Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 6081 2023-08-15 16:20:27 |
2 format change + 4 word(s) 6085 2023-08-16 02:52:41 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Werlinger, F.; Segura, C.; Martínez, J.; Osorio-Roman, I.; Jara, D.; Yoon, S.J.; Gualdrón-Reyes, A.F. Development of Active Layers for Efficient Photovoltaic Devices. Encyclopedia. Available online: https://encyclopedia.pub/entry/48094 (accessed on 24 July 2024).
Werlinger F, Segura C, Martínez J, Osorio-Roman I, Jara D, Yoon SJ, et al. Development of Active Layers for Efficient Photovoltaic Devices. Encyclopedia. Available at: https://encyclopedia.pub/entry/48094. Accessed July 24, 2024.
Werlinger, Francisca, Camilo Segura, Javier Martínez, Igor Osorio-Roman, Danilo Jara, Seog Joon Yoon, Andrés Fabián Gualdrón-Reyes. "Development of Active Layers for Efficient Photovoltaic Devices" Encyclopedia, https://encyclopedia.pub/entry/48094 (accessed July 24, 2024).
Werlinger, F., Segura, C., Martínez, J., Osorio-Roman, I., Jara, D., Yoon, S.J., & Gualdrón-Reyes, A.F. (2023, August 15). Development of Active Layers for Efficient Photovoltaic Devices. In Encyclopedia. https://encyclopedia.pub/entry/48094
Werlinger, Francisca, et al. "Development of Active Layers for Efficient Photovoltaic Devices." Encyclopedia. Web. 15 August, 2023.
Development of Active Layers for Efficient Photovoltaic Devices
Edit

Photovoltaics (PVs) has become one of the emerging alternatives to progressively supply/replace conventional energy sources, considering the potential exploitation of solar energy. Depending on the nature of the light harvester to influence on its light-absorption capability and the facility to produce electricity, different generations of solar devices have been fabricated. Early studies of organic molecules (dye sensitizers) with good absorption coefficients, going through metal chalcogenides and, lastly, the timely emergence of halide perovskites, have promoted the development of novel and low-cost solar cells with promising photoconversion efficiency (PCE), close to the well-established Si-based devices.

light harvesting organic molecules metal chalcogenides metal halide perovskites active layer

1. Introduction

Photovoltaics (PVs) is one of the most important topics in the development of prominent strategies for finding sustainable energies, with sunlight being an unlimited energy source [1]. This factor has been pivotal considering that both the high energy demand around the world and the subsequent replacement of traditional energy sources such as fossil fuels, coke, and natural gas, among others, have been promoted to mitigate the growing environmental contamination and global warming [2][3]. To guarantee a maximum capability of light absorption, the main component of solar devices is based on semiconductor materials, which collect the sunlight to produce free electrons (sunlight-to-current transformation), which are transported through the fabricated cell [4][5]. In order to optimize the absorption of sunlight, various light trapping techniques, including texturing [6], anti-reflective coatings [7][8] and plasmonic effects [9], can be employed. These methodologies are implemented to enhance the capture of light within the active layer of a solar cell. Furthermore, it is relevant to mention that to certify the performance of the fabricated devices, some operational parameters are established, such as the simulated sunlight intensity, which should contain the AM 1.5 G filter, and the direction of the voltage scanning (forward/reverse) during the photovoltaic measurements. For this contribution, the researchers show the maximum photoconversion efficiency developed by the different solar cells, independent of the direction of voltage scanning, considering that the reported performance is established for optimized devices (ideally similar efficiency in forward and reverse measurements).
For decades, silicon (Si)-based solar cells have dominated the photovoltaics market because these devices can reach photoconversion efficiencies (PCEs) of more than 20%, the photomaterial is non-toxic, providing good photoconductivity properties, resistance to corrosion, and high light intensities and temperatures [4][5][10]. However, even though Si-based solar cells can be obtained by using different Si active layers (monocrystalline, polycrystalline, nanostructured and amorphous Si films), the material preparation and processing are high energy demanding processes to melt the Si source, which increases the fabrication cost for Si-based solar cells [11]. Hence, the final costs of the device production and installation increases, which is a huge challenge in solar technologies. With the purpose of offering a step forward in the fabrication of lower-cost solar cells with promising PCE, another kind of semiconductor has been introduced, as in the case of II-VI type CdTe thin films [12][13]. This type of chalcogenide can be directly deposited as a polycrystalline film through low-cost techniques including close vapor transport [14], laser ablation [15], sputtering [16] and electrodeposition [17]. CdTe thin films have also been prepared by using chalcogenide nanoparticles dispersed in organic or aqueous solutions and deposited by dip-coating/spin-coating techniques, followed by an annealing process [18][19][20].
On the other hand, CdTe shows an optimal band gap ~1.42 eV (indicating a suitable light harvesting feature in a wide range of the UV-vis-NIR spectrum), direct band transitions, and a high absorption coefficient ~104 cm–1 [21]. This feature is adequate for absorbing ~90% of incident illumination for 1 µm thickness CdTe film. Additionally, this chalcogenide can be produced with n- or p-type semiconductivity, making it ideal for establishing heterojunctions for carrier transfer. Nevertheless, some disadvantages, such as a high resistance to carrier transport and fast surface recombination, hinder electron mobility through the solar devices, decreasing their operational performance [22]. In this way, reasonable approaches such as the preparation of chalcogenide-based heterostructures, for instance, CdS/CdTe ones, are known to enhance carrier injection and transfer into the solar cell, which increases the net PCE [23][24]. Unfortunately, the recent PCE is far from the established record of solid-state solar cells, some of them based on tandem perovskite-Si devices (whose main properties are highlighted below), and it is deductible that a wide understanding of the required components of the solar cells and the quality of the active layer should be pivotal in order to increase the PV efficiency.
The first step to maximizing the sunlight harvesting in low-cost PV devices was introduced by Grätzel in 1991 [25], to fabricate dye-sensitized solar cells (DSSCs) by using natural/synthetic organic materials. These molecules favored carrier generation and transport, obtaining an initial PCE of 10%. DSSCs are based on the deposition of ruthenium (II) polypyridyl complex monolayers on TiO2 films acting as working electrodes (photoanodes). Meanwhile, platinum-transparent glass is used as a counter electrode (also called a cathode), being coupled with the active layer to achieve the expected device. The operational performance of the DSSCs to produce electricity depends on the efficiency of exciton formation. In this scenario, the photogenerated electrons coming from the highest occupied molecular orbital (HOMO) of the organic complex (hereafter named sensitizer) are accumulated in the lowest unoccupied molecular orbital (LUMO), and then transferred to the conduction band (CB) of the TiO2 layer. Then, these carriers are transported to the external circuit of the PV cell to reach the cathode, generating photocurrent. The electron depletion in the HOMO of the dye is filled by electrons from the electrolyte, which is composed of a redox system based on the I/I3 pair. Lastly, iodide is regenerated by reducing the triiodide species through electrons collected by the cathode, closing the circuit where the current is produced. This contribution opened the door to new organic photomaterials based on dyes, polymers, and small molecules with suitable electronic properties, giving the possibility to fabricate organic solar cells [26][27][28]. Although the selective contacts (electron and hole transporter layers, ETL and HTL, respectively) impact carrier generation and transport into the PV device, organic molecules exhibit limited light harvesting in a wide range of the energy spectrum and restrained exciton diffusion [29]. These features favor a decrease in carrier density to be transported into the solar cells, hampering their operational performance.
Attending to the above device configuration from the DSSCs, organic molecules have been replaced by inorganic hybrid sensitizers, as in the case of Cd-free chalcogenides (in form of quantum dots, QDs) [30][31][32][33] and then halide perovskite (HPs) type active layers (in the form of bulk) [34][35], with fascinating features such as high absorption coefficient, long carrier diffusion length, and modulable optical properties depending on the photomaterial size and composition. These advantages allow for the extension of the the light harvesting even to the IR range of the energy spectrum, improving the capability of the sensitizer to generate the photocarriers required to favor a high PCE [36]. In this context, the band gap and relative band positions of the QDs and bulk perovskites can be modified to match the ETL and HTL and thereby facilitate the carrier injection to the selective contacts to generate the expected photoresponse [37][38]. Chalcogenide QDs have reached a maximum PCE of 14.4% [39], while HP bulk films have achieved PCE close to 26% [40][41], overpassing the actual performance of the Si- and CdTe-based technologies. Interestingly, most of the abovementioned features of both kind of active layers are also useful for another kind of research application, such as the fabrication of photodetectors with high responsivity and specific detectivity [42][43][44][45][46][47][48][49][50].
The combination of HPs-based solar cells (PSCs) and Si-solar devices has promoted the development of tandem solar cells, with PCE values of 33.2% [51]. The transparency and thickness control of the active layers and the band gap tuning have been considered to be promising devices for future commercialization. Despite the light-to-current transformation being eventually improved in current prototypes of solar cells, the use of inorganic materials is found to be restrained due to their ionic nature, which produces photocorrosion in chalcogenides [52] and the formation of non-photoactive crystalline phases in HPs, as in the case of iodide-HPs.
As shown above, reports have widely discussed the fundamentals of the photovoltaic effect in solar cells, which have been adopted to describe the way that the active layers, or sensitizers, capture the sunlight to produce electricity through the formation of photocarriers, which are separated and transported through the devices. This fact has led some researchers to believe that the emergence of the following generation of solar cells offers more advantages than the previous one, but it has not been deeply explained from the point of view of individual components how the fabrication of innovative solar devices can be more efficient and exhibit enhanced long-term stability. More specifically, few reports have consolidated the advantages/disadvantages of using different types of materials as active layers, where the photovoltaic effect starts, and some explanation about how the integration of these sensitizers into a coupled solar system can magnify the net performance of a solar cell to be competitive with the actual commercialized Si-based ones has not been widely highlighted.

2. Progress in the Development of Active Layers for Efficient PV Devices

2.1. Organic Solar Cells

Organic solar cells (OSCs) have recently attracted attention worldwide since they are one of the most promising energy sources due to their distinctive advantages, such as light-weight mechanical flexibility, being colorful, translucent, nontoxic and being solution-processable for huge area fabrication [53][54][55][56][57]. Over the last three decades, the area of OSCs has made impressive progress in efficiency, which probably makes them a competitive alternative to inorganic solar cells in the near future. Here, the researches are going to classify OSCs into three different groups based on their structure: binary, ternary, or tandem OSCs and dye-sensitized solar cells (DSCs).

