luo, C.; et al. All-Sprayed-Processable, Large-Area, and Flexible Perovskite/MXene-Based Photodetector Arrays for Photocommunication. Adv. Opt. Mater. 2019, 7, 1801521. [Google Scholar] [CrossRef]
  74. Hasan, M.M.; Hossain, M.M.; Chowdhury, H.K. Two-dimensional MXene-based flexible nanostructures for functional nanodevices: A review. J. Mater. Chem. A 2021, 9, 3231–3269. [Google Scholar] [CrossRef]
  75. Abdolhossein Zadeh, S.; Jiang, X.; Zhang, H.; Qiu, J.; Zhang, C.J. Perspectives on solution processing of two-dimensional MXenes. Mater. Today 2021, 48, 214–240. [Google Scholar] [CrossRef]
  76. Le, F.; Brandl, D.W.; Urzhumov, Y.A.; Wang, H.; Kundu, J.; Halas, N.J.; Aizpurua, J.; Nordlander, P. Metallic nanoparticle arrays: A common substrate for both surface-enhanced Raman scattering and surface-enhanced infrared absorption. ACS Nano 2008, 2, 707–718. [Google Scholar] [CrossRef]
  77. Li, H.; Li, Z.; Liu, S.; Li, M.; Wen, X.; Lee, J.; Lin, S.; Li, M.-Y.; Lu, H. High performance hybrid MXene nanosheet/CsPbBr3 quantum dot photodetectors with an excellent stability. J. Alloys Compd. 2022, 895, 162570. [Google Scholar] [CrossRef]
  78. Liu, Y.; Elbanna, A.; Gao, W.; Pan, J.; Shen, Z.; Teng, J. Interlayer excitons in transition metal dichalcogenide semiconductors for 2D optoelectronics. Adv. Mater. 2022, 34, 2107138. [Google Scholar] [CrossRef]
  79. Jin, L.C.; Xiao, Y.; Zhang, D.A.; Zhang, H.W.; Tang, X.L.; Zhong, Z.Y.; Yang, Q.H. Giant optical absorption and low dark current characteristics in wrinkled single layer graphene/bismuth nanorods heterostructures. Carbon. 2018, 127, 596–601. [Google Scholar] [CrossRef]
  80. Calvo, M.E. Materials chemistry approaches to the control of the optical features of perovskite solar cells. J. Mater. Chem. A 2017, 5, 20561–20578. [Google Scholar] [CrossRef]
  81. Bati, A.S.; Batmunkh, M.; Shapter, J.G. Emerging 2D layered materials for perovskite solar cells. Adv. Energy Mater. 2020, 10, 1902253. [Google Scholar] [CrossRef]
  82. Huang, X.; Yin, Z.; Wu, S.; Qi, X.; He, Q.; Zhang, Q.; Yan, Q.; Boey, F.; Zhang, H.J. Graphene-based materials: Synthesis, characterization, properties, and applications. Small 2011, 7, 1876–1902. [Google Scholar] [CrossRef] [PubMed]
  83. Cho, A.N.; Park, N.G. Impact of interfacial layers in perovskite solar cells. ChemSusChem 2017, 10, 3687–3704. [Google Scholar] [CrossRef]
  84. Xiang, W.; Liu, S.; Tress, W. Interfaces and interfacial layers in inorganic perovskite solar cells. Angew. Chem. 2021, 133, 26644–26657. [Google Scholar] [CrossRef]
  85. Zhu, Z.; Ma, J.; Wang, Z.; Mu, C.; Fan, Z.; Du, L.; Bai, Y.; Fan, L.; Yan, H.; Phillips, D.L.; et al. Efficiency enhancement of perovskite solar cells through fast electron extraction: The role of graphene quantum dots. J. Am. Chem. Soc. 2014, 136, 3760–3763. [Google Scholar] [CrossRef]
  86. Agresti, A.; Pescetelli, S.; Palma, A.L.; Del Rio Castillo, A.E.; Konios, D.; Kakavelakis, G.; Razza, S.; Cinà, L.; Kymakis, E.; Bonaccorso, F.; et al. Graphene Interface Engineering for Perovskite Solar Modules: 12.6% Power Conversion Efficiency over 50 cm2 Active Area. ACS Energy Lett. 2017, 2, 279–287. [Google Scholar] [CrossRef]
  87. Wang, Y.; Wu, T.; Barbaud, J.; Kong, W.; Cui, D.; Chen, H.; Yang, X.; Han, L. Stabilizing heterostructures of soft perovskite semiconductors. Science 2019, 365, 687–691. [Google Scholar] [CrossRef]
  88. Zhao, E.; Gao, L.; Yang, S.; Wang, L.; Cao, J.; Ma, T. In situ fabrication of 2D SnS2 nanosheets as a new electron transport layer for perovskite solar cells. Nano Res. 2018, 11, 5913–5923. [Google Scholar] [CrossRef]
  89. Chu, W.; Li, X.; Li, S.; Hou, J.; Jiang, Q.; Yang, J. High-Performance Flexible Perovskite Solar Cells with a Metal Sulfide Electron Transport Layer of SnS2 by Room-Temperature Vacuum Deposition. ACS Appl. Energy Mater. 2018, 2, 382–388. [Google Scholar] [CrossRef]
