Skip to content
Psc solar cell. Psc solar cell

Psc solar cell. Psc solar cell

    Psc solar cell

    Perovskite, an organic–inorganic hybrid material, tends to be a promising light-harvesting material. PSCs (organic-inorganic perovskite solar cells) are considered a significant breakthrough in photovoltaics and have received great attention. Due to the inherent advantage of perovskite thin films that can be fabricated using simple solution techniques at low temperatures, PSCs are regarded as one of the most important low-cost and mass-production prospects.

    Contributors MDPI registered users’ name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :

    Introduction

    Perovskite, an organic–inorganic hybrid material, tends to be a promising light-harvesting material. In 2009, the Miyasaka group [1] reported a perovskite solar cell (PSC) of 3.8% using a DSSC system configuration with liquid electrolyte based on MAPbI3 (MA = CH3NH3 ). Park group [2] obtained a nearly doubled power conversion efficiency (PCE) of 6.5% in 2011 using a high concentration perovskite precursor solution. Since then, several efforts have been made to enhance PSC photovoltaic efficiency from various angles, including perovskite layer fabrication methods, interface engineering, cell architecture design, and development of the hole transporting materials (HTM), an electron transporting materials (ETM). A certified PCE of 22.7% has already been achieved via the above optimization [3]. Although the PCE currently available is appealing, the PSCs still have low stability (thermal, light, and moisture stability), which hinders their commercialization.

    The discovery of high-efficiency and highly stable perovskite solar cells has sparked extensive research, which is still ongoing [4] [5]. Particularly, organometallic semiconducting perovskite has a direct Band gap with high absorption coefficients [6] that enables efficient light absorption in ultra-thin films. Furthermore, it has a long diffusion length [7] [8] [9]. low exciton binding energy [10] [11]. high carrier mobility [12] [13]. and simple and easy preparation techniques [14] that help to get high efficiency and low-cost showing promising alternative to the conventional crystalline silicon-based solar cell. over, perovskite materials can be implemented in two different cell structures, either as planer (n-i-p) or inverted (p-i-n) architecture. over, both architectures could be (i) regular structures in which no mesoporous layer is employed, and (ii) mesoscopic structures where a mesoporous layer is needed. The significant improvement in efficiency already achieved in all kinds of architecture, and the stability of PSCs remain the key concerns for the researchers at present time. Many changes were made to the working electrode, the electron transport layer (ETL), and the hole transport layer (HTL) to improve their stability and charge transport properties. The hole transporting materials is a very much important factor in PSCs to achieve high efficiency and performance. It acts as the mediator to transfer positive charges (Holes) between the perovskite and counter electrode [15]. Particularly, highest efficiency (PSCs) are achieved with organic HTL such as 2,2,7,7-tetrakis-(N,N-di-pmethoxyphenylamine)-9,90-spiro-biuorene(spiro-MeOTAD) [16]. The other most commonly used organic HTMs are poly(3,4-ethylene dioxythiophene) (PEDOT) or poly(3,4-ethylene dioxythiophene):poly(styrene sulfonate) (PEDOT:PSS) [17]. poly-[[9-(1-octylnonyl)-9H-carbazole-2,7-diyl]-2,5-thiophenediyl-2,1,3-benzothiadiazole-4,7-diyl-2,5 thiophenediyl] (PCDTBT) [18] [19]. poly-[3-hexylthiophene-2,5-diyl] (P3HT) [20] [21]. 4-(diethylamino)-benzaldehyde diphenylhydrazone (DEH) [22]. poly-triarylamine (PTAA) [19] [23]. N,N-dialkyl perylene diimide (PDI) [24]. polypyrrole (PPy), polyaniline (PANI) [25]. etc. From a commercial standpoint, the production of solar cells utilizing an organic hole transport layer has encountered numerous challenges, the most significant of which are material cost and stability. Particularly, high purity spiro-OMeTAD is more costly than novel metals such as gold and platinum, which are commonly used as a counter electrodes. Commercially available spiro-OMeTAD is nearly ten times more expensive than platinum and gold. On the other hand, organic HTMs are typically hygroscopic in nature and that’s why it has an impact on the PSCs’ general stability.