2.1.1. Binary, Ternary, and Tandem Organic Solar Cells

The first manufacture of organic photovoltaic solar cells was based on single organic films inserted between two metal electrodes of different work functions [58][59]. Initially, organic materials with p- and n-type conductivities are deposited on a transparent electrode, and then the metal electrode is placed over them. The operation of OSCs consists of the fact that sunlight is absorbed by organic layers, and produced excitons are dissociated to electrons and holes at the interface amongst p- and n-type organic layers. Finally, holes and electrons are diffused at the electrodes, generating electricity [58][59][60]. The PCEs reported by these devices were really poor (around 0.01%), although they reached a significant 0.7% with the incorporation of merocyanine dyes [61][62]. The next evolution in this area was performed by introducing the heterojunction concept, in which two organic films with special electron- or hole-transporting properties were sandwiched between the electrodes. Tang and coworkers [63] manufactured the first organic photovoltaic device based on planar heterojunction (PHJ) of two organic materials (phthalocyanine and perylene derivatives), one carrier donor, and one carrier acceptor, which were used to distribute the excitons. These organic layers were sandwiched between a transparent conducting oxide and a semitransparent metal electrode, affording effective charge extraction. Although there were very few donor/acceptor interfaces, this device was quite limited. This solar cell afforded around 1% of PEC, which was the best result obtained for almost 10 years.
On the other hand, the development of conjugated polymers grew enormously in the early 1990s, although their incorporation as single-layer devices was not initially successful since the values of PCEs obtained by this type of material were less than 0.1% [64][65]. Nevertheless, the discovery of photoinduced electronic transference between optically excited conjugated polymers and the fullerene molecule (C60) [66][67], along with the elevated photoconductivities shown by the addition of C60 to the conjugated polymers [68], allowed the fabrication of devices based on polymer-fullerene bilayer heterojunction [69][70] or bulk heterojunction (BHJ) [71][72][73]. This advance significantly increased the values of PCEs of the corresponding devices for almost two decades [74][75][76][77][78][79][80][81]. The photoinduced electron mobility favorably happens in the polymer-fullerene system when the electron in the excited state of the polymer is transferred to the fullerene, which is quite more electronegative since the electron is injected from a p-type hole belonging to the conducting polymer, donor (D), to the n-type electron from the conducting C60 molecule, acceptor (A) [82]. This occurs because fullerene derivatives present a fully conjugated structure, which gives them powerful electron-accepting and isotropic electron-transport abilities and favors electron delocalization at the D:A interfaces [83][84]. Some time ago, it was believed that fullerene materials such as PC61BM ([6,6]-phenyl-C61-butyric acid methyl ester) [85], F1 ([6,6]-phenyl-C61-propionic acid methyl ester) [86], and F [87] were a basic component for efficient operation of OSCs. However, this scenario has recently changed with the appearance of non-fullerene acceptors (NFA) such as ITIC (3,9-bis(2-methylene-(3-(1,1-dicyanomethylene)-indanone)-5,5,11,11-tetrakis(4-hexylphenyl)-ditheno [2,3-d:2′,3′-d′]-s-indaceno [1,2-b:5,6-b′]-dithiophene) [88], ITIC-Th1 (3,9-bis(2-methylene-(5&6-fluoro-(3-(1,1-dicyanomethylene)-indanone)-5,5,11,11-tetrakis(4-hexylthienyl)-dithieno [2,3-d:2′,3′-d′]-s-indaceno [1,2-b:5,6-b′]-dithiophene) [89], FOIC [90], and Y6 ((2,20-((2Z,20Z)-((12,13-bis(2-ethylhexyl)-3,9-diundecyl-12,13-dihydro-[1,2,5]thiadiazolo [3,4-e]thieno [2,”30′:4′,50]thieno [20,30:4,5]pyrolo [3,2-g]thieno [20,30:4,5]thieno [3,2-b]indole-2,10-diyl)bis(methanylylidene))bis(5,6-difluoro-3-oxo-2,3-dihydro-1H-indene-2,1-diylidene))dimalononitrile)) [91], which have increased notably the values of PCEs obtained [83][92]. The performance of the NFA-based OSCs is closely related to the morphology of the functional layers and the material processing, although in some cases the change in the stratification of the donor and acceptor in the BHJ [93][94][95] and the chemical stability at the BHJ/contact interfaces [96] have also played a decisive role in increasing the PCE values in these devices. The fast development of OSCs based on NFA has caused these devices have crossed the PCE barrier of 19% over the last years [97][98][99][100][101][102][103][104]. Nevertheless, the broad range of donor molecules previously used for fullerene OSCs provided a splendid library for direct use in NFA OSCs. From the device perspective, excitons in NFA OSCs can be effectively separated by negligible driving energies, which contributes to increasing their PCEs. Consequently, this type of devices regularly shows elevated photocurrent and low voltage losses simultaneously; nonetheless, fullerene-based OSCs present problems under low driving energies frequently [98].

2.1.2. Dye-Sensitized Solar Cells

Conceptually similar to the BHJ, there is another broad research field based on DSSCs. The early development of these devices was performed by the Grätzel research group [25][105][106][107]. As discussed previously, DSSCs have greatly focused the attention of the scientific community over the last few years due to their promising performance and low fabrication cost [108][109]. It is important to highlight that water-based electrolytes DSSCs are being developed owing to their interesting advantages such as reduced volatility, non-flammability, and environmental compatibility, affording a PCE value over 7% [110]. Additionally, DSSCs are able to be produced on a large scale and present a commercial application for self-powering systems and portable electronics [111][112]. On the other hand, specifically DSSCs prepared with copper-complex-based redox mediator have shown great progress in obtaining PCE values close to 13%, since their Voc can achieve values over 1.0 V [111][112], lower compared to the values observed by high performance tandem OSCs. However, these devices exhibit low Jsc values, considerably reducing their PCE performance. Probably, the principal disadvantage of DSSCs compared with other OSCs is related to the relatively low PCE values reported for these devices; it is difficult to find DSSCs with PCE values over 14%. On the contrary, the most important advantage of DSSCs is their ability to exhibit outstanding stability, maintaining high percentages of their initial PCE after long periods of time.

2.1.3. Comparison of J–V Curves and Normalized PCE vs. Time for Different Types of Organic Solar Cells

Binary, ternary, and tandem OSCs or DSSCs are devices with different characteristics that are determined by their values of Voc, fill factor (FF) and Jsc. These variables are defined by the current density and voltage (J–V curves) properties of the solar cells [98]. These parameters are closely related to the photoactive materials (acceptor and donor materials). Thus, the number of free carriers collected at the electrodes at zero applied potential (Voc = 0) is what determines the value of Jsc. The energy difference amongst the HOMO and LUMO orbitals of the acceptor and donor materials influences the highest voltage that a solar cell could extract for an external circuit, ascribed to the value of Voc. For example, in PM1:BTP-eC9 systems, device efficiency improved from 17.86% to 19.10% with TCB processing. It is important to comment that the TCB-ISM device exhibits the highest PCE value (19.31%) reported to date for a binary OSCs [97]. Additionally, the ternary blend PBQx-TCl:PBDB-TF:eC9-2Cl displays a maximum PCE of 19.51%, a Voc of 0.886 V, a Jsc of 27.15 mA cm−2, and an FF of 81.14% [113]. The tandem OSC based on ICLs of e-TiOx/PEDOT:PSS system achieved a Voc value over 2.0 V, which is considerably higher than the values of Voc showed by binary or ternary OSCs, affording a PCE value of 20.27% [114]. It is important to highlight that DSSCs with ZS4 performed a Jsc of 16.3 mA cm−2, as well as a great Voc value of 1.05 V, accomplishing an impressive PCE of 13.2% under AM1.5G sunlight. This efficiency is the highest one reported for copper-electrolyte-based DSSCs with a single sensitizer [108].

2.2. Chalcogenide Solar Cells

Chalcogenide solar cells are a class of thin-film solar cells that utilize semiconductor materials from the chalcogenide group (e.g., sulfur, selenium, tellurium) to absorb the solar radiation, produce photocarriers, and thereby generate electricity. These solar cells have the advantage of being cost-effective, with good conversion efficiencies. They typically consist of a substrate, transparent conductive oxide layer, p-type and n-type chalcogenide layers, and an intrinsic layer. Here, the researchers will explain some characteristics of binary, ternary, and quaternary chalcogenides usually employed for the fabrication of solar devices.

2.2.1. Binary, Ternary, and Quaternary Chalcogenides

Binary chalcogenide solar cells use semiconductor materials composed of two elements, one from the chalcogenide group, such as sulfur (S), selenium (Se), or tellurium (Te) and a metal. These materials are abundant and relatively low-cost, making them attractive for solar cell applications. The structure of binary chalcogenide solar cells typically consists of a substrate, a transparent conductive oxide layer (TCO), a p-type chalcogenide layer, an intrinsic layer, and an n-type chalcogenide layer. The TCO layer allows light to pass through while providing electrical conductivity. The p-type and n-type chalcogenide layers help in creating an electric field and facilitating the movement of charge carriers. When sunlight is absorbed by the chalcogenide layers, electron-hole pairs are created, which are then collected to generate an electric current [115].

2.2.2. Chalcogenide Tandem Solar Cells

Chalcogenide tandem solar cells have emerged as a promising candidate for achieving high-efficiency photovoltaic devices. In this type of solar cells, two or more different chalcogenide-based semiconductors with diverse bandgaps are assembled to absorb a broader range of the solar spectrum [116]. The upper cell absorbs the higher-energy photons, while the lower cell absorbs the lower-energy photons that pass through the upper cell [117]. This allows for a more efficient utilization of the solar spectrum and can lead to a higher overall efficiency. Chalcogenide-based materials, such as copper indium gallium selenide (CIGS), cadmium telluride (CdTe) [118], lead sulfide (PbS) [119][120] and earth-alternative materials such as kesterite-based chalcogenides [121][122] have been extensively studied for use in tandem solar cells. CIGS is particularly attractive due to its high absorption coefficient and tunable bandgap, making it suitable for use in both the top and bottom cells of the tandem device. Besides, CIGS tandem solar cells have high carrier mobility, scalability, potential for low-cost, and compatibility with alternative materials.

2.2.3. New Chalcogenide Materials

II-IV-N2 materials such as MgSnN2 (MTN) and ZnSnN2 (ZTN) are being investigated for their potential as absorber layers in solar cells [123][124]. Both materials are nitrides that belong to the wurtzite crystal structure and are composed of earth-abundant elements [125]. MTN and ZTN have bandgap energies in the visible range, making them well-suited for solar absorption [125]. They also have high carrier mobility, which is important for efficient charge transport in solar cells [125]. However, there are still challenges to the manufacture of these materials. For instance, to create high-quality thin films of MTN and ZTN high-temperature growth techniques such as molecular beam epitaxy or metalorganic chemical vapor deposition are used, which can be expensive and difficult to scale up for large-scale production. Moreover, the properties of the films can be sensitive to the growth conditions, leading to variations in their electronic and optical properties. Another challenge is the development of suitable heterojunctions and contacts for efficient charge collection in solar cells based on MTN and ZTN. Several studies have explored the use of different materials, such as graphene or molybdenum disulfide, as contacts for MTN and ZTN solar cells, but further research is needed to optimize their performance. Efficiencies of ZTN solar cells reach 1.5% [126], which is still low compared to conventional solar cells.