  90. Zhao, X.; Liu, S.; Zhang, H.; Chang, S.-Y.; Huang, W.; Zhu, B.; Shen, Y.; Shen, C.; Wang, D.; Yang, Y.; et al. 20% Efficient Perovskite Solar Cells with 2D Electron Transporting Layer. Adv. Funct. Mater. 2019, 29, 1805168. [Google Scholar] [CrossRef]
  91. Wang, D.; Elumalai, N.K.; Mahmud, M.A.; Yi, H.; Upama, M.B.; Lee Chin, R.A.; Conibeer, G.; Xu, C.; Haque, F.; Duan, L.; et al. MoS2 incorporated hybrid hole transport layer for high performance and stable perovskite solar cells. Synth. Met. 2018, 246, 195–203. [Google Scholar] [CrossRef]
  92. Kung, P.K.; Li, M.H.; Lin, P.Y.; Chiang, Y.H.; Chan, C.R.; Guo, T.F.; Chen, P. A review of inorganic hole transport materials for perovskite solar cells. Adv. Mater. Interfaces 2018, 5, 1800882. [Google Scholar] [CrossRef]
  93. Wang, S.; Jiang, Y.; Juarez-Perez, E.J.; Ono, L.K.; Qi, Y. Accelerated degradation of methylammonium lead iodide perovskites induced by exposure to iodine vapour. Nat. Energy 2016, 2, 1–8. [Google Scholar] [CrossRef]
  94. Capasso, A.; Matteocci, F.; Najafi, L.; Prato, M.; Buha, J.; Cinà, L.; Pellegrini, V.; Carlo, A.D.; Bonaccorso, F. Few-Layer MoS2 Flakes as Active Buffer Layer for Stable Perovskite Solar Cells. Adv. Energy Mater. 2016, 6. [Google Scholar] [CrossRef]
  95. Muduli, S.K.; Varrla, E.; Kulkarni, S.A.; Han, G.; Thirumal, K.; Lev, O.; Mhaisalkar, S.; Mathews, N. 2D black phosphorous nanosheets as a hole transporting material in perovskite solar cells. J. Power Sources 2017, 371, 156–161. [Google Scholar] [CrossRef]
  96. Wang, Z.; Zhang, Z.; Xie, L.; Wang, S.; Yang, C.; Fang, C.; Hao, F. Recent advances and perspectives of photostability for halide perovskite solar cells. Adv. Opt. Mater. 2022, 10, 2101822. [Google Scholar] [CrossRef]
  97. Wang, Y.; Zhang, H.; Zhang, T.; Shi, W.; Kan, M.; Chen, J.; Zhao, Y. Photostability of MAPbI3 perovskite solar cells by incorporating black phosphorus. Sol. RRL 2019, 3, 1900197. [Google Scholar] [CrossRef]
  98. Yang, L.; Wang, B.; Dall’Agnese, C.; Dall’Agnese, Y.; Chen, G.; Jena, A.K.; Wang, X.-F.; Miyasaka, T. Engineering, Hybridization of SnO2 and an In-Situ-Oxidized Ti3C2Tx MXene Electron Transport Bilayer for High-Performance Planar Perovskite Solar Cells. ACS Sustain. Chem. Eng. 2021, 9, 13672–13680. [Google Scholar] [CrossRef]
  99. Chen, T.; Tong, G.; Xu, E.; Li, H.; Li, P.; Zhu, Z.; Tang, J.; Qi, Y.; Jiang, Y.J. Accelerating hole extraction by inserting 2D Ti3C2-MXene interlayer to all inorganic perovskite solar cells with long-term stability. J. Mater. Chem. A 2019, 7, 20597–20603. [Google Scholar] [CrossRef]
  100. Yang, L.; Dall’Agnese, C.; Dall’Agnese, Y.; Chen, G.; Gao, Y.; Sanehira, Y.; Jena, A.K.; Wang, X.F.; Gogotsi, Y.; Miyasaka, T. Surface-modified metallic Ti3C2Tx MXene as electron transport layer for planar heterojunction perovskite solar cells. Adv. Funct. Mater. 2019, 29, 1905694. [Google Scholar] [CrossRef]
  101. Lin, L.; Jones, T.W.; Yang, T.C.-J.; Duffy, N.W.; Li, J.; Zhao, L.; Chi, B.; Wang, X.; Wilson, G.J. Inorganic electron transport materials in perovskite solar cells. Adv. Funct. Mater. 2021, 31, 2008300. [Google Scholar] [CrossRef]
  102. Sherkar, T.S.; Momblona, C.; Gil-Escrig, L.; Avila, J.; Sessolo, M.; Bolink, H.J.; Koster, L.J.A. Recombination in perovskite solar cells: Significance of grain boundaries, interface traps, and defect ions. ACS Energy Lett. 2017, 2, 1214–1222. [Google Scholar] [CrossRef]
  103. Chueh, C.-C.; Li, C.-Z.; Jen, A.K.Y. Recent progress and perspective in solution-processed Interfacial materials for efficient and stable polymer and organometal perovskite solar cells. Energy Environ. Sci. 2015, 8, 1160–1189. [Google Scholar] [CrossRef]
