Next Article in Journal
Galvanic Deposition of Calcium Phosphate/Bioglass Composite Coating on AISI 316L
Next Article in Special Issue
Dual Strategy Based on Quantum Dot Doping and Phenylethylamine Iodide Surface Modification for High-Performance and Stable Perovskite Solar Cells
Previous Article in Journal
Coating Condition Detection and Assessment on the Steel Girder of a Bridge through Hyperspectral Imaging
Previous Article in Special Issue
Investigation on Film Quality and Photophysical Properties of Narrow Bandgap Molecular Semiconductor Thin Film and Its Solar Cell Application
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Lead Leakage of Pb-Based Perovskite Solar Cells

1
School of Chemistry and Chemical Engineering, Henan Normal University, Xinxiang 453007, China
2
Henan Key Laboratory of Photovoltaic Materials, School of Physics, Henan Normal University, Xinxiang 453007, China
3
College of Material Science and Engineering, Henan Normal University, Xinxiang 453007, China
*
Authors to whom correspondence should be addressed.
Coatings 2023, 13(6), 1009; https://doi.org/10.3390/coatings13061009
Submission received: 7 April 2023 / Revised: 24 May 2023 / Accepted: 26 May 2023 / Published: 29 May 2023
(This article belongs to the Special Issue Perovskite Photovoltaics: From Materials to Device Applications)

Abstract

:
As a novel technology, perovskite solar cells (PSCs) have attracted worldwide attention due to their high photoelectric conversion efficiency (PCE) and low fabricating cost. Moreover, with the development of this technology, PSCs have achieved a great breakthrough in PCE. However, the heavy metal element Pb in PSCs does harm to human health and ecological environments, which restricts the further application of Pb-based PSCs. Under certain circumstances, the leakage of lead will cause serious pollution to the environment. The purpose of this review is to summary and discuss the way of lead leakage suppression. Among them, we pay more attention to the method of packaging technology, chemisorption procession and the limitations of each method. Finally, strategies of highly PCE and non-toxic perovskite devices are proposed.

1. Introduction

Perovskite solar cells (PSCs) have attracted wide attention from scientists due to their excellent photoelectric characteristics, such as long carrier length, high carrier mobility, high absorption coefficient, and low trap density. In terms of manufacturing costs, perovskite solar cells have low prices because of the low cost of rich materials (Ti, Pb, I, Cl, Br, etc.), making them the most competitive technology in the solar cell industry [1,2,3]. The power conversion efficiency (PCE) of single-junction perovskite solar cells has exceeded 25%, surpassing the record set by copper indium gallium selenium (CIGS) solar cells and approaching that of the best crystalline silicon solar cells [4,5,6]. Furthermore, the PCE of perovskite/silicon tandem solar cells have exceeded 29% [7,8,9,10,11]. Increasing teams are devoted to the research of PSCs, which undoubtedly accelerates the further commercial development of PSCs.
Lead element is often involved in PSCs thanks to their high crystal symmetry, unique atomic electron fusion, and strong spin orbital coupling capacity. In general, the common device structure containing Pb is APbX3, where A is an organic or inorganic cation (such as methylammonium, formamidine, or Cs+) and X is a halide ion (mainly I or Br) [12,13,14]. The results show that the Pb-based perovskite possessed the highest photoelectric conversion efficiency in all the perovskite solar cells [13,15]. Although the photovoltaic properties of Pb-based perovskite solar modules are excellent, the safety hazard brought by them cannot be ignored at the same time. It is worth noting that Pb-based perovskite contains much water-soluble lead salts as degradation products [16]. Even a small amount of Pb leaking into the environment is toxic to human health and the environment. Lead or its compounds can be absorbed by the human body through breathing, eating, skin absorption, and the like. According to the relevant studies, 20% to 80% of the ingested lead can be absorbed by a human body, and children have a higher absorption capacity of lead. Excessive lead intake can inhibit the normal synthesis of proteins and cause a number of healthy problems. For example, the increased levels of lead in the blood have a harmful effect on infants’ and children’s behavior, cognitive performance, pubertal development and hearing ability [17]. As for adults, lead can cause a series of cardiovascular, central nervous system, kidney and fertility problems. Worse still, lead can also stunt early fetal growth during pregnancy. In the meantime, lead also does harm to the environment, and can inhibit the growth of growing plants [17,18,19,20,21]. The main toxic substance released by perovskite solar cells when they are decomposed is lead iodide. To illustrate this point, Kwak et al. studied the toxicity of PbI2 to the embryos of two species of fish (zebrafish and medaka) and focused on the chemical speciation of PbI2 in the culture medium of embryos to characterize the toxicity of lead iodide to organisms. Research results reveal that the mortality, deformity, hatching failure, growth inhibition, and other pathological changes are increased in fish which are exposed to lead iodide [11,22]. Hasan Ul Banna et al. conducted a test of lead exposure on mice and found that Pb element can induce anxious behavior and memory learning disorders in mice. The memory learning disability of mice exposed to Pb could be improved by training, but that of mice exposed to Pb environment from the embryonic stage was always lower than the initial level of mice who without contact to Pb (shown in Figure 1). Mice exposed to lead as fetuses suffered more severe neurobehavioral changes and liver damage [23]. Therefore, in order to realize the commercial application of perovskite solar cells in a large area, we must first focus on solving the problem of lead toxicity [24].
Normally, the lead in perovskite crystals is contained in the PV module and does not leak into the environment. However, in extreme cases, such as hail, earthquakes, and extreme temperatures, PV modules can be damaged. Perovskite materials are affected by moisture and oxygen, and lead-based perovskites tend to release toxic lead iodide as degradation products that can be transmitted from soil to homes and/or drinking water [25,26]. Scientists have investigated the strategies to prevent lead leakage, holding that the ideal prevention strategy is to trap lead ions quickly. For example, perovskite can run for a long time when broken down by rain. In addition to protecting against outside factors such as rain, dust, or UV rays, there are many other effective ways to prevent Pb leaks. The encapsulation method can greatly increase the perovskite module’s ability to withstand the external impact force and protect the device module from damage. Chemisorption can realize the adsorption of Pb2+ in the process of Pb2+ leakage, and fix Pb2+ in the component to avoid the release of lead into the atmosphere. Nonetheless, these methods cannot fundamentally reduce the concentration of lead ions released into the environment, so the study of lead-free perovskite can approach the environmental pollution of lead ions from the root. In the following part of this paper, we mainly introduce these three methods to solve the lead leakage issue and put forward the prospect of further commercial application after solving the lead leakage problem.

2. Mechanism of Lead Leakage from Pb-Based Perovskite

Poor stability of the Pb-based perovskite solar cells is the main cause of the lead leakage. There are two factors affecting the stability of PSCs: internal factors and external factors. Internal instability factors include perovskite structure, transport layer materials, defects caused by ion migration, and degradation of perovskite materials. It can be improved by regulating the crystal structure of perovskite, controlling ion migration, passivation defects, and adopting relatively stable transport layer materials. External factors such as moisture and oxygen generally addressed through encapsulation technology. For example, HI generated by hydrolysis of CH3NH3PbI3 (MAPbI3) would decompose into I2 under ultraviolet or light irradiation, which further promotes the degradation process [27]. The decomposed Pb2+ exists in the form of PbI2, which easily leaked into the environment and caused environmental pollution [28]. Hailegnaw et al. proved that lead loss rate was as high as 72% after 5 min of rain [29].
CH 3 NH 3 PbI 3   H 2 O h v   CH 3 NH 3 I   +   PbI 2
CH 3 NH 3 I     CH 3 NH 2   +   HI
2 HI   U V   H 2   +   I 2
4 HI   +   O 2     2 I 2   +   2 H 2 O
Compared to MA, FA is a larger molecule with a smaller dipole moment. For stronger binding between FA and halides, FA-based perovskite has higher stability attributed by less halide ion migration [30]. However, when the temperature exceeds 230 °C, such as at the scene of a fire, decomposition still occurs, and the decomposition reaction is shown by the equation below:
FAPbI 3     PbI 2   +   FAI
It can be seen that PbI2 is the main decomposition product of MAPbI3 and FAPbI3 under hot and humid conditions [31]. Normally, the Pb-based perovskite solar cell is safe even if the PSCs degrades. However, in extreme conditions such as wind, snow, hail, fire etc., the PV module may be subjected to strong mechanical shocks, which would result in leak into the environment. Therefore, it is extremely important to find the ideal packaging material.