    In contrast, several low-cost inorganic HTLs were also proposed and implemented for enhancing the stability of PSCs, among them, some of the HTMs are CuSCN [26]. NiOx [27]. Cu2O or CuO [28]. CuI [29]. CuGaO3 [30] and CuAlO2 [31]. MoOx [32]. CuS [33]. MoS2 [34]. and polymer electrolyte [35]. The above-mentioned HTMs have shown potential as they offer suitable properties for application in PSCs including the suitable Band-to-Band alignment with the perovskite layer, low resistivity, and low-cost solution-process ability [2]. In the case of inorganic HTM, increased demand for inorganic HTM will certainly lower the cost of large-scale manufacturing, while organic HTM will likely stay expensive due to the preparation processes and materials with very high purity required for solar cell applications. These are the primary reason why researchers have concentrated their efforts on the development of an inorganic HTM. Consecutively, the quest for the perfect HTM is a great topic yet. There is a lot of literature on various HTMs, but only a few of them show promise in terms of improving the overall efficiency and stability of the PSCs. Several approaches have evolved to utilize inorganic p-type semiconductor materials, such as NiOx, CuOx, etc., focusing on developing non-hygroscopic and highly conductive HTMs [36]. over, carbon-based materials, including graphene, activated carbon, carbon black, graphite powder, carbon nanotube (CNT), etc., have been employed in the case of HTM-free PSC structures [37] [38] [39] [40]. In particular, current approaches to HTMs are low cost, high mobility, low absorption in the visible region, ease of synthesis, and good chemical stability that could ensure high efficiency and stable PSCs. In recent years, several review works have been published on inorganic metal oxide hole-transporting materials for perovskite solar cells in different formats and among them, some are focused on fabrication way, some are on efficiency and some are focused on stability. A list of some important review articles on inorganic metal oxide-based PSCs for the year 2015–2021 is shown in Table 1.

    No. Title Journal Year References
    01 Recent progress of inorganic hole transport materials for efficient and stable perovskite solar cells Nano Select 2021 [41]
    02 A brief review of hole transporting materials commonly used in perovskite solar cells Rare Metals 2021 [42]
    03 Nickel Oxide for Perovskite Photovoltaic Cells Advanced Photonics Research 2021 [43]
    04 Toward efficient and stable operation of perovskite solar cells: Impact of sputtered metal oxide interlayers Nano Select 2021 [44]
    05 Inorganic hole transport layers in inverted perovskite solar cells: A review Nano Select 2021 [45]
    06 Progress, highlights, and perspectives on NiO in perovskite photovoltaics Chemical Science 2020 [46]
    07 A review on the classification of organic/inorganic/carbonaceous hole-transporting materials for perovskite solar cell application Arabian Journal of Chemistry 2020 [47]
    08 Review of current progress in inorganic hole-transport materials for perovskite solar cells Applied Materials Today 2019 [48]
    09 Recent progress of inorganic perovskite solar cells Energy Environmental Science 2019 [49]
    10 Inorganic hole transporting materials for stable and high efficiency perovskite solar cells The Journal of Physical Chemistry C 2018 [50]
    11 Analysing the prospects of perovskite solar cells within the purview of recent scientific advancements Crystals 2018 [51]
    12 Recent progress in stability of perovskite solar cells Journal of Semiconductors 2017 [52]
    13 Emerging of inorganic hole transporting materials for perovskite solar cells The Chemical Record 2017 [53]
    14 Recent advances in the inverted planar structure of perovskite solar cells Accounts of chemical research 2016 [54]
    15 The progress of interface design in perovskite-based solar cells Advanced Energy Materials 2016 [55]
    16 Recent progress on hole-transporting materials for emerging organometal halide perovskite solar cells Advanced Energy Materials 2015 [36]

    Perovskite Solar Cell

    PSCs (organic-inorganic perovskite solar cells) are considered a significant recent breakthrough in photovoltaics and have recently received great attention [49]. The power conversion efficiency (PCE) of PSCs has already enhanced from 3.8 percent to 25.8 percent through the system engineering and materials design regarding the correct optoelectronic aspects in just 10 years [56]. Thus, PSCs are recognized as the best alternative approach for replacing the costly and market-dominant crystalline silicon solar cells [51] [57] [58] [59] [60]. over, PSCs are more cost-effective than conventional inorganic semiconductor thin-film solar cells, such as CIGS and CdTe [52]. The real obstacle to commercialization, however, is maintaining long-term stability. PSCs are particularly susceptible to deterioration when exposed to moisture, oxygen, heat, and light, and they must address before they can use in practical applications. Perovskite is itself very reactive due to the presence of vacancies in its structure. This is the defect of perovskite and it can encourage ion migration through the perovskite layer. Furthermore, the organic cations which are used in PSCs are hygroscopic in nature. When the PSCs are contacted with moisture, the water molecule reacts with it and forms a weak hydrogen bond with the cation which results in the formation of a hydrated perovskite phase [52]. Oxygen, heat, and UV influence this chemical reaction and favor the instability of PSCs. For commercialization, PSCs must be able to operate without major degradation for almost 25 years in outdoor conditions [61]. PSCs have so far been claimed to have one-year stability, which is considerably less than the PV systems that are already on the market. Thus, it is evident that the stability and limited longevity of PSC PV are the main factors impeding its commercialization [62].