2.3. Halide Perovskite Based Solar Cells

2.3.1. Pb-Based Perovskites Active Layers

Since Miyasaka and coworkers integrated HPs as promising light harvesters to provide a step forward in a new generation of solar cells in 2009 [127], these materials have been consolidated as the most prominent semiconductors for photovoltaics. Typically, HPs show a 3D structure based on ABX3 formula (A = Cs+, cesium; MA+, methylammonium; FA+, formamidinium; B = Pb2+, Sn2+; X = Cl, Br, I, or combinations), where the composition and particle size engineering can promote the preparation of HPs with diverse photophysical, chemical, and electronic properties [128][129][130]. Due to the fact that a solar device should collect as much incident illumination as possible in a broad range of the energy spectrum to reach an efficient light-to-current transformation, low-band gap HPs are ideal for this process. Therefore, HPs with high iodide content are in the focus of many research groups, obtaining different strategies to reach a suitable active layer, compatible and stable with the selective contacts of the solar cell, to facilitate the carrier mobility through the device [131][132]. Although an acceleration in the number of publications have been reported to show current alternatives to prepare adequate HPs active layers (open the door to an extensive number of reviews), here, the researchers will highlight some of these contributions where the modification of the perovskite layer has been vital to increasing the performance of the solar devices.
Similar to the generations of solar devices shown above, PSCs are composed by an electron transporter layer (ETL) and hole transporter layer (HTL), which favor the carrier transport through the device [133]. Moreover, a transparent conducting glass as fluorine-doped tin oxide (FTO) is used to promote the carrier mobility and allow the direct illumination of the active layer, while on top of the PSC architecture, metals such as silver or gold are deposited [134]. As the researchers previously mentioned, iodide HPs are the most relevant photomaterials to achieve a high photoconversion efficiency (PCE) in PSCs, especially perovskites based on FAPbI3. This material presents a band gap around 1.4 eV, indicating a broad absorption capability of solar energy in the UV-vis-NIR spectrum [135][136][137]. This band gap value also corresponds to a theoretical PCE ~32.3%, near to the Shockley–Queisser limit of efficiency [138]. Unfortunately, this feature is achieved for the photoactive phase of the FAPbI3, denominated as black α-phase, which is stable at high temperatures, but is eventually transformed into the yellow δ-phase at room conditions [139]. To suppress the fast α-to-δ phase conversion into this kind of perovskite, Meng and coworkers [140] have prepared bulk films based on FA0.94MA0.06Pb(I0.94,Br0.06)3, which have been coupled with carbon electrodes on top, in a n-i-p device configuration, to accelerate the carrier separation and transport into the fabricated solar cell. Here, bromide anions were incorporated into the perovskite to provide better stability to the material since iodide species are more labile, being rapidly diffused from the perovskite lattice to create structural defects. However, the presence of Br also offers the enlargement of the band gap, hampering the sunlight absorption capability of the perovskite. In this context, a high FA content ~94% was introduced into the material, to guarantee a narrow band gap. Attending to the final device, the PCE was estimated to be 20.04% at 1 sun illumination. At this point, the 94% of the initial performance of the solar cell was maintained for 1000 h.
On the other hand, Kim and coworkers [141] have deeply investigated the impact of the addition of MACl as an additive for the stabilization of FAPbI3 to obtain highly-performance solar cells. Here, they demonstrated that the α-to-δ phase transition is promoted by the weak interaction between FA+ cations and the iodide species from the [PbI6]4− octahedra building blocks, which is one of the key factors to trigger the octahedra tilting [142][143][144]. Thus, the presence of Cl provides a better p orbital localization from I, improving the FA-I interactions beyond compensating halide vacancies into the perovskite structure. Then, MA+ cations also facilitate the perovskite stabilization, causing a contraction of the cubo-octahedra sites in the material. MA+ shows a higher dipolar moment than that of FA cations (~10 times), strengthening the volume shrinkage, making closer a direct interaction between FA and I.
To date, the octahedral tilting is induced by the loss of halide species such as iodide anions during the perovskite film preparation [139][145][146], generating a high density of halide vacancies (thereby, high density of carrier traps is produced), leading to the degradation of the photophysical and electronic features of the material. More specifically, the I anions released from the perovskite structure are easily oxidized to obtain I0, which is the main species to trigger chemical chain reactions and accelerate the quenching of the intrinsic properties of the active layer [146]. For this purpose, different organic agents have been introduced to avoid the formation of halide vacancies and strengthen the Pb-I bonds in the octahedra building blocks, as in the case of caffeine [147], theobromine [148], alkylammonium halides [148], phosphonopropionic acid [149] and some amino-based organic ligands [150]. Although some of these organic agents show a high binding capability to the [PbI6]4− octahedra units, avoiding the halide deficiency, they also exhibit fast decomposition under heating or can acts as barrier layers (due to their long carbon chains). This fact hinders the carrier extraction from the perovskite, hampering the net performance of the solar device. Therefore, with the purpose of reducing the iodide migration and thereby inhibit the formation of halide defect sites in the perovskite layer, Zhao and coworkers [151] have added 3-amidinopyridine (3AP) during the material preparation, inducing the coordination between amidino moieties and the Pb-I frameworks. In this scenario, a maximum PCE of 25.3% was obtained in optimized PSCs under 1 sun irradiation, preserving ~92% of the initial efficiency for 5000 h in ambient air.
Other alternatives to stabilize the perovskite active layer have been focused on the direct incorporation of large alkylammonium species to react with the 3D HPs to form multidimensional 2D/3D systems with longer durability [150][152][153]. Considering that big cations do not match with the Goldschmidt tolerance factor, breaking the 3D preferential crystalline structure of the perovskite, the [PbX6]4− octahedra in form of stacking layers are separated by these of organic agents [154][155]. The 2D perovskites are environmentally stable and promote Vander Waals interactions with the Pb-X frameworks. However, these materials exhibit a large band gap and poor carrier transport capability, decreasing the PCE in modified PSCs. In this way, it is favorable the mixture of 2D/3D systems to facilitate the separation/mobility of electrons through the device. Grancini and coworkers [156] have proposed the incorporation of 2-thiophenemethylammonium halides, 2-TMAX (X = Cl, Br and I) during the preparation of triple-cation [(FAPbI3)0.87(MAPbBr3)0.13]0.92(CsPbI3)0.08 perovskite, where the thiophene organic species react with the photomaterial to create the 2D layer. By adding 2-TMABr and 2-TMAI agents, a suitable band structure (valence band shift) in between the 2D-3D perovskites is mediated, favoring the hole extraction into the heterojunction, contrary to 2-TMACl, where the carrier diffusion was hampered. Accordingly, PSCs with a maximum PCE of 20.8% (under 1 sun) were fabricated, keeping 74% of their initial performance for 1000 h of operation without any encapsulation. Another example is mentioned by Zhou and coworkers [155], where the 2-thiophenemethylammonium (ThMA) spacer was also added during the preparation of FA/MA based 3D perovskite. Here perovskite crystals were orientally grown, inducing the passivation of a high density of carrier traps to avoid the non-radiative recombination losses. Here, devices provided high PCE up to 21.5%; preserving 99% of its initial performance after 1680 h ambient air.
Then, a mixture of organic spacers such as iso-butylammonium iodide (i-BAI) and FA-iodide (FAI) reported by Cho and coworkers [157] suppressed the interfacial carrier traps and thereby reduced the interfacial energy barrier to avoid carrier diffusion. A PCE as high as 21.7% was obtained for the optimized solar device, keeping 87% of the original efficiency after 912 h with a relative humidity of 75%.
To maximize the absorption ability and electronic features of the bulk HPs, these materials have also been integrated with other types of solar devices such as Si, copper indium gallium selenide, different PSCs, DSSCs and QDs-based solar cells to fabricate tandem devices, as shown above [158]. Some works have reached near to 30%, as the case of the contribution reported by Albrecht and coworkers [159], where a monolithic perovskite tandem solar cell (PTSC) was fabricated. Here, a wide-band gap Cs0.05(FA0.77MA0.23)0.95Pb(I0.77Br0.23)3 bulk film was integrated with a Si-based solar device, also incorporating diverse hole transporter layers such as C60, 2PACz and Me-4PACz, to accelerate the carrier extraction. A maximum of PCE ~29% was achieved for the optimized PTSC, retaining the 95% of the initial performance after 300 h under operation. Interestingly, De wolf and coworkers [160] have incorporated MgF2 as an interlayer between the perovskite based on triple cation Cs0.05FA0.8MA0.15Pb(I0.755Br0.255)3 and C60 to improve the electron extraction (better adjustment of the relative energy positions in the perovskite) and separate the C60 from the active layer surface to restrain the non-radiative recombination dynamics. Accordingly, a stabilized PCE of 29.3% was achieved, where the optimized PTSC device retained ~95% of its initial performance after 1000 h of operation, even carrying out damp-heat testing at 85 °C with 85% relative humidity. Recently, this research group provided a new record of the PCE ~33.2% using a similar perovskite layer on top of the cell, while a textured Si solar device is in the bottom of the configuration.
Even though the surface passivation with interesting alkylammonium ligands and the incorporation of 2D/3D heterostructures enhance the operational performance of the solar devices, it is reasonable that some differences in the PCE values are obtained. So, the highest PCE value is reached by filling/replacing structural defects in the active layer, improving its optical and electronic properties, suppressing carrier traps, and facilitating carrier transport. Although 2D/3D assembly can also achieve defect passivation in the perovskite, the appearance of a heterostructure can promote the emergence of interfacial energy levels where the carriers can be trapped, hampering the development of a higher device efficiency.