3. Encapsulation

The degradation of perovskite solar cells is mainly caused by oxygen and water [32,33]. Encapsulation is a common protection method that can inhibit moisture from entering equipment, reduce the oxygen and water penetration to an almost negligible level, and curb the outflow of decomposed lead elements. As a qualified packaging layer, it should have a high performance of shielding oxygen and moisture, and can work well under extreme weather to improve the life of a device [34,35,36]. At the same time, it should also have great thermal stability and good light transmittance. A few typical packaging methods to suppress lead leakage are given below.
Ethylene vinyl acetate (EVA), as a copolymer of ethylene and ethylene acetate, is a common solar cell packaging material (covering layer, pooping agent and substrate). Moreover, EVA also features desirable light transmittance and elasticity, adhesion strength with glass, melt fluidity, and other advantages. Accordingly, EVA is widely used as a packaging material, with nearly 80% of photovoltaic (PV) modules utilizing EVA as an encapsulation material. Common EVA encapsulation method is shown in Figure 2a. Crosslinked EVA wafer is characterized by favorable transparency, and its transmittance in 400~1100 nm is about 92.8%, which can help realize the protection of components without affecting the device efficiency [37,38]. The degree of crosslinking is an important index of EVA encapsulation agent, which represents the degree of crosslinking of a polymer chain. If the crosslinking degree is too high, EVA will become brittle and will not withstand the external impact force. By contrast, if the crosslinking degree is too low, the aging resistance will be reduced, and it cannot meet the requirement of creep resistance. Taken together, the appropriate crosslinking degree is in the range of 75%~90% [39]. Crosslinked EVA tablets have such advantages as low cost and good transmittance. However, as a packaging material, an EVA sheet still faces a serious problem: EVA sheet will be aged and degraded in a long-term ultraviolet and hot atmosphere (shown in Figure 2b) making the color of the polymer film change from transparent and colorless to yellow or brown [40]. This color change will greatly reduce the light absorption range and the efficiency of solar cells. Researchers have found that the performance of EVA can be improved by adding antioxidants, ultraviolet absorbers, and light stabilizers, but this means cost increase, robbing EVA sheet of the advantage of low cost. Moreover, when the temperature of the EVA-encapsulated battery over 120 °C, the perovskite material turns yellow, indicating that the perovskite layer is decomposed. This decomposition mechanism is attributed to acetic acid in EVA [21,41]. To reduce the affection of acetic acid in EVA on the perovskite layers, M. D. McGehee further replaced EVA with polyolefin (POE) [42]. However, it was found that both the perovskite absorber and the charge transport layer degraded at high temperatures, which leads to the photovoltaic characteristics of the cells significantly reducing at 140–160 °C. Therefore, the traditional POE and EVA seem not the idea package materials for perovskite solar cells. Compared with these two packaging materials, PU (polyurethane) shows its significant advantages. The PU packaging is carried out at a relatively mild temperature of 80 °C, at which temperature the perovskite absorption layer would not decompose. Moreover, the perovskite absorption layer packaging with PU was not destroyed even at 120 °C. Zhengyang Fu proved that there is no performance degradation observed after keeping the PU-encapsulated battery at 85 °C for more than 325 h [43]. In summary, PU can be used as a simple and effective way to package perovskite solar cell module.
In addition, Kim et al. found that there is no PbI2 generated after exposed MAPbI3 based perovskite solar cell to the air for 26 days when the Au electrodes of the device were attached by Poly(p-chloro-xylylene) [44]. Further study proved that poly(p-chloro-xylylene) encapsulation can effectively prevent water into the PSCs module, and then inhibit the decomposition of perovskite. In the same way, Lei Shi et al. encapsulate PSCs device by polyolefin on the top of the electrode. Then, the device was covered by a piece of glass. The experiment results indicate that the encapsulated equipment has no Pb precipitation for more than 1800 h in the humid heat environment. This method not only prevents the entry of moisture, but also inhibits the output of decomposing gases such as CH3I or NH3 [45]. Moreover, Jie Yin et al. used UV glue to wrap PSCs. The UV-vis absorption spectrum results showed that the Pb-based perovskite solar cells packaged by UV glue could remain their photoelectric properties over 5 years [21,41,46].
Thermosetting epoxy resin has high mechanical strength, excellent dimensional stability, and good chemical resistance, so it is widely used in the packaging of photovoltaic devices [47]. For example, Yan Jiang et al. [48] found that the key to reducing lead leakage was the self-healing ability of certain polymers when heated above their glass transition temperature and inferred that an increase in the polymer’s self-healing ability would significantly reduce lead leakage. Epoxy resin (ER) is another effective encapsulation material. The team demonstrated that ER’s self-healing properties and improved mechanical strength could effectively prevent lead leakage, and lead leakage from lead halide perovskite PV products can be reduced by a factor of 375 if properly packaged. The ER film can be formed by a mixture of diglycidyl ether bisphenol A (DGEBA), n-octylamine (OA), and diphenyl phenylenediamine (MXDA) diglyceride ether. Jiang Yan et al. studied self-healing ER (DGEBA: OA: MXDA = 4:2:1) membranes by physical cross-linking of DGEBA and OA and chemical cross-linking of MXDA and DGEBA. In a contrast test, the test subjects were divided into four groups: A, B, C, and D. The encapsulation method of the test was as followed: (A) No package, (B) top package with perovskite solar modules/UV resin/glass, (C) top and bottom package with glass/Surlyn resin/perovskite solar modules/UV resin/glass, and (D) top and bottom package with glass/epoxy resin/perovskite solar modules/UV resin/glass. By simulating the amount of lead leakage under extreme weather conditions, such as acid rain and hail, it proved that an ER film had a self-healing characteristic and could realize self-healing using the heat provided by the sun. Therefore, self-healing could be achieved during the operation of solar cells, thereby greatly reducing the amount of lead leakage. By simulating the extreme weather test data, it proved that the lead leakage rate of the perovskite solar cell (method D) was greatly reduced by ER thin film encapsulation. Compared with the encapsulation method using glass cover at the module edge (method B), the lead leakage rate of Pb could be decreased from 30 to 0.08 mg h−1m−2, and the lead leakage rate of the ER encapsulation method was reduced by 375 times (shown in Figure 3) [48]. Through comparison tests under four different weather conditions, the team noted that the lead leakage in group D was much smaller than that in other groups, which could effectively control the total leakage of self-healing polymer packaging under extreme weather conditions in slow response time. The high mechanical strength and self-healing mechanism of ER films can effectively reduce the leakage of lead in perovskite solar cells, which provides a new insight for the packaging method of perovskite solar cells. By applying self-healing and lead-adsorbed ion gel sealers to the front glass surface and between the electrodes and the packaging adhesive glass, the perovskite module can be physically prevented from seeping water into the perovskite module if the packaging glass is damaged, when chemically trapping lead that may leak [49].
Atomic layer deposition (ALD) technology looks like an ideal packaging method for the perovskite solar cells, which can prepare thin film encapsulation layer (TFE) with multilayer under the condition of adjusting appropriate parameters. The most common approach is depositing encapsulation layer with inorganic (Al2O3)/organic (pV3D3) alternately structure via ALD onto the top of the PSCs (Figure 4a) [50]. While the inorganic layer blocks the external environment, the organic layer enhances the flexibility of the film and increases the smoothness of the PSC surface (shown in Figure 4c). This special structure can effectively delay the penetration of water and oxygen, greatly improving the barrier performance of the packaging layer, minimizing equipment damage, and consequently improving the long-term stability of the device. However, an ignored problem is that long worktime of ALD at 90 °C will bring irreversible damage to PSCs, because this temperature is too high for depositing perovskite films (Figure 4b,d). Therefore, the research on the deposition of low-temperature ALD or even room temperature ALD is an urgent problem to be solved [51,52,53,54].
Flexible perovskites usually need to be encapsulated in organic materials because organic packaging materials can be synthesized through organic molecules with specific properties by changing the energy level, molecular weight and solubility of their own molecules [55]. For instance, Monojit Bag et al. demonstrated that a cross-linked self-healing polymer network of Polyisobutene (PIB) could be used as a raw material for a fully printable flexible solar cell as a self-healing sealant to protect perovskite cell components. These sealants could be applied to flexible perovskite solar modules by drip casting, rotary coating, or blade coating. The team indicated that PIB based devices with cross-linked sealants enjoyed good self-healing property and stability [56]. It can be seen that self-healing polymers can be directly used as packaging materials to achieve good self-healing ability, thus improving the durability of PSCs and preventing lead leakage [57].