  104. Xiao, M.; Hao, M.; Lyu, M.; Moore, E.G.; Zhang, C.; Luo, B.; Hou, J.; Lipton-Duffin, J.; Wang, L.J. Surface ligands stabilized lead halide perovskite quantum dot photocatalyst for visible light-driven hydrogen generation. Adv. Funct. Mater. 2019, 29, 1905683. [Google Scholar] [CrossRef]
  105. Chen, S.; Takata, T.; Domen, K. Particulate photocatalysts for overall water splitting. Nat. Rev. Mater. 2017, 2, 1–17. [Google Scholar] [CrossRef]
  106. Spasiano, D.; Marotta, R.; Malato, S.; Fernandez-Ibanez, P.; Di Somma, I. Solar photocatalysis: Materials, reactors, some commercial, and pre-industrialized applications. A comprehensive approach. Appl. Catal. B Environ. 2015, 170, 90–123. [Google Scholar] [CrossRef]
  107. Tong, H.; Ouyang, S.; Bi, Y.; Umezawa, N.; Oshikiri, M.; Ye, J. Nano-photocatalytic materials: Possibilities and challenges. Adv. Mater. 2012, 24, 229–251. [Google Scholar] [CrossRef]
  108. Li, H.; Chen, Z.-H.; Zhao, L.; Yang, G.-D. Synthesis of TiO2@ZnIn2S4 hollow nanospheres with enhanced photocatalytic hydrogen evolution. Rare Met. 2019, 38, 420–427. [Google Scholar] [CrossRef]
  109. Adinolfi, V.; Peng, W.; Walters, G.; Bakr, O.M.; Sargent, E.H. The electrical and optical properties of organometal halide perovskites relevant to optoelectronic performance. Adv. Mater. 2018, 30, 1700764. [Google Scholar] [CrossRef]
  110. Tasleem, S.; Tahir, M. Current trends in strategies to improve photocatalytic performance of perovskites materials for solar to hydrogen production. Renew. Sustain. Energy Rev. 2020, 132, 110073. [Google Scholar] [CrossRef]
  111. Park, S.; Chang, W.J.; Lee, C.W.; Park, S.; Ahn, H.-Y.; Nam, K.T. Photocatalytic hydrogen generation from hydriodic acid using methylammonium lead iodide in dynamic equilibrium with aqueous solution. Nat. Energy 2016, 2, 1–8. [Google Scholar] [CrossRef]
  112. Yang, N.; Liu, Y.; Wen, H.; Tang, Z.; Zhao, H.; Li, Y.; Wang, D. Photocatalytic properties of graphdiyne and graphene modified TiO2: From theory to experiment. Acs Nano 2013, 7, 1504–1512. [Google Scholar] [CrossRef] [PubMed]
  113. Guan, W.; Li, Y.; Zhong, Q.; Liu, H.; Chen, J.; Hu, H.; Lv, K.; Gong, J.; Xu, Y.; Kang, Z.; et al. Fabricating MAPbI3/MoS2 Composites for Improved Photocatalytic Performance. Nano Lett. 2021, 21, 597–604. [Google Scholar] [CrossRef] [PubMed]
  114. Xu, Y.F.; Yang, M.Z.; Chen, B.X.; Wang, X.D.; Chen, H.Y.; Kuang, D.B.; Su, C.Y. A CsPbBr3 Perovskite Quantum Dot/Graphene Oxide Composite for Photocatalytic CO2 Reduction. J. Am. Chem. Soc. 2017, 139, 5660–5663. [Google Scholar] [CrossRef]
  115. Wang, T.; Yue, D.; Li, X.; Zhao, Y. Lead-free double perovskite Cs2AgBiBr6/RGO composite for efficient visible light photocatalytic H2 evolution. Appl. Catal. B Environ. 2020, 268, 118399. [Google Scholar] [CrossRef]
  116. Li, R.; Li, X.; Wu, J.; Lv, X.; Zheng, Y.-Z.; Zhao, Z.; Ding, X.; Tao, X.; Chen, J.-F. Few-layer black phosphorus-on-MAPbI3 for superb visible-light photocatalytic hydrogen evolution from HI splitting. Appl. Catal. B Environ. 2019, 259, 118075. [Google Scholar] [CrossRef]
  117. Wu, C.; Zhang, J.; Tong, X.; Yu, P.; Xu, J.Y.; Wu, J.; Wang, Z.M.; Lou, J.; Chueh, Y.L. A critical review on enhancement of photocatalytic hydrogen production by molybdenum disulfide: From growth to interfacial activities. Small 2019, 15, 1900578. [Google Scholar] [CrossRef] [PubMed]
  118. Wang, F.; Liu, X.; Zhang, Z.; Min, S. A noble-metal-free MoS2 nanosheet-coupled MAPbI3 photocatalyst for efficient and stable visible-light-driven hydrogen evolution. Chem. Commun. 2020, 56, 3281–3284. [Google Scholar] [CrossRef] [PubMed]
  119. Zhao, X.; Chen, S.; Yin, H.; Jiang, S.; Zhao, K.; Kang, J.; Liu, P.F.; Jiang, L.; Zhu, Z.; Cui, D.; et al. Perovskite Microcrystals with Intercalated Monolayer MoS2 Nanosheets as Advanced Photocatalyst for Solar-Powered Hydrogen Generation. Matter 2020, 3, 935–949. [Google Scholar] [CrossRef]