    The basic building block of the perovskite structure, ABX3, is shown in Figure 1, where A and B are cations with different sizes (A being larger than B) and X is an anion [63]. Figure 1 represents the simplest structure made up of cubic symmetry of corner-sharing BX6 octahedra, where the B cations are in the middle of the octahedron and the X anions are at the corners [64] [65]. In the gap of cuboctahedra, the A cations are located at interstices, surrounded by eight octahedral, and form a cubic Pm3m crystal structure [66]. In the case of frequently used perovskites in solar cells are organo-metal halide perovskite materials, where ‘A’ may be an organic or inorganic cation (i.e., MA. FA. Cs. K. and Rb ), while ‘B’ is a metal cation (i.e., Pb 2 or Sn 2 ), and ‘X’ is a halide anion (i.e., Cl, Br, I, etc.) [67] [68].

    Figure 1. Crystal structure of perovskite with a general chemical formula of ABX3 (in the case of CH3NH3PbI3, A represents the CH3NH3, B represents the Pb, and X represents I).

    It should be mentioned that the A, B, and X ions must satisfy this formula, t = (RA RX)/2 (RB RX), where RA, RB, and RX are the corresponding ionic radii and t = 1, is the tolerance factor. For most cubic perovskite structures, 0.8 t 0.9 is found quantitatively. In the case of lower symmetry, the value of “t” is very small and then the film structure will be tetragonal or orthorhombic. Alternatively, if t ≥ 1, hexagonal structures are formed, and layers of face-sharing octahedra are added to the structure [67] [68]. over, organometal halide perovskites have already been proven several outstanding optoelectronic properties, such as a large absorption coefficient, direct bandgap, small exciton-binding energy, ambipolar semiconducting characteristics, long charge-carrier diffusion length with high charge-carrier mobility [67] [68]. Furthermore, the researcher proposed hybrid organometal perovskite material with structure ABX3−xYx, for example, MAPbI3−xClx and MAPbI3−xBrx, which has tunable optical properties. The tunable optical properties make it easier to experiment with device performance and improve PSCs’ overall performance [68]. On the other hand, perovskite films can be prepared by versatile low cost and simple film deposition methods, such as spin-coating [69] [70]. sequential deposition [71] [72]. and evaporation [73] [74] techniques. Low-temperature spin coating is the simplest method to fabricate low-cost and high-efficiency PVSC devices. However, it is very challenging to form continuous perovskite films means non-fully covered perovskite films by spin-coating via the direct methyl ammonium halide and lead iodide (PbI2) mixed precursor solution [75] [76]. All the above process has their own limitation and commercial viability.

    Miyasaka and co-workers first reported the liquid-electrolyte-based dye-sensitized solar cells (DSCs) of PCE as a maximum of 3.8% using MAPbI3 and MAPbBr3 perovskites as light absorbers [1]. However, due to the dissolution of the perovskites in the liquid electrolyte, the system was found to be very unstable. In 2012, a significant advance was made independently by Grätzel et al. [60] and Snaith et al. [77] where the liquid electrolyte was replaced with a small-molecule-based hole-transporting material (HTM), 2,2′,7,7′-tetrakis(N,N-di-p methoxyphenylamine)-9,9′-spirobifluorene(spiro-OMeTAD). The perovskite is penetrated the mesoporous TiO2 (mp-TiO2) scaffold with an additional capping layer as shown in Figure 2a, which is covered with a thin layer of the HTM in a typical mesoscopic PSC. Finally, a metal electrode, preferably gold (Au), is deposited on the top of the HTM [61] [77]. Instead of TiO2, Al2O3 insulating scaffold can also be used in this mesoscopic structure [77]. The device has been found to work well, signifying that the perovskite could serve as a light harvester as well as an electron transporter (ETM). This finding led to a planar PSC configuration without the mesoporous scaffold as shown in Figure 2b. Particularly, in planar PSCs, the perovskite is simply sandwiched between a thin layer of HTM and a compact ETM, such as TiO2, ZnO, SnO2, etc. [78]. over, HTM-free PSC also reported where the perovskite works as a hole transporter as well as a light absorber [79]. over, ambipolar semiconducting characteristics of the perovskite support fabricating PSC in an inverted fashion, which is typically known as inverted PSCs. Figure 2c represents the mesoscopic inverted PSCs where a p-type mesoporous matrix (such as NiO) is used to deposit the perovskite, and then, a thin layer of ETM is deposited on top of the perovskite [80]. Finally, fabrication has been completed by depositing a metal electrode, such as silver (Ag), by the thermal evaporation technique. Analogous to usual architectures, the PSCs in inverted structure can be fabricated as shown in Figure 2d, where the perovskite layer is sandwiched by an ETM, such as PCBM, and a thin HTM, such as poly(3,4-ethylene dioxythiophene):poly(styrene sulfonic acid) (PEDOT:PSS) [54].

    Figure 2. Device architectures of perovskite solar cells; (a) normal mesoscopic, (b) normal planar, (c) inverted mesoscopic, and (d) inverted planar structure.