4. Chemical Adsorption

While packaging methods can greatly reduce lead leakage from devices into the environment, there still be some lead leakage during extreme weather when solar panels can be severely damaged. The use of packaging technology often entails increased costs. Therefore, finding other methods for curbing lead leakage is of urgent necessity. In this case, chemisorption is an economical and effective method, given its high binding energy with lead ions in a mesoporous structure. Cation exchange resin (CER) with a high adsorption capacity and adsorption rate for lead has aroused wide interest from researchers. For illustration, Shangshang Chen et al. further investigated CER by adding sulfonic acid groups to CER materials to make Pb2+ adsorbents. Because the sulfonic acid group had a strong bond for Pb2+, Pb2+ could be firmly adsorbed on the surface of CER to prevent the outflow of Pb2+ (shown in Figure 5a). The mesoporous CER layer consists of a number of nanoparticles with an average size of 50 nm that form a dense layer of CER on top of the copper electrode (as shown in Figure 5c–e). This mesoporous structure can significantly increase the surface area to adsorb more lead. The team tested the lead adsorption rate of CER films rinsed with lead-contaminated running water by preparing the CER films on a glass sheet based on ultrasonic-assisted suspension of CER powder in isopropyl alcohol, and then the CER precursor solution was coated on a glass substrate with a thickness ranging from 300 to 1300 nm, it was found that the 300 nm thick CER layer coated on the glass immediately reduced the lead content of the running water by 30% and was independent of the initial lead concentration in the solution. The concentration of Pb2+ in water did not decrease significantly with the increase of thickness, indicating that the diffusion rate of Pb2+ to the CER surface was limited (Figure 5b). The water contact angle of ~13° was so small that water could easily penetrate into the porous structure of CER, thus promoting the adsorption of Pb2+. The PSCs with a CER layer maintained the lead absorption capacity after 600 h of irradiation in the experiment of strong ultraviolet radiation simulation, showing good photostability. In the drip test, lead leakage was reduced by 98% to only 14.3 ppb (Figure 5f,g) [58]. This low lead leakage indicated that CER adsorbed a large amount of Pb2+.
Various defects in perovskite films can form deep energy wells that affect carrier transport and device performance. The defect surface has large surface energy and decomposition energy, which can lead to spontaneous relaxation of perovskite surface and leakage of lead ions [59]. As the recombination center, unsaturated Pb can cooperate with a mercaptan group to reduce recombination. Qingrui Wang et al. used 1,2-ethanol instead of mercaptan (1,2-EDT) for surface treatment of perovskite films to form Pb-S bonds. The modification could strengthen the Pb-I bond and passivate the Pb-suspended bond, and the perovskite film delayed the dissolution of lead iodide into the water, thus inhibiting the leakage of Pb. Compared with the conventional film (49.19), the treated perovskite film had a larger water contact angle (71.83), indicating better hydrophobicity and moisture resistance, which could prevent the decomposition of perovskite caused by water entering the devices. The treated perovskite film postponed the dissolution of lead iodide into the water, ultimately reducing lead leakage [60].
The chelating resin features good selectivity, pre-concentration factor, binding energy, and mechanical stability for the removal of heavy ions. It can smoothly carry out the regeneration of multiple adsorption-desorption cycles and enjoys good reproducibility in terms of adsorption characteristics [61]. XAD resins have good physical properties, such as porosity, uniform high surface area, pore size distribution, and chemically uniform non-ionic structure. They are used as a carrier for fixed chelating agents to remove metals. V Tharanitharan et al. prepared a new modified Amber lead XAD-7HP resin using harmless sodium dioctyl succinate (DOSSS) and EDTA-disodium salt (chelating agent). After 7% sodium chloride treatment, the modified XAD-7HP resin could successfully recover lead (II) ions, and the removal rate of Pb (II) could reach 99% at the adsorbent dosage of 0.9 g/100 mL [62].
Xun Li et al. deposited a transparent Pb-absorbing P, P′-di(2-ethylhexyl) methanediphosphonic acid (DMDP) film on the glass. DMDP coatings with a thickness ranging from 0.7 to 6.89 μm were highly transparent and had good light transmit ability, which would not have adverse effects on the efficiency of the device [63]. The two phosphate groups in each DMDP molecule could be strongly bound to a Pb2+ ion, and when soaked in water, the suction plates on either side expanded to absorb lead rather than dissolve it, thus maintaining structural integrity for easy collection of lead after damage. Li Xun et al. proposed a new method for suppressing lead leakage using a standard solar vinyl acetate (EVA) film and pre-laminated P, P′-di(2-ethylhexyl) methanediphosphonic acid (DMDP) to form a thin layer similar to tape. This layer could be attached to the sides of perovskite solar cells. Whether the solar cell was an n-i-p structure or a p-i-n structure, it could be tightly bonded with the glass surface and became transparent, so it did not affect the transmittance and exerted no adverse effect on the normal photovoltaic performance. The tape could be integrated into the packaging material of the unit at a later stage, so there was no strong dependence on the PSC manufacturing process. The process of making DMDP absorption layer is shown in Figure 6. Our research group’s experiments displayed that the presence of lead absorption bands did not reduce the efficiency and stability of the PV module. When PV module was damaged due to extreme weather, the tape absorbed large amounts of leaked lead and maintained SQE above 99.9% for 7 days, demonstrating excellent lead leak suppression properties [64].
Chen Junjun modified tin dioxide (tin oxide) layer with sodium phosphate (Na3PO4), which provided a new way to solve the problem of lead leakage (Figure 7a). PO43+ groups can chelate with dissolved Pb ions to form water-insoluble compounds of Pb2PO4I, thus reducing Pb leakage. From the XRD patterns in Figure 7b, we found that SnO2-based perovskite film was decomposed into pure PbI2 after soaking in water, but no lead iodide diffraction peak was detected in the XRD patterns of SnO2: Na3PO4 sample, which revealing that PO43+ groups could chelate with dissolved Pb2+. In addition, when dipped the samples into water, the Pb2+concentration of Na3PO4 containing film slightly increased to 0.2 ppm in 30 min, while that of the control gradually approached the maximum of 1 ppm in 30 min (Figure 7c). Furthermore, Na3PO4 containing layer not only improved the performance of the device, but also captured most of the dissolved lead in the water. The inclusion of phosphate facilitated charge transfer and passivated the buried perovskite interface, resulting in a substantial increase in device efficiency of up to 23% with negligible hysteresis. More importantly, the phosphorylated tin oxide layer had a high lead adsorption capacity. Due to the numerous anchor points of oxygen solitary pairs, the isolation efficiency reached 79.6%, which could convert dissolved lead into insoluble compounds in water, the concentration of Pb2+ decreased to 2 ppm after 20 min (Figure 7d) [65].
It is worth noting that both encapsulation and chemisorption have advantages and disadvantages (shown in Table 1). The encapsulation layer can effectively inhibit the water vapor and oxygen into the battery, prevent the leakage of lead, and thus significantly enhance the stability of the device. Compared to the encapsulated technology, chemisorption has a more obvious result. As the perovskite layer decomposes, the chemisorption layer can absorb lead and/or form chelates in the first place to avoid lead leakage. However, when the chemisorption layer absorbs lead in large quantities or forms chelates, the perovskite lattice may collapse due to the reduction of lead, eventually result in the PCE dropped [66]. Therefore, more efforts should be made to design a new chemisorption layer which can suppress the lead leakage and protect the perovskite lattice as well. In addition, the encapsulation layer and the chemisorption layer can be used together, which would realize a dual lock system.