  120. Ulmer, U.; Dingle, T.; Duchesne, P.N.; Morris, R.H.; Tavasoli, A.; Wood, T.; Ozin, G.A. Fundamentals and applications of photocatalytic CO2 methanation. Nat. Commun. 2019, 10, 1–12. [Google Scholar] [CrossRef]
  121. Wu, H.L.; Li, X.B.; Tung, C.H.; Wu, L.Z. Semiconductor quantum dots: An emerging candidate for CO2 photoreduction. Adv. Mater. 2019, 31, 1900709. [Google Scholar] [CrossRef] [PubMed]
  122. Eigler, S.; Dotzer, C.; Hirsch, A. Visualization of defect densities in reduced graphene oxide. Carbon 2012, 50, 3666–3673. [Google Scholar] [CrossRef]
  123. Wang, X.; Li, K.; He, J.; Yang, J.; Dong, F.; Mai, W.; Zhu, M. Defect in reduced graphene oxide tailored selectivity of photocatalytic CO2 reduction on Cs4PbBr6 pervoskite hole-in-microdisk structure. Nano Energy 2020, 78, 105388. [Google Scholar] [CrossRef]
  124. Wang, X.D.; Huang, Y.H.; Liao, J.F.; Jiang, Y.; Zhou, L.; Zhang, X.Y.; Chen, H.Y.; Kuang, D.B. In Situ Construction of a Cs2SnI6 Perovskite Nanocrystal/SnS2 Nanosheet Heterojunction with Boosted Interfacial Charge Transfer. J. Am. Chem. Soc. 2019, 141, 13434–13441. [Google Scholar] [CrossRef]
  125. Shen, Z.-K.; Yuan, Y.-J.; Pei, L.; Yu, Z.-T.; Zou, Z. Black phosphorus photocatalysts for photocatalytic H2 generation: A review. Chem. Eng. J. 2020, 386, 123997. [Google Scholar] [CrossRef]
  126. Wang, X.; He, J.; Li, J.; Lu, G.; Dong, F.; Majima, T.; Zhu, M. Immobilizing perovskite CsPbBr3 nanocrystals on Black phosphorus nanosheets for boosting charge separation and photocatalytic CO2 reduction. Appl. Catal. B Environ. 2020, 277, 119230. [Google Scholar] [CrossRef]
  127. Pan, A.; Ma, X.; Huang, S.; Wu, Y.; Jia, M.; Shi, Y.; Liu, Y.; Wangyang, P.; He, L.; Liu, Y. CsPbBr3 Perovskite Nanocrystal Grown on MXene Nanosheets for Enhanced Photoelectric Detection and Photocatalytic CO2 Reduction. J. Phys. Chem. Lett. 2019, 10, 6590–6597. [Google Scholar] [CrossRef]
  128. Boyd, C.C.; Cheacharoen, R.; Leijtens, T.; McGehee, M.D. Understanding degradation mechanisms and improving stability of perovskite photovoltaics. Chem. Rev. 2018, 119, 3418–3451. [Google Scholar] [CrossRef]
  129. Zhang, Z.; Qiao, L.; Meng, K.; Long, R.; Chen, G.; Gao, P. Rationalization of passivation strategies toward high-performance perovskite solar cells. Chem. Soc. Rev. 2023, 52, 163–195. [Google Scholar] [CrossRef]
  130. Kim, J.H.; Oh, C.M.; Hwang, I.W.; Kim, J.; Lee, C.; Kwon, S.; Ki, T.; Lee, S.; Kang, H.; Kim, H. Efficient and Stable Quasi-2D Ruddlesden–Popper Perovskite Solar Cells by Tailoring Crystal Orientation and Passivating Surface Defects. Adv. Mater. 2023, 35, 2302143. [Google Scholar] [CrossRef]
  131. Xu, Z.; Lu, D.; Dong, X.; Chen, M.; Fu, Q.; Liu, Y. Highly Efficient and Stable Dion−Jacobson Perovskite Solar Cells Enabled by Extended π-Conjugation of Organic Spacer. Adv. Mater. 2021, 33, 2105083. [Google Scholar] [CrossRef] [PubMed]
  132. Wu, X.-J.; Chen, J.; Tan, C.; Zhu, Y.; Han, Y.; Zhang, H. Controlled growth of high-density CdS and CdSe nanorod arrays on selective facets of two-dimensional semiconductor nanoplates. Nat. Chem. 2016, 8, 470–475. [Google Scholar] [CrossRef] [PubMed]
  133. Huang, X.; Zeng, Z.; Bao, S.; Wang, M.; Qi, X.; Fan, Z.; Zhang, H. Solution-phase epitaxial growth of noble metal nanostructures on dispersible single-layer molybdenum disulfide nanosheets. Nat. Commun. 2013, 4, 1444. [Google Scholar] [CrossRef] [PubMed]
  134. Ke, W.; Kanatzidis, M.G. Prospects for low-toxicity lead-free perovskite solar cells. Nat. Commun. 2019, 10, 965. [Google Scholar] [CrossRef]
Figure 1. False-colored SEM images of (a) CH3NH3PbI3/graphene, (b) CH3NH3PbI3/MoS2, and (c) CH3NH3PbI3/h-BN heterostructures, respectively [26]. (d) Schematic illustration of the synthesis of MoSe2/CsPbBr3 [28].