    As previously mentioned, PSCs use primarily two types of system structures (normal and inverted) and obviously, transparent conductive oxide (TCO) (such as ITO or FTO), HTM, perovskite layer, ETM, and contact electrodes (like Au and Ag) are the main components of both structures as shown in Figure 2a,b. The energy Band diagram of a normal configuration, shown schematically in Figure 3, depicts the transporting trajectory of electrons and holes during the action. Excitons are produced and then separated into free carriers when sunlight illuminates the perovskite active layer. The generated electrons and holes can then be transported to each interface and injected into ETM and HTM, respectively. Finally, counter electrodes capture electrons and holes in ETM and HTM, respectively, transport them to an external circuit, and generate current [55] [81]. Charge separation between MAPbI3 and HTM such as spiro-MeOTAD was observed in transient absorption spectroscopy, but electron injection at open-circuit conditions was not detected yet [61]. It has already been confirmed that HTM plays a crucial role in carrier separation and transport in PSCs [50] which will be discussed in their study for most of the inorganic HTMs used in PSCs.

    Figure 3. Energy level diagram and the carrier transport mechanism of perovskite solar cell in normal configuration (Interfaces in planar PSCs showing (1) HTL/perovskite interface, (2) perovskite/ETL interface, (3) ETL/cathode interface, and (4) HTL/anode interface).

    Particularly, there are primarily four types of interfaces in the inverted and/or normal structure of PSCs as shown in Figure 3. Each of the interfaces is methodically related to interfacial carrier dynamics including charge separation, charge injection, charge transport, charge collection, and recombination processes, and consequently affects how well the device functions in the end. Charge transport, extraction, and collection in real-time operation of PSCs are usually accompanied by charge recombination, which is closely related to PCE, stability, and hysteresis. It clearly shows that interface engineering is essential for developing effective and reliable PSCs