5. Lead-Free Perovskite Solar Cells

Although the above methods can effectively reduce lead leakage, there exists a small risk of lead leakage. In order to completely avoid the risk of lead leakage contaminating the environment, the most effective method is to find ions that can retain the unique photoelectric properties of lead halide perovskite and have low or even non-toxic materials to replace lead. In recent years, researchers have become interested in lead-free perovskite solar cells, which, although not as efficient as lead-based perovskite solar cells at this stage, show promising prospects.
Because lead halide perovskite material has excellent photoelectric performance, it is not only difficult to find substitutes for lead with low toxicity and easy processing properties, but also it requires direct band gap absorption of strong light and photon cycling to show excellent photovoltaic properties [67]. In fact, the promising alternatives to Pb are Sn/Ge halides, some double perovskites, and some Bi/Sb halides with perovskite-like structures (Figure 8). Specifically speaking, tin is the most promising element to replace lead. Tin halide perovskite has similar crystal structure to lead perovskite with ABX3 lattice [68,69]. Sn-based perovskite has a narrower optical band gap and a higher carrier mobility, so it should theoretically have better optical properties [69,70]. Tin-based perovskites are represented by methylammonium tin oxide (MASnI3), formamidine tin iodide (FASnI3) and cesium tin iodide (CsSnI3), which have direct band gaps of about 1.20, 1.41, and 1.3 eV, and are narrower than that of lead-based perovskites [71]. Sn has outer lone pair electrons. Theoretically, all Pb in MAPbI3 can be replaced by Sn to form MASnI3, which has a smaller band gap and a larger absorption coefficient than MAPbI3. In theory, it should have better photovoltaic performance. All-halide MASnX3 perovskite films, which can be processed from solution and exhibit good crystal quality, are expected to compete with existing photovoltaic technologies [72,73]. Unfortunately, compared with the heavier Pb elements in the 14th group of the periodic table, Sn elements with an electronic structure of ns2np2 have weaker interactions and are easily oxidized from Sn2+ to Sn4+ [68,74]. Therefore, tin-based perovskite solar cells are usually treated in glove boxes to prevent rapid degradation in the air. The researchers found that this effect could be suppressed by adding a reducing agent. For example, Hoshi et al. introduced HOOC(CH2) 4NH3I (5-AVAI) into MASnI3 to alleviate the oxidation of Sn2+, which significantly improved the stability of the device [75]. FASnI3 is another tin-based perovskite with excellent properties. The cationic radius of FA is slightly larger than that of MA. Replacing MA with FA can effectively reduce the oxidation degree from Sn2+ to Sn4+, making FASnI3 more stable than MASnI3 at room temperature [76]. The crystallization rate of FASnI3 is much faster, resulting in abundant trap states and lower open circuit voltage (VOC). To resolve this problem, Meng Xiang Yue et al. introduced hydrogen bonds into FASnI3 by adding polyvinyl alcohol (PVA) [77]. These hydrogen bonds introduced nucleation sites, which could slow the growth of crystals, guide crystal orientation, and reduce trap states. The inhibition of iodide migration greatly improved the stability of FASnI3, with the conversion efficiency reaching 8.9%. CsSnX3 is an all-inorganic Sn-based perovskite, where X can be I or Br. The black phase of CsSnX3 is a direct bandgap semiconductor, with a hole mobility of µh ≈ 585 cm2 V−1 s−1 and a carrier concentration of ≈1017 cm−3, showing good photoelectric performance [78]. Maning Liu et al. found that Ge2+ could not only stabilize Sn2+ cations, but also enhance their optical and physical properties. They effectively filled the high-density ground Sn vacancy, reduced surface defects and increased photoluminescence quantum yield by partially replacing Sn atoms in nanostructures with Ge atoms [79]. As an all-inorganic perovskite, CsSnX3 has better thermal stability than traditional organic-inorganic perovskite. As a less toxic element, Sn2+ degrades to environmentally sound tin oxide when exposed to air, which is more beneficial to the environment than lead-based perovskite solar cells [80].
Germanium, which belongs to the same main group as lead, is another element of interest to scientists. However, considering the small ionic radius of Germanium, poor solubility in polar solvents, and relatively wide band gap (1.6 eV), the PSC development rate of Germanium is much lower than that of tin PSC, Germanium PCE accounts for less than 5% [80,81,82].
Lead-free double perovskite has attracted much attention as a promising environmentally friendly photovoltaic material due to its inherent thermodynamic stability, appropriate band gap, small carrier effective mass, and low exciton binding energy [83]. An innovative strategy is to replace two Pb2+ ions with a univalent cation and a trivalent cation to form a chartable double perovskite in the shape of A2BIBIIIX6 [84]. Cuncun Wu et al. successfully produced high-quality, highly stable double perovskite Cs2AgBiBr6 thin films using the low-pressure assisted (LPA) method (Figure 9a,b), and employed the films to produce planar heterojunction solar cells with an efficiency of 1.44% [85]. Scanning electron microscopy (SEM) images and film photographs (Figure 9c,d) showed that the LPA film presented a dense and smooth state. It was exciting that Cs2AgBiBr6 films exhibited good moisture, photostability, and thermal stability. The crystals lasted 240 days in the ambient atmosphere, showing no signs of decomposition, and even the chemical bonds remained the same. Besides, double perovskites composed of Cs2AgInCl6 are usually doped with various elements and have attracted attention for their superior optical properties, namely self-trapping exciton (STE) emission and dopant induced photoluminescence. By alloying or doping in Cs2AgInCl6, it is possible to break the dual forbidden transition, change the band gap, and ultimately enhance the optical emission characteristics. Zhiguo Xia et al. studied Cs2AgInCl6 and proved that Cs2AgInCl6 had a direct band gap and a long carrier life and could be easily handled. The team demonstrated that Cs2AgInCl6 perovskite was thermodynamically stable and had several advantages over organic-inorganic perovskite [83].
Lead-free inorganic copper-silver-bismuth halide materials may become a new development direction in the field of lead-free perovskite because of their environmental friendliness, high element abundance and low cost. Recently, Erchuang Fan et al. prepared inorganic lead-free CuaAgm1Bim2In absorption layers with a direct band gap of 1.78 eV by using low temperature gas-solid phase dispersion induced direct metal Surface element reaction (DMSER). At the same time, CuaAgm1Bim2In/CuI bilayer films were prepared by one step low temperature gas-solid phase dispersion induced elemental reaction. The team through the FTO/TiO2/CuaAgm1Bim2In/CuI/carbon structure solar cells, and obtained the PCE of 2.76%, for high efficiency, environmental protection provides a possible way to photovoltaic field [86].

6. Conclusions and Outlook

It remains a problem as to how to prevent the leakage of lead and effectively reduce the harm to the environment in order to achieve the wide commercial application of PSCs. Although the content of lead in PSCs is highly small, it is far from negligible in the current context. In general, the main methods to solve lead leakage at this stage are packaging, chemical adsorption, and lead-free PSCs. Firstly, packaging is the most common method to protect devices, which can greatly improve the strength of devices, powerfully resist the impact of the external environment, and effectively prevent lead leakage and improve the stability of devices. However, an important issue faced by packaging method is the aging of devices. Antioxidants need to be added to improve the anti-aging and stability of packaging materials. At the same time, if the PV module is damaged and cannot be recovered for a long time, some lead will leak into the environment. Secondly, compared with the encapsulation method, chemical adsorption has such characteristics as low cost and high absorption efficiency. Chemisorption reduces the amount of lead leakage by 99.9% in certain cases through a series of chemical reactions during the lead leakage process to adsorb the lead in the device, or to adsorb the lead in the water. In the meantime, chemisorption can also be combined with physical packaging to achieve the lowest amount of lead leakage and reduce the pollution to the water source and the environment. Though chemisorption method has a good application prospect, further research is in need to achieve its commercial application. Thirdly, lead-free PSCs has developed rapidly in the past few years. At present, tin-based perovskite solar cells are the most competitive products to replace lead-based perovskite. Meanwhile, Germanium-based perovskite and double perovskite are developing stably. The biggest challenges confronting these lead-free perovskite products are the issues of product stability, conversion efficiency, and cost, which should be further explored by researchers. If these problems are solved, lead-free perovskite solar cells will be commercialized as low-toxicity or even non-toxic perovskite solar cells. We believe that PSCs can be made safer and more reliable through a combination of physical packaging and chemisorption, and by finding a strategy to recover lead from the environment. The search should continue for new lead-free, non-toxic materials to address concerns about the safety of lead.

Author Contributions

K.M. contributed to methodology, investigation, writing—original draft preparation; X.L. contributed to validation, data curation, formal analysis; F.Y. contributed to data curation, formal analysis, writing—review and editing; H.L. contributed to writing—review and editing, supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Henan Province college youth backbone teacher project (No.2020GGJS062).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare that they had no known competing financial interests or personal relationships that could had appeared to influence the work reported in this paper.