Figure 1. False-colored SEM images of (a) CH3NH3PbI3/graphene, (b) CH3NH3PbI3/MoS2, and (c) CH3NH3PbI3/h-BN heterostructures, respectively [26]. (d) Schematic illustration of the synthesis of MoSe2/CsPbBr3 [28].
Crystals 13 01566 g001
Figure 2. (a) Schematic of the perovskite/graphene/Au NPs PD. (b) Transfer characteristics of the CH3NH3PbI3/graphene/Au NPs PDs. (c) The generation, diffusion, and transfer of photo-induced carriers in the perovskite layer. (d) The relationships between photo-responsivities and light intensity, as well as external quantum efficiency and light intensity, for both perovskite/graphene and perovskite/graphene/Au NPs devices [51]. (e) Schematic of the graphene/Au/perovskite hybrid-structure PDs, with the SEM images of Au square nanoarray plasmonic substrates. (f) Schematic of the graphene/Au/perovskite hybrid-structure PDs. (g) The responsivity of the device [54].
Figure 2. (a) Schematic of the perovskite/graphene/Au NPs PD. (b) Transfer characteristics of the CH3NH3PbI3/graphene/Au NPs PDs. (c) The generation, diffusion, and transfer of photo-induced carriers in the perovskite layer. (d) The relationships between photo-responsivities and light intensity, as well as external quantum efficiency and light intensity, for both perovskite/graphene and perovskite/graphene/Au NPs devices [51]. (e) Schematic of the graphene/Au/perovskite hybrid-structure PDs, with the SEM images of Au square nanoarray plasmonic substrates. (f) Schematic of the graphene/Au/perovskite hybrid-structure PDs. (g) The responsivity of the device [54].
Crystals 13 01566 g002
Figure 3. (a) Device structure and energy level diagram of MAPbI3/MoS2 PD. (b) APTES doping process. (c) Responsivity and detectivity versus illumination intensity [58]. (d) Schematic illustration of the PtSe2/perovskite heterojunction PD. (e) Energy band diagram of the PtSe2/FA0.85Cs0.15PbI3 perovskite heterojunction under light illumination at zero bias. (f) Steady−state PL of the perovskite and PtSe2/perovskite hybrid. (g) A single normalized cycle of the photoresponse for rise and fall times [60].
Figure 3. (a) Device structure and energy level diagram of MAPbI3/MoS2 PD. (b) APTES doping process. (c) Responsivity and detectivity versus illumination intensity [58]. (d) Schematic illustration of the PtSe2/perovskite heterojunction PD. (e) Energy band diagram of the PtSe2/FA0.85Cs0.15PbI3 perovskite heterojunction under light illumination at zero bias. (f) Steady−state PL of the perovskite and PtSe2/perovskite hybrid. (g) A single normalized cycle of the photoresponse for rise and fall times [60].
Crystals 13 01566 g003
Figure 4. (a) Cross−section of the vertical heterostructure device. (b) Electronic band alignment for a WS2 monolayer and PEPI monolayer. (c) Comparison of the incident laser power dependence of the photocurrent. (d) Comparison of the incident laser power dependence of the photoresponsivity [66]. (e) Schematic diagram of the PD with the MoS2/2DRP heterostructure. (f) Schematic illustration of the interband excitation process in the MoS2/2DRP vdW heterostructure. (g) PL spectra between individual components and the heterojunction area. (h) Photodetection performance of the PD at λ = 1550 nm [67].
Figure 4. (a) Cross−section of the vertical heterostructure device. (b) Electronic band alignment for a WS2 monolayer and PEPI monolayer. (c) Comparison of the incident laser power dependence of the photocurrent. (d) Comparison of the incident laser power dependence of the photoresponsivity [66]. (e) Schematic diagram of the PD with the MoS2/2DRP heterostructure. (f) Schematic illustration of the interband excitation process in the MoS2/2DRP vdW heterostructure. (g) PL spectra between individual components and the heterojunction area. (h) Photodetection performance of the PD at λ = 1550 nm [67].