    References

    • 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.
    • Im, J.-H.; Lee, C.-R.; Lee, J.-W.; Park, S.-W.; Park, N.-G. 6.5% efficient perovskite quantum-dot-sensitized solar cell. Nanoscale 2011, 3, 4088–4093.
    • Green, M.A.; Hishikawa, Y.; Dunlop, E.D.; Levi, D.H.; Hohl-Ebinger, J.; Ho-Baillie, A.W. Solar cell efficiency tables (version 52). Prog. Photovolt. Res. Appl. 2018, 26, 427–436.
    • Saliba, M.; Matsui, T.; Seo, J.-Y.; Domanski, K.; Correa-Baena, J.-P.; Nazeeruddin, M.K.; Zakeeruddin, S.M.; Tress, W.; Abate, A.; Hagfeldt, A. Cesium-containing triple cation perovskite solar cells: Improved stability, reproducibility and high efficiency. Energy Environ. Sci. 2016, 9, 1989–1997.
    • Burschka, J.; Pellet, N.; Moon, S.-J.; Humphry-Baker, R.; Gao, P.; Nazeeruddin, M.K.; Grätzel, M. Sequential deposition as a route to high-performance perovskite-sensitized solar cells. Nature 2013, 499, 316–319.
    • Hao, F.; Stoumpos, C.C.; Cao, D.H.; Chang, R.P.; Kanatzidis, M.G. Lead-free solid-state organic–inorganic halide perovskite solar cells. Nat. Photon. 2014, 8, 489–494.
    • Dong, Q.; Fang, Y.; Shao, Y.; Mulligan, P.; Qiu, J.; Cao, L.; Huang, J. Electron-hole diffusion lengths 175 μm in solution-grown CH3NH3PbI3 single crystals. Science 2015, 347, 967–970.
    • Stranks, S.D.; Eperon, G.E.; Grancini, G.; Menelaou, C.; Alcocer, M.J.; Leijtens, T.; Herz, L.M.; Petrozza, A.; Snaith, H.J. Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber. Science 2013, 342, 341–344.
    • Xing, G.; Mathews, N.; Sun, S.; Lim, S.S.; Lam, Y.M.; Grätzel, M.; Mhaisalkar, S.; Sum, T.C. Long-range balanced electron-and hole-transport lengths in organic-inorganic CH3NH3PbI3. Science 2013, 342, 344–347.
    • Miyata, A.; Mitioglu, A.; Plochocka, P.; Portugall, O.; Wang, J.T.-W.; Stranks, S.D.; Snaith, H.J.; Nicholas, R.J. Direct measurement of the exciton binding energy and effective masses for charge carriers in organic–inorganic tri-halide perovskites. Nat. Phys. 2015, 11, 582–587.
    • Lin, Q.; Armin, A.; Nagiri, R.C.R.; Burn, P.L.; Meredith, P. Electro-optics of perovskite solar cells. Nat. Photon. 2015, 9, 106–112.
    • Brenner, T.M.; Egger, D.A.; Rappe, A.M.; Kronik, L.; Hodes, G.; Cahen, D. Are mobilities in hybrid organic–inorganic halide perovskites actually “high”? J. Phys. Chem. Lett. 2015, 6, 4754–4757.
    • Motta, C.; El-Mellouhi, F.; Sanvito, S. Charge carrier mobility in hybrid halide perovskites. Sci. Rep. 2015, 5, 12746.
    • Park, N.-G. Methodologies for high efficiency perovskite solar cells. Nano Converg. 2016, 3, 15.
    • Paek, S. Novel Anthracene HTM Containing TIPs for Perovskite Solar Cells. Processes 2021, 9, 2249.
    • Wang, S.; Yuan, W.; Meng, Y.S. Spectrum-dependent spiro-OMeTAD oxidization mechanism in perovskite solar cells. ACS Appl. Mater. Interfaces 2015, 7, 24791–24798.
    • Chiang, C.-H.; Wu, C.-G. Bulk heterojunction perovskite–PCBM solar cells with high fill factor. Nat. Photon. 2016, 10, 196–200.
    • Heo, J.H.; Im, S.H.; Noh, J.H.; Mandal, T.N.; Lim, C.-S.; Chang, J.A.; Lee, Y.H.; Kim, H.-j.; Sarkar, A.; Nazeeruddin, M.K. Efficient inorganic–organic hybrid heterojunction solar cells containing perovskite compound and polymeric hole conductors. Nat. Photon. 2013, 7, 486–491.
    • Seo, J.; Noh, J.H.; Seok, S.I. Rational strategies for efficient perovskite solar cells. Acc. Chem. Res. 2016, 49, 562–572.
    • Chen, H.-W.; Huang, T.-Y.; Chang, T.-H.; Sanehira, Y.; Kung, C.-W.; Chu, C.-W.; Ikegami, M.; Miyasaka, T.; Ho, K.-C. Efficiency enhancement of hybrid perovskite solar cells with MEH-PPV hole-transporting layers. Sci. Rep. 2016, 6, 34319.
    • Xiao, J.; Shi, J.; Liu, H.; Xu, Y.; Lv, S.; Luo, Y.; Li, D.; Meng, Q.; Li, Y. Efficient CH3NH3PbI3 Perovskite Solar Cells Based on Graphdiyne (GD)-Modified P3HT Hole-Transporting Material. Adv. Energy Mater. 2015, 5, 1401943.
    • Bi, D.; Yang, L.; Boschloo, G.; Hagfeldt, A.; Johansson, E.M. Effect of different hole transport materials on recombination in CH3NH3PbI3 perovskite-sensitized mesoscopic solar cells. J. Phys. Chem. Lett. 2013, 4, 1532–1536.
    • Son, D.-Y.; Im, J.-H.; Kim, H.-S.; Park, N.-G. 11% efficient perovskite solar cell based on ZnO nanorods: An effective charge collection system. J. Phys. Chem. C 2014, 118, 16567–16573.
    • Zhang, H.; Xue, L.; Han, J.; Fu, Y.Q.; Shen, Y.; Zhang, Z.; Li, Y.; Wang, M. New generation perovskite solar cells with solution-processed amino-substituted perylene diimide derivative as electron-transport layer. J. Mater. Chem. A 2016, 4, 8724–8733.
    • Xiao, Y.; Han, G.; Chang, Y.; Zhou, H.; Li, M.; Li, Y. An all-solid-state perovskite-sensitized solar cell based on the dual function polyaniline as the sensitizer and p-type hole-transporting material. J. Power Sources 2014, 267, 1–8.
    • Arora, N.; Dar, M.I.; Hinderhofer, A.; Pellet, N.; Schreiber, F.; Zakeeruddin, S.M.; Grätzel, M. Perovskite solar cells with CuSCN hole extraction layers yield stabilized efficiencies greater than 20%. Science 2017, 358, 768–771.
    • Jung, J.W.; Chueh, C.C.; Jen, A.K.Y. A low-temperature, solution-processable, Cu-doped nickel oxide hole-transporting layer via the combustion method for high-performance thin-film perovskite solar cells. Adv. Mater. 2015, 27, 7874–7880.
    • Sun, W.; Li, Y.; Ye, S.; Rao, H.; Yan, W.; Peng, H.; Li, Y.; Liu, Z.; Wang, S.; Chen, Z. High-performance inverted planar heterojunction perovskite solar cells based on a solution-processed CuOx hole transport layer. Nanoscale 2016, 8, 10806–10813.