References

  1. Seigo, I. Research Update: Overview of progress about efficiency and stability on perovskite solar cells. APL Mater. 2016, 4, 91504. [Google Scholar]
  2. Krishna, B.G.; Sundar Ghosh, D.; Tiwari, S. Progress in ambient air-processed perovskite solar cells: Insights into processing techniques and stability assessment. Sol. Energy 2021, 224, 1369–1395. [Google Scholar] [CrossRef]
  3. Elumalai, N.; Mahmud, M.; Wang, D.; Uddin, A. Perovskite Solar Cells: Progress and Advancements. Energies 2016, 9, 861. [Google Scholar] [CrossRef]
  4. Wu, T.; Qin, Z.; Wang, Y.; Wu, Y.; Chen, W.; Zhang, S.; Cai, M.; Dai, S.; Zhang, J.; Liu, J.; et al. The Main Progress of Perovskite Solar Cells in 2020-2021. Nano Micro Lett. 2021, 13, 152. [Google Scholar] [CrossRef]
  5. Park, N.-G. Perovskite solar cells: An emerging photovoltaic technology. Mater. Today 2015, 18, 65–72. [Google Scholar] [CrossRef]
  6. Shai, X.-X.; Li, D.; Liu, S.-S.; Li, H.; Wang, M.-K. Advances and Developments in Perovskite Materials for Solar Cell Applications. Acta Phys. Chim. Sin. 2016, 32, 2159–2170. [Google Scholar] [CrossRef]
  7. Gao, Y.; Hu, Y.; Yao, C.; Zhang, S. Recent Advances in Lead-Safe Perovskite Solar Cells. Adv. Funct. Mater. 2022, 32, 2208225. [Google Scholar] [CrossRef]
  8. Fang, Z.; Zeng, Q.; Zuo, C.; Zhang, L.; Xiao, H.; Cheng, M.; Hao, F.; Bao, Q.; Zhang, L.; Yuan, Y.; et al. Perovskite-based tandem solar cells. Sci. Bull. 2021, 66, 621–636. [Google Scholar] [CrossRef]
  9. Qarony, W.; Hossain, M.I.; Salleo, A.; Knipp, D.; Tsang, Y.H. Rough versus planar interfaces: How to maximize the short circuit current of perovskite single and tandem solar cells. Mater. Today Energy 2019, 11, 106–113. [Google Scholar] [CrossRef]
  10. Torabi, N.; Behjat, A.; Zhou, Y.; Docampo, P.; Stoddard, R.J.; Hillhouse, H.W.; Ameri, T. Progress and challenges in perovskite photovoltaics from single- to multi-junction cells. Mater. Today Energy 2019, 12, 70–94. [Google Scholar] [CrossRef]
  11. Jin, X.; Yang, Y.; Zhao, T.; Wu, X.; Liu, B.; Han, M.; Chen, W.; Chen, T.; Hu, J.-S.; Jiang, Y. Mitigating Potential Lead Leakage Risk of Perovskite Solar Cells by Device Architecture Engineering from Exterior to Interior. ACS Energy Lett. 2022, 7, 3618–3636. [Google Scholar] [CrossRef]
  12. Assadi, M.K.; Bakhoda, S.; Saidur, R.; Hanaei, H. Recent progress in perovskite solar cells. Renew. Sustain. Energy Rev. 2018, 81, 2812–2822. [Google Scholar] [CrossRef]
  13. Heo, J.H.; Im, S.H. CH3 NH3 PbBr3 -CH3 NH3 PbI3 Perovskite-Perovskite Tandem Solar Cells with Exceeding 2.2 V Open Circuit Voltage. Adv. Mater. 2016, 28, 5121–5125. [Google Scholar] [CrossRef]
  14. Basumatary, P.; Agarwal, P. A short review on progress in perovskite solar cells. Mater. Res. Bull. 2022, 149, 111700. [Google Scholar] [CrossRef]
  15. Poespawati, N.R.; Sulistianto, J.; Abuzairi, T.; Purnamaningsih, R.W.; Mehta, V. Performance and Stability Comparison of Low-Cost Mixed Halide Perovskite Solar Cells: CH3NH3PbI3-xClx and CH3NH3PbI3-xSCNx. Int. J. Photoenergy 2020, 2020, 8827917. [Google Scholar] [CrossRef]
  16. Hong, Q.-M.; Xu, R.-P.; Jin, T.-Y.; Tang, J.-X.; Li, Y.-Q. Unraveling the light-induced degradation mechanism of CH3NH3PbI3 perovskite films. Org. Electron. 2019, 67, 19–25. [Google Scholar] [CrossRef]
  17. 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. [Google Scholar] [CrossRef]
  18. Kumar, A.; Kumar, A.; Cabral-Pinto, M.M.S.; Chaturvedi, A.K.; Shabnam, A.A.; Subrahmanyam, G.; Mondal, R.; Gupta, D.K.; Malyan, S.K.; Kumar, S.S.; et al. Lead Toxicity: Health Hazards, Influence on Food Chain, and Sustainable Remediation Approaches. Int. J. Environ. Res. Public Health 2020, 17, 2179. [Google Scholar] [CrossRef]
  19. Ravi, V.K.; Mondal, B.; Nawale, V.V.; Nag, A. Don’t Let the Lead Out: New Material Chemistry Approaches for Sustainable Lead Halide Perovskite Solar Cells. ACS Omega 2020, 5, 29631–29641. [Google Scholar] [CrossRef]
  20. Wan, J.; Yu, X.; Zou, J.; Li, K.; Chen, L.; Peng, Y.; Cheng, Y.-B. Lead contamination analysis of perovskite modules under simulated working conditions. Sol. Energy 2021, 226, 85–91. [Google Scholar] [CrossRef]
  21. Luo, H.; Li, P.; Ma, J.; Han, L.; Zhang, Y.; Song, Y. Sustainable Pb Management in Perovskite Solar Cells toward Eco-Friendly Development. Adv. Energy Mater. 2022, 12, 2201242. [Google Scholar] [CrossRef]
  22. Babayigit, A.; Duy Thanh, D.; Ethirajan, A.; Manca, J.; Muller, M.; Boyen, H.G.; Conings, B. Assessing the toxicity of Pb- and Sn-based perovskite solar cells in model organism Danio rerio. Sci. Rep. 2016, 6, 18721. [Google Scholar] [CrossRef]
  23. Banna, H.U.; Anjum, A.; Biswas, S.; Mondal, V.; Siddique, A.E.; Roy, A.K.; Nikkon, F.; Haque, A.; Himeno, S.; Salam, K.A.; et al. Parental Lead Exposure Promotes Neurobehavioral Disorders and Hepatic Dysfunction in Mouse Offspring. Biol. Trace Elem. Res. 2022, 200, 1171–1180. [Google Scholar] [CrossRef] [PubMed]
  24. Lee, S.W.; Bae, S.; Kim, D.; Lee, H.S. Historical Analysis of High-Efficiency, Large-Area Solar Cells: Toward Upscaling of Perovskite Solar Cells. Adv. Mater. 2020, 32, e2002202. [Google Scholar] [CrossRef] [PubMed]
  25. Ono, L.K.; Qi, Y.; Liu, S. Progress toward Stable Lead Halide Perovskite Solar Cells. Joule 2018, 2, 1961–1990. [Google Scholar] [CrossRef]
  26. Aslan, E.; Ates Turkmen, T.; Alturk, E. High stability perovskite solar cells under ambient conditions. IET Renew. Power Gener. 2020, 14, 3160–3163. [Google Scholar] [CrossRef]
  27. Wang, K.; Huo, J.; Cao, L.; Yang, P.; Muller-Buschbaum, P.; Tong, Y.; Wang, H. Fully Methylammonium-Free Stable Formamidinium Lead Iodide Perovskite Solar Cells Processed under Humid Air Conditions. ACS Appl. Mater. Interfaces 2023, 15, 13353–13362. [Google Scholar] [CrossRef]
  28. Ma, L.; Guo, D.; Li, M.; Wang, C.; Zhou, Z.; Zhao, X.; Zhang, F.; Ao, Z.; Nie, Z. Temperature-Dependent Thermal Decomposition Pathway of Organic–Inorganic Halide Perovskite Materials. Chem. Mater. 2019, 31, 8515–8522. [Google Scholar] [CrossRef]
  29. Hailegnaw, B.; Kirmayer, S.; Edri, E.; Hodes, G.; Cahen, D. Rain on Methylammonium Lead Iodide Based Perovskites: Possible Environmental Effects of Perovskite Solar Cells. J. Phys. Chem. Lett. 2015, 6, 1543–1547. [Google Scholar] [CrossRef]
  30. Burwig, T.; Heinze, K.; Pistor, P. Thermal decomposition kinetics of FAPbI3 thin films. Phys. Rev. Mater. 2022, 6, 065404. [Google Scholar] [CrossRef]
  31. Conings, B.; Babayigit, A.; Boyen, H.-G. Fire Safety of Lead Halide Perovskite Photovoltaics. ACS Energy Lett. 2019, 4, 873–878. [Google Scholar] [CrossRef]