Crystals 13 01566 g004
Figure 5. (a) Perovskite/MoS2/BP heterojunction device structure. (b) Band diagram and photocarriers transfer under laser illumination. (c) PL spectra for the MAPbI3 perovskite film and the various hybrid structures. (d) The Ids−Vds characteristics of devices under dark and light conditions [72]. (e) Schematic of the single PD with MXene electrode. (f) Under light illumination, photon-generated electron–hole pairs migrate from CsPbBr3 nanosheets to the MXene electrodes. (g) I−t curves of the PDs under the different bending angles [73].
Figure 5. (a) Perovskite/MoS2/BP heterojunction device structure. (b) Band diagram and photocarriers transfer under laser illumination. (c) PL spectra for the MAPbI3 perovskite film and the various hybrid structures. (d) The Ids−Vds characteristics of devices under dark and light conditions [72]. (e) Schematic of the single PD with MXene electrode. (f) Under light illumination, photon-generated electron–hole pairs migrate from CsPbBr3 nanosheets to the MXene electrodes. (g) I−t curves of the PDs under the different bending angles [73].
Crystals 13 01566 g005
Figure 6. (a) Cross−sectional SEM image of CH3NH3PbI3/GQDs/TiO2 cell. (b) Schematic illustration of electron generation and extraction at PT and PGT interfaces. (c) J−V curves for the best-performing solar cells [85].
Figure 6. (a) Cross−sectional SEM image of CH3NH3PbI3/GQDs/TiO2 cell. (b) Schematic illustration of electron generation and extraction at PT and PGT interfaces. (c) J−V curves for the best-performing solar cells [85].
Crystals 13 01566 g006
Figure 7. (a) Diagram of PEN/ITO/ETL/MAPbI3/Spiro/Au device. (b) Steady−state PCE and photocurrent density at maximum power point as a function of time for 60 nm SnS2. (c) Stability of the flexible PKSCs with and without SnS2 as ETL, respectively [89]. (d) Schematic diagram of cell with MoS2 blended PEDOT:PSS HTL. (e) The J−V curves of different concentration of MoS2 blended PEDOT:PSS and pristine PEDOT:PSS HTL based device. (f) The stabilized J−V and efficiency curves tracked at their maximum power point for different samples with different MoS2 concentrations in PEDOT:PSS HTL [91].
Figure 7. (a) Diagram of PEN/ITO/ETL/MAPbI3/Spiro/Au device. (b) Steady−state PCE and photocurrent density at maximum power point as a function of time for 60 nm SnS2. (c) Stability of the flexible PKSCs with and without SnS2 as ETL, respectively [89]. (d) Schematic diagram of cell with MoS2 blended PEDOT:PSS HTL. (e) The J−V curves of different concentration of MoS2 blended PEDOT:PSS and pristine PEDOT:PSS HTL based device. (f) The stabilized J−V and efficiency curves tracked at their maximum power point for different samples with different MoS2 concentrations in PEDOT:PSS HTL [91].
Crystals 13 01566 g007
Figure 8. (a) The SEM image of the cross-section MAPbI3/BP−based device. Scale bars represent 500 nm. (b) Photostability under continuous white light LED illumination in a N2 glovebox. (c) The J−V characteristics of PSCs based on MAPbI3 with and without the incorporation of BP [97]. (d) Device architecture of a PSC. (e) Box diagrams of PCE obtained from 18 PSCs based on SnO2 and O−Ti3C2Tx/SnO2. (f) Stability results in air without encapsulation at room temperature [98]. (g) Schematic of CsPbBr3/Ti3C2−MXene based solar cell. (h) J−V curves. (i) Steady power output curves of devices with and without Ti3C2− MXene [99].
Figure 8. (a) The SEM image of the cross-section MAPbI3/BP−based device. Scale bars represent 500 nm. (b) Photostability under continuous white light LED illumination in a N2 glovebox. (c) The J−V characteristics of PSCs based on MAPbI3 with and without the incorporation of BP [97]. (d) Device architecture of a PSC. (e) Box diagrams of PCE obtained from 18 PSCs based on SnO2 and O−Ti3C2Tx/SnO2. (f) Stability results in air without encapsulation at room temperature [98]. (g) Schematic of CsPbBr3/Ti3C2−MXene based solar cell. (h) J−V curves. (i) Steady power output curves of devices with and without Ti3C2− MXene [99].
Crystals 13 01566 g008
Figure 9. (a) Comparison of the H2 evolution activities of MAPbI3, MAPbI3/Pt, and MAPbI3/rGO. (b) H2 evolution activity of MAPbI3/rGO during 20 cycles. (c) Nyquist plots of MAPbI3 and MAPbI3/rGO measured at 0 V versus Ag/AgCl electrode under light irradiation [33]. (d) Schematic diagram of energy band alignment of MAPbI3/MoS2 composites and the mechanism of hydrogen evolution [113]. (e) Schematic diagram of CO2 photoreduction over the CsPbBr3 QD/GO photocatalyst. (f) PL spectra with an excitation wavelength of 369.6 nm. (g) EIS Nyquist plots recorded under 150 mW cm−2 illumination at a bias of −0.4 VAg/AgCl [114].