2.3.2. Partially Pb-Substituted/Pb-Free Perovskite Active Layers

As is highlighted above, a series of approaches are prominent to reach long-term stability and suitable electronic properties in perovskites to favor carrier separation/extraction and mobility through the fabricated device. However, it is also known that one of the main drawbacks of the PSCs is the inherent toxicity of Pb, which hampers the future commercialization of this technology [161][162]. Therefore, the fact of substituting Pb partially or completely with less toxic metals, preserve/increase the final properties of the active layer, and thereby, reaching high performance in the solar device, is still a huge challenge. One of the most attractive metals to replace Pb is Sn, which also generate [SnX6]4− octahedra units to generate multidimensional HPs structures as the case of Pb and provide HPs lower band gap than that of Pb-analogous [163]. Nevertheless, Sn-HPs exhibit a poor stability consequence of the fast Sn2+-to-Sn4+ oxidation under ambient conditions, degrading the net efficiency in solar devices [164][165]. Hence, the researchers will focus on the fabrication of PSCs based on a mixture of Sn/Pb- or Sn-HPs introducing novel strategies to avoid Sn oxidation in the active layer and thereby increase the performance of Sn-PSCs.
Firstly, for Sn/Pb-HPs systems, Wang and coworkers [166] have studied the addition of antioxidants such as tea polyphenol (TP) during the preparation of optimized CsPb0.5Sn0.5I2Br bulk film to avoid the rapid Sn oxidation. Through the coordination bonds between high Lewis acidity Sn sites and hydroxybenzene groups, the Sn4+/Sn2+ ratio into the HP was decreased from 68.1% to 5.85%, deducing that TP is a good reducing agent. In addition, this antioxidant mediated an improved HP crystallization, making that at 0.5 wt% TP, a homogeneous film with a low content of pinholes is obtained. Therefore, low density of grain boundaries appears to hinder the carrier extraction. A maximum PCE of 8.1% was reached under 1 sun illumination, retaining 95% of original performance after ~1440 h into a N2-filled glove box. Meanwhile, Guo and coworkers [167] have used the mixture of 2D/3D HPs in order to prepare suitable heterostructures for PSCs. Here, authors have introduced bulky alkylammonium based on 2-(4-fluorophenyl)ethylammonium iodide (FPEAI) during the preparation of (MAPbI3)0.75(FASnI3)0.25 films. In presence of the organic cation, highly oriented crystals are grown, and phase segregation occurring as the result of the Sn oxidation (emerging two distinguished signals from MAPbI3 and FASnI3 HPs), is suppressed. This fact makes that the recombination lifetime is longer than that of the absence of the organic molecules, associated with the reduction of high density of trap states. In context, a high PCE of 17.51% was reached, keeping 90% of the initial performance after 1200 h into a glove box, and 70% of initial efficiency close to 400 h under ambient air.
Then, Xu and coworkers [168] employed the surface passivation concept to prolong the stability of the FA0.7MA0.3Pb0.7Sn0.3I3 HPs and remove structural defects from the material using guanidinium bromide (GABr). Considering that GA+ cations provide more charge density than FA+ and MA+ cations, this could be one of the main explanations to restrain the fast Sn2+-to-Sn4+ transformation. In addition, iodide vacancies are formed by the depletion of Sn2+, creating high content of halide vacancies into the perovskite. Thus, Br domains fill/replace these empty states to enhance the final properties of the prepared active layer. The optimized device has reached a maximum PCE of 20.63% without any encapsulation, conserving over 85% of the original efficiency after 1000 h under ambient air, and 80% of the original efficiency after 24 h of operation at 80 °C. Lastly, Huang and coworkers [160] have demonstrated that the addition of an alkylammonium pseudo-halide additive based on octylammonium tetrafluoroborate (OA+BF4) for the preparation of MA-free Sn/Pb HPs. While OA+ cations can fill some A-site cations into the HPs, BF4 species compensate the iodide vacancies generated by the Sn oxidation, also reducing the likelihood of emerging interstitial iodide and iodine. The suppression of defect density, main factor to promote carrier trapping, facilitates the fabrication of PSCs with PCE up to 23.7% (a record for Sn/Pb-PSCs), keeping over 88% of the initial performance after 1000 h of continuous operation at 50 °C in air and tracking the device under MPP.
For the case of the stabilization of Sn-based PSCs, some examples can be addressed. Mora-Seró and coworkers [169] have made use of the chemical engineering strategy, where Dipropylammonium iodide (DipI) and sodium borohydride (NaBH4) (known as a good reducing agent) have been added to the preparation of FASnI3 films and prevent the Sn oxidation. While DipI can block the loss of iodine in form of I2, NaBH4 reduce the released I2 to produce iodide into the HPs, avoiding the halide migration. Therefore, the Sn2+/Sn4+ and I/I2 ratios are increased, inhibiting the formation of interstitial iodine/iodide vacancies, which are the main energy states where the hole/electrons can be trapped. In this alternative for stabilizing Sn, optimized devices boost a high PCE of 10.61%, providing up to 96% of the initial performance after 1300 h of continuous operation at MPP in N2 condition. Another approach to stabilizing the black phase of the FASnI3, Mi and coworkers [170] have introduced trimethylthiourea (3T) during the spin-coating process for HPs preparation. The presence of this bifunctional ligand induces coordination bonds with Sn2+ cations through the formation of Sn-S species. Simultaneously, N-H moiety from the organic agent produces hydrogen bond iodide species from the octahedra building blocks to avoid their diffusion out from the HPs lattice. Due to these effects, the carrier recombination is prolonged in comparison to the pristine HPs, hindering the emergence of carrier traps. In this context, PSCs were able to develop a maximum PCE up to 14%, maintaining 100% of the original efficiency for more than 700 h, under N2 atmosphere. On the other hand, He and coworkers [171] have included 4-fluoro-phenethylammonium bromide (FPEABr) into the precursor solution to prepare 2D/3D Sn-HPs with low density of structural defects. FPEA cations are incorporated into the perovskite surface, while F- and Br-anions fill/replace halide vacancies. In terms of the device architecture, the 2D Sn-perovskite acts as an interface layer between the active layer and PEDOT:PSS hole transporter layer, restraining the iodide migration into the device. In presence of 10 wt% FPEABr, encapsulated 2D/3D PSCs generate a PCE up to 14.81%, exhibiting over 80% initial performance after 432 h under operation at room temperature, and 60% of initial efficiency after 1 h at 80 °C, both kinds of tests under ambient air.