  32. Boyd, C.C.; Cheacharoen, R.; Leijtens, T.; McGehee, M.D. Understanding Degradation Mechanisms and Improving Stability of Perovskite Photovoltaics. Chem. Rev. 2019, 119, 3418–3451. [Google Scholar] [CrossRef] [PubMed]
  33. Xudong, G.; Guangda, N.; Liduo, W. Chemical Stability Issue and Its Research Process of Perovskite Solar Cells with High Efficiency. Chin. J. Chem. 2015, 73, 211. [Google Scholar]
  34. Goetz, K.P.; Taylor, A.D.; Hofstetter, Y.J.; Vaynzof, Y. Sustainability in Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2021, 13, 1–17. [Google Scholar] [CrossRef] [PubMed]
  35. Cheng, Y.; Yang, Q.-D.; Ding, L. Encapsulation for perovskite solar cells. Sci. Bull. 2021, 66, 100–102. [Google Scholar] [CrossRef] [PubMed]
  36. Seethamraju, S.; Ramamurthy, P.C.; Madras, G. Encapsulation for Improving the Efficiencies of Solar Cells. Mater. Process. Sol. Fuel Prod. 2014, 174, 23–40. [Google Scholar]
  37. Jing, J.; Chen, S.; Zhang, J. Investigation of UV aging influences on the crystallization of ethylene-vinyl acetate copolymer via successive self-nucleation and annealing treatment. J. Polym. Res. 2010, 17, 827–836. [Google Scholar]
  38. El Amrani, A.; Mahrane, A.; Moussa, F.Y.; Boukennous, Y. Solar Module Fabrication. Int. J. Photoenergy 2007, 2007, 027610. [Google Scholar] [CrossRef]
  39. Jiang, S.; Wang, K.; Zhang, H.; Ding, Y.; Yu, Q. Encapsulation of PV Modules Using Ethylene Vinyl Acetate Copolymer as the Encapsulant. Macromol. React. Eng. 2015, 9, 522–529. [Google Scholar] [CrossRef]
  40. Tucker, R.T.; Yorgensen, R.S. A review of field performance of EVA-based encapsulants. In Proceedings of the IEEE Photovoltaic Specialists Conference, New Orleans, LA, USA, 19–24 May 2002. [Google Scholar]
  41. Kim, H.C.; Jang, T.W.; Chae, H.J.; Choi, W.J.; Ha, M.N.; Ye, B.J.; Kim, B.G.; Jeon, M.J.; Kim, S.Y.; Hong, Y.S. Evaluation and management of lead exposure. Ann. Occup. Environ. Med. 2015, 27, 30. [Google Scholar] [CrossRef]
  42. Cheacharoen, R.; Boyd, C.C.; Burkhard, G.F.; Leijtens, T.; Raiford, J.A.; Bush, K.A.; Bent, S.F.; McGehee, M.D. Encapsulating perovskite solar cells to withstand damp heat and thermal cycling. Sustain. Energy Fuels 2018, 2, 2398–2406. [Google Scholar] [CrossRef]
  43. Fu, Z.; Xu, M.; Sheng, Y.; Yan, Z.; Meng, J.; Tong, C.; Li, D.; Wan, Z.; Ming, Y.; Mei, A. Encapsulation of Printable Mesoscopic Perovskite Solar Cells Enables High Temperature and Long-Term Outdoor Stability. Adv. Funct. Mater. 2019, 29, 1809129.1809121–1809129.1809127. [Google Scholar] [CrossRef]
  44. Kim, H.; Lee, J.; Kim, B.; Byun, H.R.; Kim, S.H.; Oh, H.M.; Baik, S.; Jeong, M.S. Enhanced Stability of MAPbI(3) Perovskite Solar Cells using Poly(p-chloro-xylylene) Encapsulation. Sci. Rep. 2019, 9, 15461. [Google Scholar] [CrossRef]
  45. Shi, L.; Bucknall, M.P.; Young, T.L.; Zhang, M.; Hu, L.; Bing, J.; Lee, D.S.; Kim, J.; Wu, T.; Takamure, N.; et al. Gas chromatography-mass spectrometry analyses of encapsulated stable perovskite solar cells. Science 2020, 368, eaba2412. [Google Scholar] [CrossRef] [PubMed]
  46. Yin, J.; Cui, J.; Zhou, H.; Cui, S.; Wang, C.; Guo, J.; Wei, J.; Zhang, X. Encapsulation of UV Glue, Hydrophobicity of Binder and Carbon Electrode Enhance the Stability of Organic–Inorganic Hybrid Perovskite Solar Cells up to 5 Years. Energy Technol. 2020, 8, 2000513. [Google Scholar] [CrossRef]
  47. Liu, W.; Wang, Z.; Chen, Z.; Zhao, L. Thermo-initiated cationic polymerization of phosphorus-containing cycloaliphatic epoxides with tunable degradable temperature. Polym. Degrad. Stab. 2012, 97, 810–815. [Google Scholar] [CrossRef]
  48. Jiang, Y.; Qiu, L.; Juarez-Perez, E.J.; Ono, L.K.; Hu, Z.; Liu, Z.; Wu, Z.; Meng, L.; Wang, Q.; Qi, Y. Reduction of lead leakage from damaged lead halide perovskite solar modules using self-healing polymer-based encapsulation. Nat. Energy 2019, 4, 585–593. [Google Scholar] [CrossRef]
  49. Wang, X.; Dong, B.; Feng, M.; Xue, D.-J.; Wang, S.-M. Sustainable management of lead in perovskite solar cells. J. Mater. Chem. A 2022, 10, 15861–15864. [Google Scholar] [CrossRef]
  50. Lee, Y.I.; Jeon, N.J.; Kim, B.J.; Shim, H.; Yang, T.-Y.; Seok, S.I.; Seo, J.; Im, S.G. A Low-Temperature Thin-Film Encapsulation for Enhanced Stability of a Highly Efficient Perovskite Solar Cell. Adv. Energy Mater. 2018, 8, 1701928. [Google Scholar] [CrossRef]
  51. Sarkar, S.; Culp, J.H.; Whyland, J.T.; Garvan, M.; Misra, V. Encapsulation of organic solar cells with ultrathin barrier layers deposited by ozone-based atomic layer deposition. Org. Electron. 2010, 11, 1896–1900. [Google Scholar] [CrossRef]
  52. Park, J.S.; Chae, H.; Chung, H.K.; Lee, S.I. Thin film encapsulation for flexible AM-OLED: A review. Semicond. Sci. Technol. 2011, 26, 034001. [Google Scholar] [CrossRef]
  53. Zhang, Y.; Yang, Y.; Mbumba, M.T.; Akram, M.W.; Rop, E.K.; Bai, L.; Guli, M. Research Progress of Buffer Layer and Encapsulation Layer Prepared by Atomic Layer Deposition to Improve the Stability of Perovskite Solar Cells. Sol. RRL 2022, 6, 2200823. [Google Scholar] [CrossRef]
  54. Clark, M.D.; Jespersen, M.L.; Patel, R.J.; Leever, B.J. Ultra-thin alumina layer encapsulation of bulk heterojunction organic photovoltaics for enhanced device lifetime. Org. Electron. 2014, 15, 1–8. [Google Scholar] [CrossRef]
  55. Zuo, C.; Bolink, H.J.; Han, H.; Huang, J.; Cahen, D.; Ding, L. Advances in Perovskite Solar Cells. Adv. Sci. 2016, 3, 1500324. [Google Scholar] [CrossRef]
  56. Bag, M.; Banerjee, S.; Faust, R.; Venkataraman, D. Self-healing polymer sealant for encapsulating flexible solar cells. Sol. Energy Mater. Sol. Cells 2016, 145, 418–422. [Google Scholar] [CrossRef]
  57. Yu, Y.; Zhang, F.; Yu, H. Self-healing perovskite solar cells. Sol. Energy 2020, 209, 408–414. [Google Scholar] [CrossRef]
  58. Chen, S.; Deng, Y.; Gu, H.; Xu, S.; Wang, S.; Yu, Z.; Blum, V.; Huang, J. Trapping lead in perovskite solar modules with abundant and low-cost cation-exchange resins. Nat. Energy 2020, 5, 1003–1011. [Google Scholar] [CrossRef]
  59. Shao, Y.; Xiao, Z.; Bi, C.; Yuan, Y.; Huang, J. Origin and elimination of photocurrent hysteresis by fullerene passivation in CH3NH3PbI3 planar heterojunction solar cells. Nat. Commun. 2014, 5, 5784. [Google Scholar] [CrossRef]
  60. Wang, Q.; Lin, Z.; Su, J.; Xu, Y.; Guo, X.; Li, Y.; Zhang, M.; Zhang, J.; Chang, J.; Hao, Y. Dithiol surface treatment towards improved charge transfer dynamic and reduced lead leakage in lead halide perovskite solar cells. EcoMat 2022, 4, e12185. [Google Scholar] [CrossRef]