    • Huangfu, M.; Shen, Y.; Zhu, G.; Xu, K.; Cao, M.; Gu, F.; Wang, L. Copper iodide as inorganic hole conductor for perovskite solar cells with different thickness of mesoporous layer and hole transport layer. Appl. Surf. Sci. 2015, 357, 2234–2240.
    • Zhang, H.; Wang, H.; Chen, W.; Jen, A.K.Y. CuGaO2: A promising inorganic hole-transporting material for highly efficient and stable perovskite solar cells. Adv. Mater. 2017, 29, 1604984.
    • Igbari, F.; Li, M.; Hu, Y.; Wang, Z.-K.; Liao, L.-S. A room-temperature CuAlO2 hole interfacial layer for efficient and stable planar perovskite solar cells. J. Mater. Chem. A 2016, 4, 1326–1335.
    • Sung, H.; Ahn, N.; Jang, M.S.; Lee, J.K.; Yoon, H.; Park, N.G.; Choi, M. Transparent conductive oxide-free graphene-based perovskite solar cells with over 17% efficiency. Adv. Energy Mater. 2016, 6, 1501873.
    • Rao, H.; Sun, W.; Ye, S.; Yan, W.; Li, Y.; Peng, H.; Liu, Z.; Bian, Z.; Huang, C. Solution-processed CuS NPs as an inorganic hole-selective contact material for inverted planar perovskite solar cells. ACS Appl. Mater. Interfaces 2016, 8, 7800–7805.
    • Singh, E.; Kim, K.S.; Yeom, G.Y.; Nalwa, H.S. Atomically thin-layered molybdenum disulfide (MoS2) for bulk-heterojunction solar cells. ACS Appl. Mater. Interfaces 2017, 9, 3223–3245.
    • Bhattacharya, B.; Singh, P.K.; Singh, R.; Khan, Z.H. Perovskite sensitized solar cell using solid polymer electrolyte. Int. J. Hydrog. Energy 2016, 41, 2847–2852.
    • Yu, Z.; Sun, L. Recent progress on hole-transporting materials for emerging organometal halide perovskite solar cells. Adv. Energy Mater. 2015, 5, 1500213.
    • Al Mamun, A.; Ava, T.T.; Zhang, K.; Baumgart, H.; Namkoong, G. New PCBM/carbon based electron transport layer for perovskite solar cells. Phys. Chem. Chem. Phys. 2017, 19, 17960–17966.
    • Aitola, K.; Sveinbjörnsson, K.; Correa-Baena, J.-P.; Kaskela, A.; Abate, A.; Tian, Y.; Johansson, E.M.; Grätzel, M.; Kauppinen, E.I.; Hagfeldt, A. Carbon nanotube-based hybrid hole-transporting material and selective contact for high efficiency perovskite solar cells. Energy Environ. Sci. 2016, 9, 461–466.
    • Cao, J.; Liu, Y.-M.; Jing, X.; Yin, J.; Li, J.; Xu, B.; Tan, Y.-Z.; Zheng, N. Well-defined thiolated nanographene as hole-transporting material for efficient and stable perovskite solar cells. J. Am. Chem. Soc. 2015, 137, 10914–10917.
    • Singh, R.; Jun, H.; Arof, A.K. Activated carbon as back contact for HTM-free mixed cation perovskite solar cell. Phase Transit. 2018, 91, 1268–1276.
    • Wang, Q.; Lin, Z.; Su, J.; Hu, Z.; Chang, J.; Hao, Y. Recent progress of inorganic hole transport materials for efficient and stable perovskite solar cells. Nano Sel. 2021, 2, 1055–1080.
    • Li, S.; Cao, Y.-L.; Li, W.-H.; Bo, Z.-S. A brief review of hole transporting materials commonly used in perovskite solar cells. Rare Met. 2021, 40, 2712–2729.
    • Park, H.; Chaurasiya, R.; Ho Jeong, B.; Sakthivel, P.; Joon Park, H. Nickel Oxide for Perovskite Photovoltaic Cells. Adv. Photon. Res. 2021, 2, 2000178.
    • Cai, L.; Zhu, F. Toward efficient and stable operation of perovskite solar cells: Impact of sputtered metal oxide interlayers. Nano Sel. 2021, 2, 1417–1436.
    • Arumugam, G.M.; Karunakaran, S.K.; Liu, C.; Zhang, C.; Guo, F.; Wu, S.; Mai, Y. Inorganic hole transport layers in inverted perovskite solar cells: A review. Nano Sel. 2021, 2, 1081–1116.
    • Di Girolamo, D.; Di Giacomo, F.; Matteocci, F.; Marrani, A.G.; Dini, D.; Abate, A. Progress, highlights and perspectives on NiO in perovskite photovoltaics. Chem. Sci. 2020, 11, 7746–7759.
    • Pitchaiya, S.; Natarajan, M.; Santhanam, A.; Asokan, V.; Yuvapragasam, A.; Ramakrishnan, V.M.; Palanisamy, S.E.; Sundaram, S.; Velauthapillai, D. A review on the classification of organic/inorganic/carbonaceous hole transporting materials for perovskite solar cell application. Arab. J. Chem. 2020, 13, 2526–2557.
    • Singh, R.; Singh, P.K.; Bhattacharya, B.; Rhee, H.-W. Review of current progress in inorganic hole-transport materials for perovskite solar cells. Appl. Mater. Today 2019, 14, 175–200.
    • Tai, Q.; Tang, K.-C.; Yan, F. Recent progress of inorganic perovskite solar cells. Energy Environ. Sci. 2019, 12, 2375–2405.
    • Chen, J.; Park, N.-G. Inorganic hole transporting materials for stable and high efficiency perovskite solar cells. J. Phys. Chem. C 2018, 122, 14039–14063.
    • Bhat, A.; Dhamaniya, B.P.; Chhillar, P.; Korukonda, T.B.; Rawat, G.; Pathak, S.K. Analysing the Prospects of Perovskite Solar Cells within the Purview of Recent Scientific Advancements. Crystals 2018, 8, 242.
    • Qin, X.; Zhao, Z.; Wang, Y.; Wu, J.; Jiang, Q.; You, J. Recent progress in stability of perovskite solar cells. J. Semicond. 2017, 38, 011002.
    • Rajeswari, R.; Mrinalini, M.; Prasanthkumar, S.; Giribabu, L. Emerging of inorganic hole transporting materials for perovskite solar cells. Chem. Rec. 2017, 17, 681–699.
    • Meng, L.; You, J.; Guo, T.-F.; Yang, Y. Recent advances in the inverted planar structure of perovskite solar cells. Acc. Chem. Res. 2016, 49, 155–165.
    • Fan, R.; Huang, Y.; Wang, L.; Li, L.; Zheng, G.; Zhou, H. The progress of interface design in perovskite-based solar cells. Adv. Energy Mater. 2016, 6, 1600460.