Figure 9. (a) Comparison of the H2 evolution activities of MAPbI3, MAPbI3/Pt, and MAPbI3/rGO. (b) H2 evolution activity of MAPbI3/rGO during 20 cycles. (c) Nyquist plots of MAPbI3 and MAPbI3/rGO measured at 0 V versus Ag/AgCl electrode under light irradiation [33]. (d) Schematic diagram of energy band alignment of MAPbI3/MoS2 composites and the mechanism of hydrogen evolution [113]. (e) Schematic diagram of CO2 photoreduction over the CsPbBr3 QD/GO photocatalyst. (f) PL spectra with an excitation wavelength of 369.6 nm. (g) EIS Nyquist plots recorded under 150 mW cm−2 illumination at a bias of −0.4 VAg/AgCl [114].
Crystals 13 01566 g009
Table 1. Summary of advantages and disadvantages of different fabrication techniques.
Table 1. Summary of advantages and disadvantages of different fabrication techniques.
Fabrication TechniquesAdvantagesDisadvantages
Solid-state methodsHigh-quality
High surface cleanness
Good thickness control
Immediate device fabrication
Complex fabrication processes
High consumption conditions
Small-scale production
Solution methodsFacile processing
Low temperature
Low energy consumption
Good scalability
Low quality and uniformity
Limited thickness control
Residual solvents and impurities
Table 2. Summary of the key parameters for PDs based on perovskite and 2D material heterostructures.
Table 2. Summary of the key parameters for PDs based on perovskite and 2D material heterostructures.
Device StructureR [AW−1]D* [Jones]On/Off RatioEQE [%]Response Time [trise,tfall.ms]Test ConditionsReference
CH3NH3PbI3/Graphene180109-5 × 10487, 540λ:520 nm, P:1 uw, VDS = 0.1 V, VGS = 0 V[45]
Graphene/(C4H9NH3)2PbBr4/Graphene2.1 × 103-103--λ:470 nm, P:10 uw, Vbias = 0.5 V[46]
Graphene/CH3NH3PbI3/Graphene22 × 10−3-2.6 × 103--λ:452 nm, P:2.1 mWcm−2, Vbias = 2 V[47]
MAPbBr3/Graphene1.017 × 1032.02 × 1013--50.9, 26λ:532 nm, P:200 mWcm−2, Vbias = 3 V[49]
CH3NH3PbI3/Graphene1.73 × 1072 × 1015-108-, 879White light, VDS = 1.5 V, VGS = 40 v[48]
graphene/Au nanoarray/(BA)2(FA)n−1PbnI3n+118.712.21 × 10136 × 104-0.33 × 103, 0.27 × 103λ:405 nm, P:10.4 mWcm−2, Vbias = 2 V[54]
CH3NH3PbI3/Graphene/Ag nanoparticle495.3---7 × 103, -λ:585 nm, P:1.27 uW, Vbias = 0.1 V[53]
CH3NH3PbI3/MoS221.1 × 1031.38 × 1010--6.17 × 103, 4.5 × 103λ:520 nm[58]
CH3NH3PbI3/MoS23421.14 × 1012--27, 21λ:520 nm, P:2.2 pW, Vbias = 2 V[59]
CH3NH3PbI3/WSe21102.2 × 1011-2.5 × 104143, 2113λ:532 nm, P:2.8 mW, Vbias = 2 V[27]
FACsPbI3/PtSe20.11772.91 × 10125.7 × 103-78 × 10−6, 60 × 10−6λ:808 nm, Vbias = 0 V[79]
CsPbBr3/MoS26403.38 × 1011≈1041.5 × 10592, -λ:532 nm, P:4.86 mWcm−2, VDS = 1 V[62]
Graphene/(PEA)2SnI4/MoS2/Graphene0.1218.09 × 10950038.234, 38λ:532 nm, P:5.46 nW, VDS = 0 V, VGS = 0 V[65]
Graphene/(C6H5C2H4NH3)2PbI4/WS2/Graphene24.2 × 10−6-1.5 × 1035.7 × 10−5200, 200λ:532 nm, Vbias = 0 V[66]
(BA)2MA3Pb4I13/MoS21214.3 × 1014--8.2 × 10−3, 28.6 × 10−3λ:860 nm, P:5 uw, Vbias = 2 V,[67]
MAPbI3-xClx/BP106–1089 × 1013--8, 17-[71]
MAPbI3/BP/MoS2111.3 × 10126 × 104800.15, 0.24λ:450 nm, Vbias = −2 V[72]
CsPbBr3/Ti3C2Tx44.9 × 10−36.4 × 1082.3 × 103-48, 18λ:450 nm, P:7.07 mWmm−2, Vbias = 10 V[73]
CsPbBr3/Ti3C2Tx nanosheet97 × 10−6---46.2, 24.6λ:490 nm, P:2.9 mWcm−2[77]
Table 3. Summary of the key parameters for solar cells based on perovskite and 2D material heterostructures.
Table 3. Summary of the key parameters for solar cells based on perovskite and 2D material heterostructures.