References

  1. Fthenakis, V.M. End-of-Life Management and Recycling of PV Modules. Energy Policy 2000, 28, 1051–1058.
  2. Carneiro, A.L.; Martins, A.A.; Duarte, V.C.M.; Mata, T.M.; Andrade, L. Energy Consumption and Carbon Footprint of Perovskite Solar Cells. Energy Rep. 2022, 8, 475–481.
  3. Dissanayake, P.D.; Yeom, K.M.; Sarkar, B.; Alessi, D.S.; Hou, D.; Rinklebe, J.; Noh, J.H.; Ok, Y.S. Environmental Impact of Metal Halide Perovskite Solar Cells and Potential Mitigation Strategies: A Critical Review. Environ. Res. 2023, 219, 115066.
  4. Mathan Kumar, P.; Das, A.; Seban, L.; Nair, R.G. Fabrication and Life Time of Perovskite Solar Cells. In Perovskite Photovoltaics Basic to Advanced Concepts and Implementation; Academic Press: Cambridge, MA, USA, 2018; pp. 231–287.
  5. Sundaram, S.; Benson, D.; Mallick, T.K. Overview of the PV Industry and Different Technologies. Sol. Photovolt. Technol. Prod. 2016, 7–22.
  6. Campbell, P.; Green, M.A. High Performance Light Trapping Textures for Monocrystalline Silicon Solar Cells. Sol. Energy Mater. Sol. Cells 2001, 65, 369–375.
  7. Khezami, L.; Al Megbel, A.O.; Jemai, A.B.; Ben Rabha, M. Theoretical and Experimental Analysis on Effect of Porous Silicon Surface Treatment in Multicrystalline Silicon Solar Cells. Appl. Surf. Sci. 2015, 353, 106–111.
  8. Karyaoui, M.; Bardaoui, A.; Ben Rabha, M.; Harmand, J.C.; Amlouk, M. Effect of Rapid Oxidation on Optical and Electrical Properties of Silicon Nanowires Obtained by Chemical Etching. EPJ Appl. Phys. 2012, 58, 20103.
  9. Tan, H.; Santbergen, R.; Smets, A.H.M.; Zeman, M. Plasmonic Light Trapping in Thin-Film Silicon Solar Cells with Improved Self-Assembled Silver Nanoparticles. Nano Lett. 2012, 12, 4070–4076.
  10. Wang, Z.; He, J.; Wang, W.; Lin, H.; Xu, Z.; Liu, Q.; Peng, S.; Hou, J.; He, D.; Gao, P. Twenty Percent Efficiency Crystalline Silicon Solar Cells with Solution-Processed Electron-Selective Contacts. ACS Appl. Energy Mater. 2021, 4, 3644–3650.
  11. Barkhouse, D.A.R.; Gunawan, O.; Gokmen, T.; Todorov, T.K.; Mitzi, D.B. Yield Predictions for Photovoltaic Power Plants:Empirical Validation,Recent Advances and Remaining Uncertainties. Prog. Photovolt. Res. Appl. 2015, 20, 6–11.
  12. Chavan, K.T.; Chandra, S.; Kshirsagar, A. Half-Metallicity in Smallest Cage-like Cluster of CdTe with Doping of Transition Metal Atoms. Mater. Today Commun. 2022, 30, 103104.
  13. Chu, T.L.; Chu, S.S. Thin Film II–VI Photovoltaics. Solid. State. Electron. 1995, 38, 533–549.
  14. Kumarasinghe, P.K.K.; Dissanayake, A.; Pemasiri, B.M.K.; Dassanayake, B.S. Thermally Evaporated CdTe Thin Films for Solar Cell Applications: Optimization of Physical Properties. Mater. Res. Bull. 2017, 96, 188–195.
  15. Rivera, L.P.; García, E.; Cardona, D.; Pérez-Centeno, A.; Camps, E.; Santana-Aranda, M.A.; Gómez-Rosas, G.; de Moure-Flores, F.; Chávez-Chávez, A.; Quiñones-Galván, J.G. CdTe:Sn Thin Films Deposited by the Simultaneous Laser Ablation of CdTe and Sn Targets. Mater. Res. Express 2020, 7, 015905.
  16. Camacho-Espinosa, E.; López-Sánchez, A.; Rimmaudo, I.; Mis-Fernández, R.; Peña, J.L. All-Sputtered CdTe Solar Cell Activated with a Novel Method. Sol. Energy 2019, 193, 31–36.
  17. Ling, J.; Zhang, X.; Mao, T.; Li, L.; Wang, S.; Cao, M.; Zhang, J.; Shi, H.; Huang, J.; Shen, Y.; et al. Electrodeposition of CdTe Thin Films for Solar Energy Water Splitting. Materials 2020, 13, 1536.
  18. Vaishanav, S.K.; Korram, J.; Pradhan, P.; Chandraker, K.; Nagwanshi, R.; Ghosh, K.K.; Satnami, M.L. Green Luminescent CdTe Quantum Dot Based Fluorescence Nano-Sensor for Sensitive Detection of Arsenic (III). J. Fluoresc. 2017, 27, 781–789.
  19. Kini, S.; Kulkarni, S.D.; Ganiga, V.; Nagarakshit, T.K.; Chidangil, S. Dual Functionalized, Stable and Water Dispersible CdTe Quantum Dots: Facile, One-Pot Aqueous Synthesis, Optical Tuning and Energy Transfer Applications. Mater. Res. Bull. 2019, 110, 57–66.
  20. Duan, J.; Song, L.; Zhan, J. One-Pot Synthesis of Highly Luminescent CdTe Quantum Dots by Microwave Irradiation Reduction and Their Hg2+-Sensitive Properties. Nano Res. 2009, 2, 61–68.
  21. Rangel-Cárdenas, J.; Sobral, H. Optical Absorption Enhancement in CdTe Thin Films by Microstructuration of the Silicon Substrate. Materials 2017, 10, 607.
  22. Kavanagh, S.R.; Walsh, A.; Scanlon, D.O. Rapid Recombination by Cadmium Vacancies in CdTe. ACS Energy Lett. 2021, 6, 1392–1398.
  23. Mendoza-Pérez, R.; Aguilar-Hernández, J.; Sastre-Hernández, J.; Ximello-Quiebras, N.; Contreras-Puente, G.; Santana-Rodríguez, G.; Vigil-Galán, O.; Moreno-García, E.; Morales-Acevedo, A. Photoluminescence Characteristics of CdS Layers Deposited in a Chemical Bath and Their Correlation to CdS/CdTe Solar Cell Performance. Sol. Energy 2006, 80, 682–686.
  24. Morales-Acevedo, A. Thin Film CdS/CdTe Solar Cells: Research Perspectives. Sol. Energy 2006, 80, 675–681.
  25. O’Regan, B.; Grätzel, M. A Low-Cost, High-Efficiency Solar Cell Based on Dye-Sensitized Colloidal TiO2 Films. Nature 1991, 353, 737–740.
  26. Azaid, A.; Raftani, M.; Alaqarbeh, M.; Kacimi, R.; Abram, T.; Khaddam, Y.; Nebbach, D.; Sbai, A.; Lakhlifi, T.; Bouachrine, M. New Organic Dye-Sensitized Solar Cells Based on the D-A-π-A Structure for Efficient DSSCs: DFT/TD-DFT Investigations. RSC Adv. 2022, 12, 30626–30638.
  27. Lee, C.P.; Li, C.T.; Ho, K.C. Use of Organic Materials in Dye-Sensitized Solar Cells. Mater. Today 2017, 20, 267–283.
  28. Kwaku Asiam, F.; Mahbubur Rahman, M.; Kumar Kaliamurthy, A.; Muthu, S.; Yadagiri, B.; Cheol Kang, H.; Chen, C.; Yoo, K.; Lee, J.-J. Role of Pi-Electron Density at the Interface of Small Molecule-Sensitized Solar Cells. J. Phys. Chem. C 2023, 127, 3928–3939.
  29. Soonmin, H.; Hardani; Nandi, P.; Mwankemwa, B.S.; Malevu, T.D.; Malik, M.I. Overview on Different Types of Solar Cells: An Update. Appl. Sci. 2023, 13, 2051.
  30. Crisp, R.W.; Kirkwood, N.; Grimaldi, G.; Kinge, S.; Siebbeles, L.D.A.; Houtepen, A.J. Highly Photoconductive InP Quantum Dots Films and Solar Cells. ACS Appl. Energy Mater. 2018, 1, 6569–6576.
  31. Wu, J.; Li, M.-H.; Jiang, Y.; Xu, Q.; Xian, L.; Guo, H.; Wan, J.; Wen, R.; Fang, Y.; Xie, D.; et al. Carrier Management via Integrating InP Quantum Dots into Electron Transport Layer for Efficient Perovskite Solar Cells. ACS Nano 2022, 16, 15063–15071.
  32. Pidluzhna, A.; Stakhira, P.; Baryshnikov, G.; Zavaraki, A.J.; Ågren, H. InP/ZnS Quantum Dots Synthesis and Photovoltaic Application. Appl. Nanosci. 2022, 13, 4969–4975.
  33. Halder, G.; Ghosh, D.; Ali, M.Y.; Sahasrabudhe, A.; Bhattacharyya, S. Interface Engineering in Quantum-Dot-Sensitized Solar Cells. Langmuir 2018, 34, 10197–10216.
  34. Young Kim, J.; Lee, J.-W.; Suk Jung, H.; Shin, H.; Park, N.-G. High-Efficiency Perovskite Solar Cells. Chem. Rev. 2020, 120, 7867–7918.
  35. Mehrabian, M.; Dalir, S.; Mahmoudi, G.; Miroslaw, B.; Babashkina, M.G.; Dektereva, A.V.; Safin, D.A. A Highly Stable All-Inorganic CsPbBr3 Perovskite Solar Cell. Eur. J. Inorg. Chem. 2019, 2019, 3699–3703.
  36. Valluvar Oli, A.; Li, Z.; Chen, Y.; Ivaturi, A. Near-Ultraviolet Indoor Black Light-Harvesting Perovskite Solar Cells. ACS Appl. Energy Mater. 2022, 5, 14669–14679.
  37. Al-Ahmed, A.; Afzaal, M.; Mahar, N.; Khan, F.; Pandey, S.; Zahir, M.H.; Al-Suliman, F.A. The Synergy of Lead Chalcogenide Nanocrystals in Polymeric Bulk Heterojunction Solar Cells. ACS Omega 2022, 7, 45981–45990.
  38. Tao, S.; Schmidt, I.; Brocks, G.; Jiang, J.; Tranca, I.; Meerholz, K.; Olthof, S. Absolute Energy Level Positions in Tin- and Lead-Based Halide Perovskites. Nat. Commun. 2019, 10, 2560.
  39. Song, H.; Lin, Y.; Zhou, M.; Rao, H.; Pan, Z.; Zhong, X. Zn-Cu-In-S-Se Quinary “Green” Alloyed Quantum-Dot-Sensitized Solar Cells with a Certified Efficiency of 14.4%. Angew. Chemie-Int. Ed. 2021, 60, 6137–6144.
  40. Min, H.; Lee, D.Y.; Kim, J.; Kim, G.; Lee, K.S.; Kim, J.; Paik, M.J.; Kim, Y.K.; Kim, K.S.; Kim, M.G.; et al. Perovskite Solar Cells with Atomically Coherent Interlayers on SnO2 Electrodes. Nature 2021, 598, 444–450.
  41. Park, J.; Kim, J.; Yun, H.-S.; Paik, M.J.; Noh, E.; Mun, H.J.; Kim, M.G.; Shin, T.J.; Seok, S. Il Controlled Growth of Perovskite Layers with Volatile Alkylammonium Chlorides. Nature 2023, 616, 724–730.
  42. Sulaman, M.; Yang, S.; Imran, A.; Zhang, Z.; Bukhtiar, A.; Ge, Z.; Song, Y.; Sun, F.; Jiang, Y.; Tang, L.; et al. Two Bulk-Heterojunctions Made of Blended Hybrid Nanocomposites for High-Performance Broadband, Self-Driven Photodetectors. ACS Appl. Mater. Interfaces 2023, 15, 25671–25683.
  43. Sulaman, M.; Yang, S.; Song, Y.; Bukhtiar, A.; Hu, J.; Zhang, Z.; Jiang, Y.; Cui, Y.; Tang, L.; Zou, B. Hybrid Nanocomposites of All-Inorganic Halide Perovskites with Polymers for High-Performance Field-Effect-Transistor-Based Photodetectors: An Experimental and Simulation Study. Adv. Mater. Interfaces 2022, 9, 2200017.
  44. Sulaman, M.; Yang, S.Y.; Zhang, Z.H.; Imran, A.; Bukhtiar, A.; Ge, Z.H.; Tang, Y.; Jiang, Y.R.; Tang, L.B.; Zou, B.S. Lead-Free Tin-Based Perovskites Nanocrystals for High-Performance Self-Driven Bulk-Heterojunction Photodetectors. Mater. Today Phys. 2022, 27, 100829.