    • 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. Perovskite solar cells with atomically coherent interlayers on SnO2 electrodes. Nature 2021, 598, 444–450.
    • Ibn-Mohammed, T.; Koh, S.; Reaney, I.; Acquaye, A.; Schileo, G.; Mustapha, K.; Greenough, R. Perovskite solar cells: An integrated hybrid lifecycle assessment and review in comparison with other photovoltaic technologies. Renew. Sustain. Energy Rev. 2017, 80, 1321–1344.
    • Ball, J.M.; Stranks, S.D.; Hörantner, M.T.; Hüttner, S.; Zhang, W.; Crossland, E.J.; Ramirez, I.; Riede, M.; Johnston, M.B.; Friend, R.H. Optical properties and limiting photocurrent of thin-film perovskite solar cells. Energy Environ. Sci. 2015, 8, 602–609.
    • Guo, T.; Lin, M.; Huang, J.; Zhou, C.; Tian, W.; Yu, H.; Jiang, X.; Ye, J.; Shi, Y.; Xiao, Y. The recent advances of magnetic nanoparticles in medicine. J. Nanomater. 2018, 2018, 7805147.
    • Grancini, G.; Roldán-Carmona, C.; Zimmermann, I.; Mosconi, E.; Lee, X.; Martineau, D.; Narbey, S.; Oswald, F.; De Angelis, F.; Graetzel, M. One-Year stable perovskite solar cells by 2D/3D interface engineering. Nat. Commun. 2017, 8, 15684.
    • Kim, H.-S.; Lee, C.-R.; Im, J.-H.; Lee, K.-B.; Moehl, T.; Marchioro, A.; Moon, S.-J.; Humphry-Baker, R.; Yum, J.-H.; Moser, J.E. Lead iodide perovskite sensitized all-solid-state submicron thin film mesoscopic solar cell with efficiency exceeding 9%. Sci. Rep. 2012, 2, 591.
    • Rong, Y.; Hu, Y.; Mei, A.; Tan, H.; Saidaminov, M.I.; Seok, S.I.; McGehee, M.D.; Sargent, E.H.; Han, H. Challenges for commercializing perovskite solar cells. Science 2018, 361, eaat8235.
    • Green, M.A.; Ho-Baillie, A.; Snaith, H.J. The emergence of perovskite solar cells. Nat. Photon. 2014, 8, 506–514.
    • Kazim, S.; Nazeeruddin, M.K.; Grätzel, M.; Ahmad, S. Perovskite as light harvester: A game changer in photovoltaics. Angew. Chem. Int. Ed. 2014, 53, 2812–2824.
    • Bretschneider, S.A.; Weickert, J.; Dorman, J.A.; Schmidt-Mende, L. Research update: Physical and electrical characteristics of lead halide perovskites for solar cell applications. APL Mater. 2014, 2, 155204.
    • Mitzi, D.B. Synthesis, structure, and properties of organic-inorganic perovskites and related materials. Prog. Inorg. Chem. 1999, 48, 1–121.
    • Correa-Baena, J.-P.; Abate, A.; Saliba, M.; Tress, W.; Jacobsson, T.J.; Grätzel, M.; Hagfeldt, A. The Rapid evolution of highly efficient perovskite solar cells. Energy Environ. Sci. 2017, 10, 710–727.
    • Jiang, X.; Yu, Z.; Lai, J.; Zhang, Y.; Lei, N.; Wang, D.; Sun, L. Efficient perovskite solar cells employing a solution-processable copper phthalocyanine as a hole-transporting material. Sci. China Chem. 2017, 60, 423–430.
    • Tzounis, L.; Stergiopoulos, T.; Zachariadis, A.; Gravalidis, C.; Laskarakis, A.; Logothetidis, S. Perovskite solar cells from small scale spin coating process towards roll-to-roll printing: Optical and morphological studies. Mater. Today Proc. 2017, 4, 5082–5089.
    • Wang, G.; Wu, F.; Wu, R.; Chen, T.; Ding, B.F.; Song, Q.L. Crystallization process of perovskite modified by adding lead acetate in precursor solution for better morphology and higher device efficiency. Org. Electron. 2017, 43, 189–195.
    • Guo, F.; He, W.; Qiu, S.; Wang, C.; Liu, X.; Forberich, K.; Brabec, C.J.; Mai, Y. Sequential Deposition of High-Quality Photovoltaic Perovskite Layers via Scalable Printing Methods. Adv. Funct. Mater. 2019, 29, 1900964.
    • Li, H.; Li, S.; Wang, Y.; Sarvari, H.; Zhang, P.; Wang, M.; Chen, Z. A modified sequential deposition method for fabrication of perovskite solar cells. Sol. Energy 2016, 126, 243–251.
    • Reinoso, M.Á.; Otálora, C.A.; Gordillo, G. Improvement Properties of Hybrid Halide Perovskite Thin Films Prepared by Sequential Evaporation for Planar Solar Cells. Materials 2019, 12, 1394.
    • Ma, Q.; Huang, S.; Wen, X.; Green, M.A.; Ho-Baillie, A.W. Hole transport layer free inorganic CsPbIBr2 perovskite solar cell by dual source thermal evaporation. Adv. Energy Mater. 2016, 6, 1502202.
    • Yang, M.; Li, Z.; Reese, M.O.; Reid, O.G.; Kim, D.H.; Siol, S.; Klein, T.R.; Yan, Y.; Berry, J.J.; Van Hest, M.F. Perovskite ink with wide processing window for scalable high-efficiency solar cells. Nat. Energy 2017, 2, 17038.
    • Jung, M.; Ji, S.-G.; Kim, G.; Seok, S.I. Perovskite precursor solution chemistry: From fundamentals to photovoltaic applications. Chem. Soc. Rev. 2019, 48, 2011–2038.
    • Lee, M.M.; Teuscher, J.; Miyasaka, T.; Murakami, T.N.; Snaith, H.J. Efficient hybrid solar cells based on meso-superstructured organometal halide perovskites. Science 2012, 338, 643–647.
    • Ke, W.; Fang, G.; Liu, Q.; Xiong, L.; Qin, P.; Tao, H.; Wang, J.; Lei, H.; Li, B.; Wan, J. Low-temperature solution-processed tin oxide as an alternative electron transporting layer for efficient perovskite solar cells. J. Am. Chem. Soc. 2015, 137, 6730–6733.
    • Etgar, L.; Gao, P.; Xue, Z.; Peng, Q.; Chandiran, A.K.; Liu, B.; Nazeeruddin, M.K.; Grätzel, M. Mesoscopic CH3NH3PbI3/TiO2 heterojunction solar cells. J. Am. Chem. Soc. 2012, 134, 17396–17399.
    • Wang, K.-C.; Jeng, J.-Y.; Shen, P.-S.; Chang, Y.-C.; Diau, E.W.-G.; Tsai, C.-H.; Chao, T.-Y.; Hsu, H.-C.; Lin, P.-Y.; Chen, P. P-type mesoscopic nickel oxide/organometallic perovskite heterojunction solar cells. Sci. Rep. 2014, 4, 4756.
    • Jung, H.S.; Park, N.G. Perovskite solar cells: From materials to devices. Small 2015, 11, 10–25.