Device StructureVoc [V]Jsc [mA cm−2]FF PCE [%]Reference
FTO/TiO2/GQDS/CH3NH3PbI3/Spiro-OMeTAD/Au0.93717.060.63510.15[85]
FTO/TiO2/G+mTiO2/GO-Li/CH3NH3PbI3/spiro-OMeTAD/Au8.57-0.64612.6[86]
ITO/SnO2/[CH(NH2)2]x[CH3NH3]1−xPb1+yI3/Cl-GO/PTAA/AU1.1223.820.7921.08[87]
FTO/SnS2/CH3NH3PbI3/Spiro-OMeTAD/Au0.9523.700.6113.63[88]
FTO/SnS2/CH3NH3PbI3/Spiro-OMeTAD/Au1.01121.700.6012.20[89]
ITO/SnO2/SnS2/FA0.75MA0.15Cs0.1PbI2.65Br0.35/Spiro-OMeTAD/Au1.16123.550.7320.12[90]
FTO/TiO2/TiO2/CH3NH3PbI3/MoS2/Spiro-OMeTAD/Au0.943 ± 0.026−18.11 ± 0.650.628 ± 0.03410.7 ± 0.4[94]
ITO/PEDOT:PSS, MoS2/MAPbI3/PC71BM/BCP/Ag0.9522 ± 0.00172.3 ± 0.90.723 ± 0.00914.2 ± 0.39[91]
FTO/TiO2/Meso-TiO2/MAPbI3/BP, Spiro-OMeTAD/AU1.0620.220.76116.4[95]
FTO/TiO2/SnO2/CH3NH3PbI3,BP/Spiro-OMeTAD/Ag1.08223.310.822520.65[97]
ITO/Ti3C2Tx/CH3NH3PbI3/Spiro-OMeTAD/Ag1.0822.630.7017.17[100]
FTO/O-Ti3C2Tx/SnO2/CH3NH3PbI3/Spiro-OMeTAD/Ag1.07323.880.78420.09[98]
FTO/CsPbBr3/Ti3C2/Au1.4448.540.73089.01[99]
Table 4. Summary of the key parameters for photocatalysis based on perovskite and 2D material heterostructures.
Table 4. Summary of the key parameters for photocatalysis based on perovskite and 2D material heterostructures.
Heterostructure TypePhotoreaction TypeReaction SolutionActivityLight SourceReference
MAPbI3/RGOH2 productionHIH2 (93.9 µmolh−1)120 mWcm−2 visible-light (λ ≥ 420 nm)[33]
Cs2AgBiBr6/rGOH2 productionHBrH2 (498 µmolg−1)visible light irradiation (λ ≥ 420 nm, 300 W Xe lamp)[115]
MAPbI3/BPH2 productionHIH2 (3742 µmolh−1g−1)visible light (λ ≥ 420 nm)[116]
MAPbI3/MoS2H2 productionHIH2 (206.1 µmolh−1)visible light[118]
MAPbI3/MoS2H2 productionHIH2 (13,600 µmolh−1g−1)visible light (λ ≥ 420 nm)[119]
MAPbI3/MoS2H2 productionHIH2 (30,000 µmolh−1g−1)visible light (λ ≥ 420 nm)[113]
CsPbBr3 QD/GOCO2 reductionEthyl acetateCO (58.7 µmol g−1)
CH4 (29.6 µmol g−1)
H2 (1.58 µmol g−1)
100 W Xe lamp with an AM 1.5 G filter[114]
Cs4PbBr6/rGOCO2 reductionEthyl acetateCO (11.4 µmol g−1h−1)300 W Xe lamp equipped with 420 nm filter[123]
Cs2SnI6 NC/SnS2CO2 reduction-CO (6.09 µmol g−1)150 mWcm−2 Xe lamp with 400 nm filter[124]
CsPbBr3 NC/BPCO2 reductionEthyl acetateCO (44.7 µmol g−1h−1)
CH4 (10.7 µmol g−1h−1)
200 mWcm−2 Xe lamp[126]
CsPbBr3 NCs/Ti3C2TxCO2 reductionEthyl acetateCO (26.32 µmol g−1h−1)
CH4 (7.25 µmol g−1h−1)
300 W Xe lamp with a cut off filter >420 nm[127]
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

Feng, S.; Li, B.; Xu, B.; Wang, Z. Hybrid Perovskites and 2D Materials in Optoelectronic and Photocatalytic Applications. Crystals 2023, 13, 1566. https://doi.org/10.3390/cryst13111566

AMA Style

Feng S, Li B, Xu B, Wang Z. Hybrid Perovskites and 2D Materials in Optoelectronic and Photocatalytic Applications. Crystals. 2023; 13(11):1566. https://doi.org/10.3390/cryst13111566

Chicago/Turabian Style

Feng, Shuo, Benxuan Li, Bo Xu, and Zhuo Wang. 2023. "Hybrid Perovskites and 2D Materials in Optoelectronic and Photocatalytic Applications" Crystals 13, no. 11: 1566. https://doi.org/10.3390/cryst13111566

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