  45. Sulaman, M.; Yang, S.; Bukhtiar, A.; Tang, P.; Zhang, Z.; Song, Y.; Imran, A.; Jiang, Y.; Cui, Y.; Tang, L.; et al. Hybrid Bulk-Heterojunction of Colloidal Quantum Dots and Mixed-Halide Perovskite Nanocrystals for High-Performance Self-Powered Broadband Photodetectors. Adv. Funct. Mater. 2022, 32, 2201527.
  46. Dastgeer, G.; Khan, M.F.; Nazir, G.; Afzal, A.M.; Aftab, S.; Naqvi, B.A.; Cha, J.; Min, K.A.; Jamil, Y.; Jung, J.; et al. Temperature-Dependent and Gate-Tunable Rectification in a Black Phosphorus/WS2 van Der Waals Heterojunction Diode. ACS Appl. Mater. Interfaces 2018, 10, 13150–13157.
  47. Perveen, A.; Hussain, S.; Xu, Y.; Raza, A.; Saeed, F.; Din, N.; Subramanian, A.; Khan, Q.; Lei, W. Solution Processed and Highly Efficient UV-Photodetector Based on CsPbBr3 Perovskite-Polymer Composite Film. J. Photochem. Photobiol. A Chem. 2022, 426, 113764.
  48. Sulaman, M.; Sulaman, M.; Song, Y.; Yang, S.; Yang, S.; Saleem, M.I.; Li, M.; Perumal Veeramalai, C.; Zhi, R.; Jiang, Y.; et al. Interlayer of PMMA Doped with Au Nanoparticles for High-Performance Tandem Photodetectors: A Solution to Suppress Dark Current and Maintain High Photocurrent. ACS Appl. Mater. Interfaces 2020, 12, 26153–26160.
  49. Saleem, M.I.; Yang, S.; Zhi, R.; Sulaman, M.; Chandrasekar, P.V.; Jiang, Y.; Tang, Y.; Batool, A.; Zou, B. Surface Engineering of All-Inorganic Perovskite Quantum Dots with Quasi Core–Shell Technique for High-Performance Photodetectors. Adv. Mater. Interfaces 2020, 7, 2000360.
  50. Saleem, M.I.; Sulaman, M.; Batool, A.; Bukhtiar, A.; Khalid, S. Suppression of Mid-Gap Trap State in CsPbBr3 Nanocrystals with Br-Passivation for Self-Powered Photodetector. Energy Technol. 2023, 11, 2300013.
  51. Pv-Magazine KAUST Claims 33.2% Efficiency for Perovskite/Silicon Tandem Solar Cell. Available online: https://www.pv-magazine.com/2023/04/13/kaust-claims-33-2-efficiency-for-perovskite-silicon-tandem-solar-cell/ (accessed on 15 June 2023).
  52. Lee, W.; Son, H.J.; Lee, D.K.; Kim, B.; Kim, H.; Kim, K.; Ko, M.J. Suppression of Photocorrosion in CdS/CdSe Quantum Dot-Sensitized Solar Cells: Formation of a Thin Polymer Layer on the Photoelectrode Surface. Synth. Met. 2013, 165, 60–63.
  53. Zhan, L.; Yin, S.; Li, Y.; Li, S.; Chen, T.; Sun, R.; Min, J.; Zhou, G.; Zhu, H.; Chen, Y.; et al. Multiphase Morphology with Enhanced Carrier Lifetime via Quaternary Strategy Enables High-Efficiency, Thick-Film, and Large-Area Organic Photovoltaics. Adv. Mater. 2022, 34, 2206269.
  54. Zheng, X.; Zuo, L.; Zhao, F.; Li, Y.; Chen, T.; Shan, S.; Yan, K.; Pan, Y.; Xu, B.; Li, C.-Z.; et al. High-Efficiency ITO-Free Organic Photovoltaics with Superior Flexibility and Upscalability. Adv. Mater. 2022, 34, 2200044.
  55. Zuo, L.; Jo, S.B.; Li, Y.; Meng, Y.; Stoddard, R.J.; Liu, Y.; Lin, F.; Shi, X.; Liu, F.; Hillhouse, H.W. Dilution Effect for Highly Efficient Multiple-Component Organic Solar Cells. Nat. Nanotechnol. 2022, 17, 53–60.
  56. Kim, T.; Kim, J.-H.; Kang, T.E.; Lee, C.; Kang, H.; Shin, M.; Wang, C.; Ma, B.; Jeong, U.; Kim, T.-S.; et al. Flexible, Highly Efficient All-Polymer Solar Cells. Nat. Commun. 2015, 6, 8547.
  57. Inganäs, O. Organic Photovoltaics over Three Decades. Adv. Mater. 2018, 30, 1800388.
  58. Wöhrle, D.; Meissner, D. Organic Solar Cells. Adv. Mater. 1991, 3, 129–138.
  59. Chamberlain, G.A. Organic Solar Cells: A Review. Sol. Cells 1983, 8, 47–83.
  60. Yeh, N.; Yeh, P. Organic Solar Cells: Their Developments and Potentials. Renew. Sustain. Energy Rev. 2013, 21, 421–431.
  61. Morel, D.L.; Ghosh, A.K.; Feng, T.; Stogryn, E.L.; Purwin, P.E.; Shaw, R.F.; Fishman, A.C. High-efficiency Organic Solar Cells. Appl. Phys. Lett. 1978, 32, 495–497.
  62. Ghosh, A.K.; Feng, T. Merocyanine Organic Solar Cells. J. Appl. Phys. 1978, 49, 5982–5989.
  63. Tang, C.W. Two-layer Organic Photovoltaic Cell. Appl. Phys. Lett. 1986, 48, 183–185.
  64. Antoniadis, H.; Hsieh, B.R.; Abkowitz, M.A.; Jenekhe, S.A.; Stolka, M. Photovoltaic and Photoconductive Properties of Aluminum/Poly (p-Phenylene Vinylene) Interfaces. Synth. Met. 1994, 62, 265–271.
  65. Marks, R.N.; Halls, J.J.M.; Bradley, D.D.C.; Friend, R.H.; Holmes, A.B. The Photovoltaic Response in Poly (p-Phenylene Vinylene) Thin-Film Devices. J. Phys. Condens. Matter 1994, 6, 1379.
  66. Smilowitz, L.; Sariciftci, N.S.; Wu, R.; Gettinger, C.; Heeger, A.J.; Wudl, F. Photoexcitation Spectroscopy of Conducting-PolymerC60 Composites: Photoinduced Electron Transfer. Phys. Rev. B 1993, 47, 13835–13842.
  67. Sariciftci, N.S.; Smilowitz, L.; Heeger, A.J.; Wudl, F. Photoinduced Electron Transfer from a Conducting Polymer to Buckminsterfullerene. Science 1992, 258, 1474–1476.
  68. Morita, S.; Kiyomatsu, S.; Yin, X.H.; Zakhidov, A.A.; Noguchi, T.; Ohnishi, T.; Yoshino, K. Doping Effect of Buckminsterfullerene in Poly (2,5-dialkoxy-p-phenylene Vinylene). J. Appl. Phys. 1993, 74, 2860–2865.
  69. Roman, L.S.; Mammo, W.; Pettersson, L.A.A.; Andersson, M.R.; Inganäs, O. High Quantum Efficiency Polythiophene. Adv. Mater. 1998, 10, 774–777.
  70. Halls, J.J.M.; Pichler, K.; Friend, R.H.; Moratti, S.C.; Holmes, A.B. Exciton Diffusion and Dissociation in a Poly (P-phenylenevinylene)/C60 Heterojunction Photovoltaic Cell. Appl. Phys. Lett. 1996, 68, 3120–3122.
  71. Yang, C.Y.; Heeger, A.J. Morphology of Composites of Semiconducting Polymers Mixed with C60. Synth. Met. 1996, 83, 85–88.
  72. Yu, G.; Gao, J.; Hummelen, J.C.; Wudl, F.; Heeger, A.J. Polymer Photovoltaic Cells: Enhanced Efficiencies via a Network of Internal Donor-Acceptor Heterojunctions. Science 1995, 270, 1789–1791.
  73. Gnida, P.; Amin, M.F.; Pajak, A.K.; Jarzabek, B. Polymers in High-Efficiency Solar Cells: The Latest Reports. Polymers 2022, 14, 1946.
  74. Labiod, A.; Ibraikulov, O.A.; Dabos-Seignon, S.; Ferry, S.; Heinrich, B.; Méry, S.; Fall, S.; Tchognia Nkuissi, H.J.; Heiser, T.; Cabanetos, C.; et al. Photo-Degradation in Bulk Heterojunction Organic Solar Cells Using a Fullerene or a Non-Fullerene Derivative Electron Acceptor. Org. Electron. 2022, 107, 106549.
  75. Kaim, A.; Piotrowski, P.; Zarębska, K.; Bogdanowicz, K.A.; Przybył, W.; Kwak, A.; Skompska, M.; Gnida, P.; Schab-Balcerzak, E.; Iwan, A. Thermal Imaging and Deep Optical and Electrochemical Study of C70 Fullerene Derivatives with Thiophene, Pyrrolidine or Indene Moieties along with Electropolymerization with Thiophene Substituted Imine: Blends with P3HT and PTB7. Electrochim. Acta 2022, 426, 140741.
  76. Liang, S.; Xiao, C.; Xie, C.; Liu, B.; Fang, H.; Li, W. 13% Single-Component Organic Solar Cells Based on Double-Cable Conjugated Polymers with Pendent Y-Series Acceptors. Adv. Mater. 2023, 35, e2300629.
  77. Shaheen, S.E.; Brabec, C.J.; Sariciftci, N.S.; Padinger, F.; Fromherz, T.; Hummelen, J.C. 2.5% Efficient Organic Plastic Solar Cells. Appl. Phys. Lett. 2001, 78, 841–843.
  78. Levitsky, A.; Schneider, S.A.; Rabkin, E.; Toney, M.F.; Frey, G.L. Bridging the Thermodynamics and Kinetics of Temperature-Induced Morphology Evolution in Polymer/Fullerene Organic Solar Cell Bulk Heterojunction. Mater. Horizons 2021, 8, 1272–1285.
  79. Sprau, C.; Kattenbusch, J.; Li, Y.; Müller, E.; Gerthsen, D.; Berger, R.; Michels, J.J.; Colsmann, A. Revisiting Solvent Additives for the Fabrication of Polymer:Fullerene Solar Cells: Exploring a Series of Benzaldehydes. Sol. RRL 2021, 5, 2100238.
  80. Grodniski, D.C.; Benatto, L.; Gonçalves, J.P.; de Oliveira, C.C.; Pacheco, K.R.M.; Adad, L.B.; Coturi, V.M.; Roman, L.S.; Koehler, M. High Photothermal Conversion Efficiency for Semiconducting Polymer/Fullerene Nanoparticles and Its Correlation with Photoluminescence Quenching. RSC Chem. Biol. 2022, 4, 486–503.
  81. Li, S.; Hamada, F.; Nishikubo, R.; Saeki, A. Quantifying the Optimal Thickness in Polymer:Fullerene Solar Cells from the Analysis of Charge Transport Dynamics and Photoabsorption. Sustain. Energy Fuels 2022, 6, 756–765.
  82. Hoppe, H.; Sariciftci, N.S. Organic Solar Cells: An Overview. J. Mater. Res. 2004, 19, 1924–1945.
  83. Hou, J.; Inganäs, O.; Friend, R.H.; Gao, F. Organic Solar Cells Based on Non-Fullerene Acceptors. Nat. Mater. 2018, 17, 119–128.
  84. Kietzke, T. Recent Advances in Organic Solar Cells. Adv. Optoelectron. 2007, 2007, 40285.
  85. Kim, H.U.; Kim, J.-H.; Kang, H.; Grimsdale, A.C.; Kim, B.J.; Yoon, S.C.; Hwang, D.-H. Naphthalene-, Anthracene-, and Pyrene-Substituted Fullerene Derivatives as Electron Acceptors in Polymer-Based Solar Cells. ACS Appl. Mater. Interfaces 2014, 6, 20776–20785.
  86. Zhao, G.; He, Y.; Xu, Z.; Hou, J.; Zhang, M.; Min, J.; Chen, H.-Y.; Ye, M.; Hong, Z.; Yang, Y.; et al. Effect of Carbon Chain Length in the Substituent of PCBM-like Molecules on Their Photovoltaic Properties. Adv. Funct. Mater. 2010, 20, 1480–1487.
  87. Mikroyannidis, J.A.; Kabanakis, A.N.; Sharma, S.S.; Sharma, G.D. A Simple and Effective Modification of PCBM for Use as an Electron Acceptor in Efficient Bulk Heterojunction Solar Cells. Adv. Funct. Mater. 2011, 21, 746–755.
  88. Xu, X.; Yu, T.; Bi, Z.; Ma, W.; Li, Y.; Peng, Q. Realizing Over 13% Efficiency in Green-Solvent-Processed Nonfullerene Organic Solar Cells Enabled by 1,3,4-Thiadiazole-Based Wide-Bandgap Copolymers. Adv. Mater. 2018, 30, 1703973.