  61. Tewari, P.K.; Singh, A.K. Preconcentration of lead with Amberlite XAD-2 and Amberlite XAD-7 based chelating resins for its determination by flame atomic absorption spectrometry. Talanta 2002, 56, 735–744. [Google Scholar] [CrossRef]
  62. Tharanitharan, V.; Srinivasan, K. Removal of Pb(II) from aqueous solutions by using dioctyl sodium sulphosuccinate-EDTA modified Amberlite XAD-7HP resin. Indian J. Chem. Technol. 2009, 16, 417–425. [Google Scholar]
  63. Li, X.; Zhang, F.; He, H.; Berry, J.J.; Zhu, K.; Xu, T. On-device lead sequestration for perovskite solar cells. Nature 2020, 578, 555–558. [Google Scholar] [CrossRef] [PubMed]
  64. Li, X.; Zhang, F.; Wang, J.; Tong, J.; Xu, T.; Zhu, K. On-device lead-absorbing tapes for sustainable perovskite solar cells. Nat. Sustain. 2021, 4, 1038–1041. [Google Scholar] [CrossRef]
  65. Chen, J.; Li, S.; Ma, T.; Wu, D.; Zhao, Y.; Wang, C.; Zhao, D.; Li, X. Managing Lead Leakage in Efficient Perovskite Solar Cells with Phosphate Interlayers. Adv. Mater. Interfaces 2022, 9, 2200570. [Google Scholar] [CrossRef]
  66. Dou, J.; Bai, Y.; Chen, Q. Challenges of lead leakage in perovskite solar cells. Mater. Chem. Front. 2022, 6, 2779–2789. [Google Scholar] [CrossRef]
  67. Giustino, F.; Snaith, H.J. Toward Lead-Free Perovskite Solar Cells. ACS Energy Lett. 2016, 1, 1233–1240. [Google Scholar] [CrossRef]
  68. Wu, T.; Cui, D.; Liu, X.; Luo, X.; Su, H.; Segawa, H.; Zhang, Y.; Wang, Y.; Han, L. Additive Engineering toward High-Performance Tin Perovskite Solar Cells. Sol. RRL 2021, 5, 2100034. [Google Scholar] [CrossRef]
  69. Ke, W.; Kanatzidis, M.G. Prospects for low-toxicity lead-free perovskite solar cells. Nat. Commun. 2019, 10, 965. [Google Scholar] [CrossRef]
  70. Schileo, G.; Grancini, G. Lead or no lead? Availability, toxicity, sustainability and environmental impact of lead-free perovskite solar cells. J. Mater. Chem. C 2021, 9, 67–76. [Google Scholar] [CrossRef]
  71. Hussain, I.; Tran, H.P.; Jaksik, J.; Moore, J.; Islam, N.; Uddin, M.J. Functional materials, device architecture, and flexibility of perovskite solar cell. Emergent Mater. 2018, 1, 133–154. [Google Scholar] [CrossRef]
  72. Yang, S.; Fu, W.; Zhang, Z.; Chen, H.; Li, C.; Mater, J. Recent advances in perovskite solar cells: Efficiency, stability and lead-free perovskite. J. Mater. Chem. A 2017, 5, 11462–11482. [Google Scholar] [CrossRef]
  73. Ke, W.; Stoumpos, C.C.; Kanatzidis, M.G. “Unleaded” Perovskites: Status Quo and Future Prospects of Tin-Based Perovskite Solar Cells. Adv. Mater. 2019, 31, e1803230. [Google Scholar] [CrossRef] [PubMed]
  74. Abate, A. Stable Tin-Based Perovskite Solar Cells. ACS Energy Lett. 2023, 8, 1896–1899. [Google Scholar] [CrossRef] [PubMed]
  75. Hoshi, H.; Shigeeda, N.; Dai, T. Improved oxidation stability of tin iodide cubic perovskite treated by 5-ammonium valeric acid iodide. Mater. Lett. 2016, 183, 391–393. [Google Scholar] [CrossRef]
  76. Stoumpos, C.C.; Malliakas, C.D.; Kanatzidis, M.G. Semiconducting tin and lead iodide perovskites with organic cations: Phase transitions, high mobilities, and near-infrared photoluminescent properties. Inorg. Chem. A Res. J. That Incl. Bioinorg. Catal. Organomet. Solid State Synth. Chem. React. Dyn. 2013, 52, 9019–9038. [Google Scholar] [CrossRef]
  77. Meng, X.; Lin, J.; Liu, X.; He, X.; Wang, Y.; Noda, T.; Wu, T.; Yang, X.; Han, L. Highly Stable and Efficient FASnI(3) -Based Perovskite Solar Cells by Introducing Hydrogen Bonding. Adv. Mater. 2019, 31, e1903721. [Google Scholar] [CrossRef]
  78. Brivio, F.; Frost, J.M.; Skelton, J.M.; Jackson, A.J.; Weber, O.J.; Weller, M.T.; Goni, A.R.; Leguy, A.; Barnes, P.; Walsh, A. Lattice dynamics and vibrational spectra of the orthorhombic, tetragonal and cubic phases of methylammonium lead iodide. Phys. Rev. B 2015, 92, 144308. [Google Scholar] [CrossRef]
  79. Liu, M.; Pasanen, H.; Ali-Lytty, H.; Hiltunen, A.; Vivo, P. B-Site Co-Alloying with Germanium Improves the Efficiency and Stability of All-Inorganic Tin-Based Perovskite Nanocrystal Solar Cells. Angew. Chem. Int. Ed. 2020, 59, 22117–22125. [Google Scholar] [CrossRef]
  80. Azhari, A.W.; Then, F.S.X.; Halin, D.S.C.; Sepeai, S.; Ludin, N.A. Tin and germanium substitution in lead free perovskite solar cell: Current status and future trends. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2020. [Google Scholar]
  81. Krishnamoorthy, T.; Ding, H.; Yan, C.; Leong, W.L.; Baikie, T.; Zhang, Z.; Sherburne, M.; Li, S.; Asta, M.; Mathews, N.; et al. Lead-free germanium iodide perovskite materials for photovoltaic applications. J. Mater. Chem. A Mater. Energy Sustain. 2015, 3, 23829–23832. [Google Scholar] [CrossRef]
  82. Chiara, R.; Morana, M.; Malavasi, L. Germanium-Based Halide Perovskites: Materials, Properties, and Applications. ChemPlusChem 2021, 86, 879–888. [Google Scholar] [CrossRef]
  83. Liu, Y.; Nag, A.; Manna, L.; Xia, Z. Lead-Free Double Perovskite Cs2 AgInCl6. Angew. Chem. 2021, 60, 11592–11603. [Google Scholar] [CrossRef] [PubMed]
  84. Dave, K.; Fang, M.H.; Bao, Z.; Fu, H.T.; Liu, R.S. Recent Developments in Lead-Free Double Perovskites: Structure, Doping, and Applications. Chem. Asian J. 2020, 15, 242–252. [Google Scholar] [CrossRef] [PubMed]
  85. Wu, C.; Zhang, Q.; Liu, Y.; Luo, W.; Guo, X.; Huang, Z.; Ting, H.; Sun, W.; Zhong, X.; Wei, S.; et al. The Dawn of Lead-Free Perovskite Solar Cell: Highly Stable Double Perovskite Cs2AgBiBr6 Film. Adv. Sci. 2018, 5, 1700759. [Google Scholar] [CrossRef] [PubMed]
  86. Fan, E.; Liu, M.; Yang, K.; Jiang, S.; Li, B.; Zhao, D.; Guo, Y.; Zhang, Y.; Zhang, P.; Zuo, C.; et al. One-Step Gas-Solid-Phase Diffusion-Induced Elemental Reaction for Bandgap-Tunable Cu(a)Ag(m1)Bi(m2)I(n)/CuI Thin Film Solar Cells. Nano Micro Lett. 2023, 15, 58. [Google Scholar] [CrossRef]
Figure 1. Effects of Pb on learning and spatial memory in mice. Mice in the Control group were not exposed to lead, mice in the Pb group were exposed to lead, and mice in the P–Pb group were exposed to lead during pregnancy and after birth. Sec represents the learning and spatial memory ability of mice, and the lower the Sec value is the stronger learning and spatial memory ability of mice. Reproduced with permission from [23]. Copyright 2021, Springer Nature.
Figure 1. Effects of Pb on learning and spatial memory in mice. Mice in the Control group were not exposed to lead, mice in the Pb group were exposed to lead, and mice in the P–Pb group were exposed to lead during pregnancy and after birth. Sec represents the learning and spatial memory ability of mice, and the lower the Sec value is the stronger learning and spatial memory ability of mice. Reproduced with permission from [23]. Copyright 2021, Springer Nature.