  89. Lin, Y.; Zhao, F.; He, Q.; Huo, L.; Wu, Y.; Parker, T.C.; Ma, W.; Sun, Y.; Wang, C.; Zhu, D.; et al. High-Performance Electron Acceptor with Thienyl Side Chains for Organic Photovoltaics. J. Am. Chem. Soc. 2016, 138, 4955–4961.
  90. Li, T.; Dai, S.; Ke, Z.; Yang, L.; Wang, J.; Yan, C.; Ma, W.; Zhan, X. Fused Tris (Thienothiophene)-Based Electron Acceptor with Strong Near-Infrared Absorption for High-Performance As-Cast Solar Cells. Adv. Mater. 2018, 30, 1705969.
  91. Liu, Q.; Jiang, Y.; Jin, K.; Qin, J.; Xu, J.; Li, W.; Xiong, J.; Liu, J.; Xiao, Z.; Sun, K.; et al. 18% Efficiency Organic Solar Cells. Sci. Bull. 2020, 65, 272–275.
  92. Zhang, S.; Ye, L.; Hou, J. Breaking the 10% Efficiency Barrier in Organic Photovoltaics: Morphology and Device Optimization of Well-Known PBDTTT Polymers. Adv. Energy Mater. 2016, 6, 1502529.
  93. Fu, Y.; Lee, T.H.; Chin, Y.-C.; Pacalaj, R.A.; Labanti, C.; Park, S.Y.; Dong, Y.; Cho, H.W.; Kim, J.Y.; Minami, D.; et al. Molecular Orientation-Dependent Energetic Shifts in Solution-Processed Non-Fullerene Acceptors and Their Impact on Organic Photovoltaic Performance. Nat. Commun. 2023, 14, 1870.
  94. Wang, Y.; Lee, J.; Hou, X.; Labanti, C.; Yan, J.; Mazzolini, E.; Parhar, A.; Nelson, J.; Kim, J.-S.; Li, Z. Recent Progress and Challenges toward Highly Stable Nonfullerene Acceptor-Based Organic Solar Cells. Adv. Energy Mater. 2021, 11, 2003002.
  95. Lan, W.; Gu, J.; Wu, S.; Peng, Y.; Zhao, M.; Liao, Y.; Xu, T.; Wei, B.; Ding, L.; Zhu, F. Toward Improved Stability of Nonfullerene Organic Solar Cells: Impact of Interlayer and Built-in Potential. EcoMat 2021, 3, e12134.
  96. Wang, X.; Zhao, C.X.; Xu, G.; Chen, Z.K.; Zhu, F. Degradation Mechanisms in Organic Solar Cells: Localized Moisture Encroachment and Cathode Reaction. Sol. Energy Mater. Sol. Cells 2012, 104, 1–6.
  97. Fu, J.; Fong, P.W.K.; Liu, H.; Huang, C.-S.; Lu, X.; Lu, S.; Abdelsamie, M.; Kodalle, T.; Sutter-Fella, C.M.; Yang, Y.; et al. 19.31% Binary Organic Solar Cell and Low Non-Radiative Recombination Enabled by Non-Monotonic Intermediate State Transition. Nat. Commun. 2023, 14, 1760.
  98. Tarique, W.B.; Uddin, A. A Review of Progress and Challenges in the Research Developments on Organic Solar Cells. Mater. Sci. Semicond. Process. 2023, 163, 107541.
  99. Adnan, M.; Irshad, Z.; Hussain, R.; Lee, W.; Kim, M.; Lim, J. Efficient Ternary Active Layer Materials for Organic Photovoltaics. Sol. Energy 2023, 257, 324–343.
  100. Li, D.; Deng, N.; Fu, Y.; Guo, C.; Zhou, B.; Wang, L.; Zhou, J.; Liu, D.; Li, W.; Wang, K.; et al. Fibrillization of Non-Fullerene Acceptors Enables 19% Efficiency Pseudo-Bulk Heterojunction Organic Solar Cells. Adv. Mater. 2023, 35, 2208211.
  101. Zhang, Y.; Lang, Y.; Li, G. Recent Advances of Non-Fullerene Organic Solar Cells: From Materials and Morphology to Devices and Applications. EcoMat 2023, 5, e12281.
  102. Gu, X.; Lai, X.; Zhang, Y.; Wang, T.; Tan, W.L.; McNeill, C.R.; Liu, Q.; Sonar, P.; He, F.; Li, W.; et al. Organic Solar Cell With Efficiency Over 20% and VOC Exceeding 2.1 V Enabled by Tandem with All-Inorganic Perovskite and Thermal Annealing-Free Process. Adv. Sci. 2022, 9, e2200445.
  103. Zhu, L.; Zhang, M.; Xu, J.; Li, C.; Yan, J.; Zhou, G.; Zhong, W.; Hao, T.; Song, J.; Xue, X.; et al. Single-Junction Organic Solar Cells with over 19% Efficiency Enabled by a Refined Double-Fibril Network Morphology. Nat. Mater. 2022, 21, 656–663.
  104. Cui, Y.; Xu, Y.; Yao, H.; Bi, P.; Hong, L.; Zhang, J.; Zu, Y.; Zhang, T.; Qin, J.; Ren, J.; et al. Single-Junction Organic Photovoltaic Cell with 19% Efficiency. Adv. Mater. 2021, 33, 2102420.
  105. Grätzel, M. Photoelectrochemical Cells. Nature 2001, 414, 338–344.
  106. Kalyanasundaram, K.; Grätzel, M. Applications of Functionalized Transition Metal Complexes in Photonic and Optoelectronic Devices. Coord. Chem. Rev. 1998, 177, 347–414.
  107. Grätzel, M. Dye-Sensitized Solar Cells. J. Photochem. Photobiol. C Photochem. Rev. 2003, 4, 145–153.
  108. Grobelny, A.; Shen, Z.; Eickemeyer, F.T.; Antariksa, N.F.; Zapotoczny, S.; Zakeeruddin, S.M.; Grätzel, M. A Molecularly Tailored Photosensitizer with an Efficiency of 13.2% for Dye-Sensitized Solar Cells. Adv. Mater. 2023, 35, 2207785.
  109. Ren, Y.; Cao, Y.; Zhang, D.; Zakeeruddin, S.M.; Hagfeldt, A.; Wang, P.; Grätzel, M. A Blue Photosensitizer Realizing Efficient and Stable Green Solar Cells via Color Tuning by the Electrolyte. Adv. Mater. 2020, 32, 2000193.
  110. Dou, J.; Chen, Q. MOFs in Emerging Solar Cells. Chin. J. Chem. 2023, 41, 695–709.
  111. Bandara, T.M.W.J.; Hansadi, J.M.C.; Bella, F. A Review of Textile Dye-Sensitized Solar Cells for Wearable Electronics. Ionics 2022, 28, 2563–2583.
  112. Devadiga, D.; Selvakumar, M.; Shetty, P.; Santosh, M.S. The Integration of Flexible Dye-Sensitized Solar Cells and Storage Devices towards Wearable Self-Charging Power Systems: A Review. Renew. Sustain. Energy Rev. 2022, 159, 112252.
  113. Bi, P.; Wang, J.; Cui, Y.; Zhang, J.; Zhang, T.; Chen, Z.; Qiao, J.; Dai, J.; Zhang, S.; Hao, X. Enhancing Photon Utilization Efficiency for High-Performance Organic Photovoltaic Cells via Regulating Phase-Transition Kinetics. Adv. Mater. 2023, 35, e2210865.
  114. Zheng, Z.; Wang, J.; Bi, P.; Ren, J.; Wang, Y.; Yang, Y.; Liu, X.; Zhang, S.; Hou, J. Tandem Organic Solar Cell with 20.2% Efficiency. Joule 2022, 6, 171–184.
  115. Barbato, M.; Artegiani, E.; Bertoncello, M.; Meneghini, M.; Ortolani, L.; Zanoni, E.; Meneghesso, G. CdTe Solar Cells: Technology, Operation and Reliability. J. Phys. D Appl. Phys. 2021, 54, 333002.
  116. Moon, S.H.; Park, S.J.; Kim, S.H.; Lee, M.W.; Han, J.; Kim, J.Y.; Kim, H.; Hwang, Y.J.; Lee, D.K.; Min, B.K. Monolithic DSSC/CIGS Tandem Solar Cell Fabricated by a Solution Process. Sci. Rep. 2015, 5, srep08970.
  117. Hadipour, A.; de Boer, B.; Blom, P.W.M. Device Operation of Organic Tandem Solar Cells. Org. Electron. 2008, 9, 617–624.
  118. Enam, F.M.T.; Rahman, K.S.; Kamaruzzaman, M.I.; Sobayel, K.; Chelvanathan, P.; Bais, B.; Akhtaruzzaman, M.; Alamoud, A.R.M.; Amin, N. Design Prospects of Cadmium Telluride/Silicon (CdTe/Si) Tandem Solar Cells from Numerical Simulation. Optik 2017, 139, 397–406.
  119. Hosokawa, H.; Tamaki, R.; Sawada, T.; Okonogi, A.; Sato, H.; Ogomi, Y.; Hayase, S.; Okada, Y.; Yano, T. Solution-Processed Intermediate-Band Solar Cells with Lead Sulfide Quantum Dots and Lead Halide Perovskites. Nat. Commun. 2019, 10, 4–6.
  120. Andruszkiewicz, A.; Zhang, X.; Johansson, M.B.; Yuan, L.; Johansson, E.M.J. Perovskite and Quantum Dot Tandem Solar Cells with Interlayer Modification for Improved Optical Semitransparency and Stability. Nanoscale 2021, 13, 6234–6240.
  121. Todorov, T.; Gershon, T.; Gunawan, O.; Sturdevant, C.; Guha, S. Perovskite-Kesterite Monolithic Tandem Solar Cells with High Open-Circuit Voltage. Appl. Phys. Lett. 2014, 105, 173902.
  122. Saha, U.; Alam, M.K. Proposition and Computational Analysis of a Kesterite/Kesterite Tandem Solar Cell with Enhanced Efficiency. RSC Adv. 2017, 7, 4806–4814.
  123. Greenaway, A.L.; Loutris, A.L.; Heinselman, K.N.; Melamed, C.L.; Schnepf, R.R.; Tellekamp, M.B.; Woods-Robinson, R.; Sherbondy, R.; Bardgett, D.; Bauers, S.; et al. Combinatorial Synthesis of Magnesium Tin Nitride Semiconductors. J. Am. Chem. Soc. 2020, 142, 8421–8430.
  124. Sun, W.; Bartel, C.J.; Arca, E.; Bauers, S.R.; Matthews, B.; Orvañanos, B.; Chen, B.R.; Toney, M.F.; Schelhas, L.T.; Tumas, W.; et al. A Map of the Inorganic Ternary Metal Nitrides. Nat. Mater. 2019, 18, 732–739.
  125. Choi, J.W.; Shin, B.; Gorai, P.; Hoye, R.L.Z.; Palgrave, R. Emerging Earth-Abundant Solar Absorbers. ACS Energy Lett. 2022, 7, 1553–1557.
  126. Javaid, K.; Wu, W.; Wang, J.; Fang, J.; Zhang, H.; Gao, J.; Zhuge, F.; Liang, L.; Cao, H. Band Offset Engineering in ZnSnN2-Based Heterojunction for Low-Cost Solar Cells. ACS Photonics 2018, 5, 2094–2099.
  127. 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.
  128. Tien, C.H.; Lai, H.Y.; Chen, L.C. Methylammonium Halide Salt Interfacial Modification of Perovskite Quantum Dots/Triple-Cation Perovskites Enable Efficient Solar Cells. Sci. Rep. 2023, 13, 5387.
  129. Dey, A.; Ye, J.; De, A.; Debroye, E.; Ha, S.K.; Bladt, E.; Kshirsagar, A.S.; Wang, Z.; Yin, J.; Wang, Y.; et al. State of the Art and Prospects for Halide Perovskite Nanocrystals. ACS Nano 2021, 15, 10775–10981.
  130. Feng, B.; Duan, J.; Tao, L.; Zhang, J.; Wang, H. Enhanced Performance in Perovskite Solar Cells via Bromide Ion Substitution and Ethanol Treatment. Appl. Surf. Sci. 2018, 430, 603–612.
  131. Niu, T.; Lu, J.; Tang, M.-C.; Barrit, D.; Smilgies, D.-M.; Yang, Z.; Li, J.; Fan, Y.; Luo, T.; McCulloch, I.; et al. High Performance Ambient-Air-Stable FAPbI3 Perovskite Solar Cells with Molecule-Passivated Ruddlesden–Popper/3D Heterostructured Film. Energy Environ. Sci. 2018, 11, 3358–3366.
  132. Jeong, J.; Kim, M.; Seo, J.; Lu, H.; Ahlawat, P.; Mishra, A.; Yang, Y.; Hope, M.A.; Eickemeyer, F.T.; Kim, M.; et al. Pseudo-Halide Anion Engineering for α-FAPbI3 Perovskite Solar Cells. Nature 2021, 592, 381–385.
  133. Hossain, M.K.; Toki, G.F.I.; Kuddus, A.; Rubel, M.H.K.; Hossain, M.M.; Bencherif, H.; Rahman, M.F.; Islam, M.R.; Mushtaq, M. An Extensive Study on Multiple ETL and HTL Layers to Design and Simulation of High-Performance Lead-Free CsSnCl3-Based Perovskite Solar Cells. Sci. Rep. 2023, 13, 2521.
  134. Juarez-Perez, E.J.; Wuβler, M.; Fabregat-Santiago, F.; Lakus-Wollny, K.; Mankel, E.; Mayer, T.; Jaegermann, W.; Mora-Sero, I. Role of the Selective Contacts in the Performance of Lead Halide Perovskite Solar Cells. J. Phys. Chem. Lett. 2014, 5, 680–685.
  135. Li, F.; Zhou, S.; Yuan, J.; Qin, C.; Yang, Y.; Shi, J.; Ling, X.; Li, Y.; Ma, W. Perovskite Quantum Dot Solar Cells with 15.6% Efficiency and Improved Stability Enabled by an α-CsPbI3/FAPbI3 Bilayer Structure. ACS Energy Lett. 2019, 4, 2571–2578.
  136. Yao, D.; Zhang, C.; Pham, N.D.; Zhang, Y.; Tiong, V.T.; Du, A.; Shen, Q.; Wilson, G.J.; Wang, H. Hindered Formation of Photoinactive δ-FAPbI3 Phase and Hysteresis-Free Mixed-Cation Planar Heterojunction Perovskite Solar Cells with Enhanced Efficiency via Potassium Incorporation. J. Phys. Chem. Lett. 2018, 9, 2113–2120.
  137. Xue, J.; Lee, J.-W.; Dai, Z.; Wang, R.; Nuryyeva, S.; Liao, M.E.; Chang, S.-Y.; Meng, L.; Meng, D.; Sun, P.; et al. Surface Ligand Management for Stable FAPbI3 Perovskite Quantum Dot Solar Cells. Joule 2018, 2, 1866–1878.
  138. Sha, W.E.I.; Ren, X.; Chen, L.; Choy, W.C.H. The Efficiency Limit of CH3NH3PbI3 Perovskite Solar Cells. Appl. Phys. Lett. 2015, 106, 221104.
  139. Masi, S.; Gualdrón-Reyes, A.F.; Mora-Seró, I. Stabilization of Black Perovskite Phase in FAPbI3 and CsPbI3. ACS Energy Lett. 2020, 5, 1974–1985.
  140. Zhang, H.; Li, Y.; Tan, S.; Chen, Z.; Song, K.; Huang, S.; Shi, J.; Luo, Y.; Li, D.; Meng, Q. High-Efficiency (>20%) Planar Carbon-Based Perovskite Solar Cells through Device Configuration Engineering. J. Colloid Interface Sci. 2022, 608, 3151–3158.
  141. Kim, M.; Kim, G.H.; Lee, T.K.; Choi, I.W.; Choi, H.W.; Jo, Y.; Yoon, Y.J.; Kim, J.W.; Lee, J.; Huh, D.; et al. Methylammonium Chloride Induces Intermediate Phase Stabilization for Efficient Perovskite Solar Cells. Joule 2019, 3, 2179–2192.
  142. Yin, J.; Xu, Z.; Hu, Q.; Teobaldi, G.; Liu, L.-M.; Prezhdo, O.V. Tuning Octahedral Tilting by Doping to Prevent Detrimental Phase Transition and Extend Carrier Lifetime in Organometallic Perovskites. J. Am. Chem. Soc. 2023, 145, 5393–5399.
  143. Wang, A.; Wang, J.; Niu, X.; Zuo, C.; Hao, F.; Ding, L. Inhibiting Octahedral Tilting for Stable CsPbI2Br Solar Cells. InfoMat 2022, 4, 2–9.
  144. Menahem, M.; Dai, Z.; Aharon, S.; Sharma, R.; Asher, M.; Diskin-Posner, Y.; Korobko, R.; Rappe, A.M.; Yaffe, O. Strongly Anharmonic Octahedral Tilting in Two-Dimensional Hybrid Halide Perovskites. ACS Nano 2021, 15, 10153–10162.
  145. Zhao, Q.; Hazarika, A.; Schelhas, L.T.; Liu, J.; Gaulding, E.A.; Li, G.; Zhang, M.; Toney, M.F.; Sercel, P.C.; Luther, J.M. Size-Dependent Lattice Structure and Confinement Properties in CsPbI3 Perovskite Nanocrystals: Negative Surface Energy for Stabilization. ACS Energy Lett. 2019, 5, 238–247.
  146. Doherty, T.A.S.; Nagane, S.; Kubicki, D.J.; Jung, Y.K.; Johnstone, D.N.; Iqbal, A.N.; Guo, D.; Frohna, K.; Danaie, M.; Tennyson, E.M.; et al. Stabilized Tilted-Octahedra Halide Perovskites Inhibit Local Formation of Performance-Limiting Phases. Science 2021, 374, 1598–1605.
  147. Wang, R.; Xue, J.; Meng, L.; Lee, J.W.; Zhao, Z.; Sun, P.; Cai, L.; Huang, T.; Wang, Z.; Wang, Z.K.; et al. Caffeine Improves the Performance and Thermal Stability of Perovskite Solar Cells. Joule 2019, 3, 1464–1477.
  148. Wang, R.; Xue, J.; Wang, K.L.; Wang, Z.K.; Luo, Y.; Fenning, D.; Xu, G.; Nuryyeva, S.; Huang, T.; Zhao, Y.; et al. Constructive Molecular Configurations for Surface-Defect Passivation of Perovskite Photovoltaics. Science 2019, 366, 1509–1513.
  149. Xie, H.; Wang, Z.; Chen, Z.; Pereyra, C.; Pols, M.; Gałkowski, K.; Anaya, M.; Fu, S.; Jia, X.; Tang, P.; et al. Decoupling the Effects of Defects on Efficiency and Stability through Phosphonates in Stable Halide Perovskite Solar Cells. Joule 2021, 5, 1246–1266.
  150. Feng, W.; Tan, Y.; Yang, M.; Jiang, Y.; Lei, B.X.; Wang, L.; Wu, W.Q. Small Amines Bring Big Benefits to Perovskite-Based Solar Cells and Light-Emitting Diodes. Chem 2022, 8, 351–383.
  151. Yang, T.; Gao, L.; Lu, J.; Ma, C.; Du, Y.; Wang, P.; Ding, Z.; Wang, S.; Xu, P.; Liu, D.; et al. One-Stone-for-Two-Birds Strategy to Attain beyond 25% Perovskite Solar Cells. Nat. Commun. 2023, 14, 839.
  152. Li, G.; Song, J.; Wu, J.; Song, Z.; Wang, X.; Sun, W.; Fan, L.; Lin, J.; Huang, M.; Lan, Z.; et al. Efficient and STable 2D@3D/2D Perovskite Solar Cells Based on Dual Optimization of Grain Boundary and Interface. ACS Energy Lett. 2021, 6, 3614–3623.
  153. Campos, T.; Dally, P.; Gbegnon, S.; Blaizot, A.; Trippé-Allard, G.; Provost, M.; Bouttemy, M.; Duchatelet, A.; Garrot, D.; Rousset, J.; et al. Unraveling the Formation Mechanism of the 2D/3D Perovskite Heterostructure for Perovskite Solar Cells Using Multi-Method Characterization. J. Phys. Chem. C 2022, 126, 13527–13538.
  154. Weidman, M.C.; Seitz, M.; Stranks, S.D.; Tisdale, W.A. Highly Tunable Colloidal Perovskite Nanoplatelets through Variable Cation, Metal, and Halide Composition. ACS Nano 2016, 10, 7830–7839.
  155. Zhou, T.; Lai, H.; Liu, T.; Lu, D.; Wan, X.; Zhang, X.; Liu, Y.; Chen, Y. Highly Efficient and Stable Solar Cells Based on Crystalline Oriented 2D/3D Hybrid Perovskite. Adv. Mater. 2019, 31, e1901242.
  156. Sutanto, A.A.; Caprioglio, P.; Drigo, N.; Hofstetter, Y.J.; Garcia-Benito, I.; Queloz, V.I.E.; Neher, D.; Nazeeruddin, M.K.; Stolterfoht, M.; Vaynzof, Y.; et al. 2D/3D Perovskite Engineering Eliminates Interfacial Recombination Losses in Hybrid Perovskite Solar Cells. Chem 2021, 7, 1903–1916.
  157. Cho, Y.; Soufiani, A.M.; Yun, J.S.; Kim, J.; Lee, D.S.; Seidel, J.; Deng, X.; Green, M.A.; Huang, S.; Ho-Baillie, A.W.Y. Mixed 3D–2D Passivation Treatment for Mixed-Cation Lead Mixed-Halide Perovskite Solar Cells for Higher Efficiency and Better Stability. Adv. Energy Mater. 2018, 8, 1703392.
  158. Li, H.; Zhang, W. Perovskite Tandem Solar Cells: From Fundamentals to Commercial Deployment. Chem. Rev. 2020, 120, 9835–9950.
  159. Al-Ashouri, A.; Köhnen, E.; Li, B.; Magomedov, A.; Hempel, H.; Caprioglio, P.; Márquez, J.A.; Vilches, A.B.M.; Kasparavicius, E.; Smith, J.A.; et al. Monolithic Perovskite/Silicon Tandem Solar Cell with >29% Efficiency by Enhanced Hole Extraction. Science 2020, 370, 1300–1309.
  160. Wang, J.; Uddin, M.A.; Chen, B.; Ying, X.; Ni, Z.; Zhou, Y.; Li, M.; Wang, M.; Yu, Z.; Huang, J. Enhancing Photostability of Sn-Pb Perovskite Solar Cells by an Alkylammonium Pseudo-Halogen Additive. Adv. Energy Mater. 2023, 13, 2204115.
  161. Ren, M.; Qian, X.; Chen, Y.; Wang, T.; Zhao, Y. Potential Lead Toxicity and Leakage Issues on Lead Halide Perovskite Photovoltaics. J. Hazard. Mater. 2022, 426, 127848.
  162. Torrence, C.E.; Libby, C.S.; Nie, W.; Stein, J.S. Environmental and Health Risks of Perovskite Solar Modules: Case for Better Test Standards and Risk Mitigation Solutions. iScience 2023, 26, 105807.
  163. Aftab, A.; Ahmad, M.I. A Review of Stability and Progress in Tin Halide Perovskite Solar Cell. Sol. Energy 2021, 216, 26–47.
  164. Mahmoudi, T.; Rho, W.Y.; Kohan, M.; Im, Y.H.; Mathur, S.; Hahn, Y.B. Suppression of Sn2+/Sn4+ Oxidation in Tin-Based Perovskite Solar Cells with Graphene-Tin Quantum Dots Composites in Active Layer. Nano Energy 2021, 90, 106495.
  165. Chen, Z.; Liu, M.; Li, Z.; Shi, T.; Yang, Y.; Yip, H.L.; Cao, Y. Stable Sn/Pb-Based Perovskite Solar Cells with a Coherent 2D/3D Interface. iScience 2018, 9, 337–346.
  166. Ban, H.; Sun, Q.; Zhang, T.; Li, H.; Shen, Y.; Wang, M. Stabilization of Inorganic CsPb0.5Sn0.5I2Br Perovskite Compounds by Antioxidant Tea Polyphenol. Sol. RRL 2020, 4, 1900457.
  167. Li, C.; Pan, Y.; Hu, J.; Qiu, S.; Zhang, C.; Yang, Y.; Chen, S.; Liu, X.; Brabec, C.J.; Khaja Nazeeruddin, M.; et al. Vertically Aligned 2D/3D Pb–Sn Perovskites with Enhanced Charge Extraction and Suppressed Phase Segregation for Efficient Printable Solar Cells. ACS Energy Lett. 2020, 5, 1386–1395.
  168. Zhou, X.; Zhang, L.; Wang, X.; Liu, C.; Chen, S.; Zhang, M.; Li, X.; Yi, W.; Xu, B. Highly Efficient and Stable GABr-Modified Ideal-Bandgap (1.35 EV) Sn/Pb Perovskite Solar Cells Achieve 20.63% Efficiency with a Record Small Voc Deficit of 0.33 V. Adv. Mater. 2020, 32, e1908107.
  169. Sanchez-Diaz, J.; Sánchez, R.S.; Masi, S.; Kreĉmarová, M.; Alvarez, A.O.; Barea, E.M.; Rodriguez-Romero, J.; Chirvony, V.S.; Sánchez-Royo, J.F.; Martinez-Pastor, J.P.; et al. Tin Perovskite Solar Cells with >1300 h of Operational Stability in N2 through a Synergistic Chemical Engineering Approach. Joule 2022, 6, 861–883.
  170. Zhu, Z.; Jiang, X.; Yu, D.; Yu, N.; Ning, Z.; Mi, Q. Smooth and Compact FASnI3 Films for Lead-Free Perovskite Solar Cells with over 14% Efficiency. ACS Energy Lett. 2022, 7, 2079–2083.
  171. Yu, B.B.; Chen, Z.; Zhu, Y.; Wang, Y.; Han, B.; Chen, G.; Zhang, X.; Du, Z.; He, Z. Heterogeneous 2D/3D Tin-Halides Perovskite Solar Cells with Certified Conversion Efficiency Breaking 14%. Adv. Mater. 2021, 33, e2102055.
More
Information
Subjects: Others
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , , ,
View Times: 242
Revisions: 2 times (View History)
Update Date: 16 Aug 2023
1000/1000
Video Production Service