Coatings 13 01009 g001
Figure 2. (a) Structure of EVA encapsulated solar cells. (b) Aging degradation mechanism of EVA. Reproduced with permission from [39]. Copyright 2015, John Wiley and Sons.
Figure 2. (a) Structure of EVA encapsulated solar cells. (b) Aging degradation mechanism of EVA. Reproduced with permission from [39]. Copyright 2015, John Wiley and Sons.
Coatings 13 01009 g002
Figure 3. The concentration of Pb leaking into the water. Three parallel experiments were carried out for each package method to reduce the error. (a) Perform a drip test on the damaged PV module. (b) The damaged perovskite solar module was dripped and then kept in simulated sunlight for 4 h (45 °C) before being injected again. (c) The damaged perovskite solar modules are heated at 45 °C for 4 h and then dripping with water. No matter what the conditions, group D always has the lowest lead leakage. Reproduced with permission from [48]. Copyright 2019, Springer Nature.
Figure 3. The concentration of Pb leaking into the water. Three parallel experiments were carried out for each package method to reduce the error. (a) Perform a drip test on the damaged PV module. (b) The damaged perovskite solar module was dripped and then kept in simulated sunlight for 4 h (45 °C) before being injected again. (c) The damaged perovskite solar modules are heated at 45 °C for 4 h and then dripping with water. No matter what the conditions, group D always has the lowest lead leakage. Reproduced with permission from [48]. Copyright 2019, Springer Nature.
Coatings 13 01009 g003
Figure 4. Perovskite solar modules packaged by the TFE method. (a) SEM image of cross section of TFE layer and schematic diagram of perovskite solar module packaged by TFE. (b) Changes of PCE in HTM layer after 60 °C and 90 °C ALD process. (c) Cross sectional SEM image of PSC encapsulated by CVD. The surface smoothness is improved. (d) PCE distribution of samples before and after encapsulation. Reproduced with permission from [50]. Copyright 2017, John Wiley and Sons.
Figure 4. Perovskite solar modules packaged by the TFE method. (a) SEM image of cross section of TFE layer and schematic diagram of perovskite solar module packaged by TFE. (b) Changes of PCE in HTM layer after 60 °C and 90 °C ALD process. (c) Cross sectional SEM image of PSC encapsulated by CVD. The surface smoothness is improved. (d) PCE distribution of samples before and after encapsulation. Reproduced with permission from [50]. Copyright 2017, John Wiley and Sons.
Coatings 13 01009 g004
Figure 5. (a) Schematic diagram of lead leakage prevention mechanism. (b) Different concentration of Pb2+ aqueous solution at different initial concentrations on glass with CER layers of different thicknesses of 15 cm long. (ce) SEM image of top view and cross section of MAPbI3 device with CER-coated and the top of the copper electrode with CER. (f,g) Results of immersion tests on damaged micromodules with or without CER coating. Reproduced with permission from [58]. Copyright 2020, Springer Nature.
Figure 5. (a) Schematic diagram of lead leakage prevention mechanism. (b) Different concentration of Pb2+ aqueous solution at different initial concentrations on glass with CER layers of different thicknesses of 15 cm long. (ce) SEM image of top view and cross section of MAPbI3 device with CER-coated and the top of the copper electrode with CER. (f,g) Results of immersion tests on damaged micromodules with or without CER coating. Reproduced with permission from [58]. Copyright 2020, Springer Nature.
Coatings 13 01009 g005
Figure 6. A schematic of DMDP applied to EVA film and a photo of PSC made with EVA tape. Reproduced with permission from [64]. Copyright 2021, Springer Nature.
Figure 6. A schematic of DMDP applied to EVA film and a photo of PSC made with EVA tape. Reproduced with permission from [64]. Copyright 2021, Springer Nature.
Coatings 13 01009 g006
Figure 7. (a) Schematic diagram of Pb adsorption. (b) XRD patterns. (c) Transformation of SnO2− and SnO2:Na3PO4 based perovskites in water. (d) Measurement of lead leakage and comparison of lead retention in damaged PSCs with and without Na3PO4. Reproduced with permission from [65]. Copyright 2022, John Wiley and Sons.
Figure 7. (a) Schematic diagram of Pb adsorption. (b) XRD patterns. (c) Transformation of SnO2− and SnO2:Na3PO4 based perovskites in water. (d) Measurement of lead leakage and comparison of lead retention in damaged PSCs with and without Na3PO4. Reproduced with permission from [65]. Copyright 2022, John Wiley and Sons.
Coatings 13 01009 g007
Figure 8. Potential solar cell absorption materials and band gaps of various materials. Reproduced with permission from [69]. Copyright 2019, Springer Nature.
Figure 8. Potential solar cell absorption materials and band gaps of various materials. Reproduced with permission from [69]. Copyright 2019, Springer Nature.
Coatings 13 01009 g008
Figure 9. Preparation and SEM image of Cs2AgBiBr6 thin films. (a) Image of Cs2AgBiBr6 powder (left) and DMSO solution (right). (b) Film production process diagram. (c,d) SEM images of thin films obtained by (c) TA and (d) LPA. Reproduced with permission from [85]. Copyright 2018, Advanced science.
Figure 9. Preparation and SEM image of Cs2AgBiBr6 thin films. (a) Image of Cs2AgBiBr6 powder (left) and DMSO solution (right). (b) Film production process diagram. (c,d) SEM images of thin films obtained by (c) TA and (d) LPA. Reproduced with permission from [85]. Copyright 2018, Advanced science.
Coatings 13 01009 g009
Table 1. Advantages and disadvantages of encapsulation and chemisorption.
Table 1. Advantages and disadvantages of encapsulation and chemisorption.
EncapsulationChemical Adsorption
CostHighLow
Whether mechanical strength can be improved×
Whether the amount of lead leaking into the environment can be reduced×
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

Ma, K.; Li, X.; Yang, F.; Liu, H. Lead Leakage of Pb-Based Perovskite Solar Cells. Coatings 2023, 13, 1009. https://doi.org/10.3390/coatings13061009

AMA Style

Ma K, Li X, Yang F, Liu H. Lead Leakage of Pb-Based Perovskite Solar Cells. Coatings. 2023; 13(6):1009. https://doi.org/10.3390/coatings13061009

Chicago/Turabian Style

Ma, Kai, Xiaofang Li, Feng Yang, and Hairui Liu. 2023. "Lead Leakage of Pb-Based Perovskite Solar Cells" Coatings 13, no. 6: 1009. https://doi.org/10.3390/coatings13061009

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