    © Text is available under the terms and conditions of the Creative Commons Attribution (CC BY) license; additional terms may apply. By using this site, you agree to the Terms and Conditions and Privacy Policy.

    Perovskite solar cells: Hero, villain or just plain fantasy?

    With my experience throughout the PV industry supply chain and thin-film commercialization, perovskite solar appeared at first glance to be yet another unsubstantiated claim of printing and spraying simple solar cells on everything to provide the world with endless, cheap energy. But what was missing from those previous efforts was a material that could be high efficiency, exponentially lower cost and highly manufacturable.

    There is a lot of misunderstanding in the industry around the perovskite solar cell (PSC), and often reporting from outside the industry adds to the confusion. As with all new technologies, irrational exuberance and naysaying abound.

    So let’s start by getting a few baseline items out of the way.

    First the pronunciation: pe·​rov·​skite | \ pə-ˈräv-ˌskīt — it took me forever to get this right!

    Record efficiencies of perovskite solar cells shown in the yellow circles. Credit: NREL

    Perovskite is comprised of a class of minerals in crystalline cube- and diamond-like structures that were first discovered over 170 years ago in the Ural Mountains of Russia. Named after Russian mineralogist Lev Perovski, it’s abundantly found mainly in the earth’s mantle and occasionally in near-surface deposits. Fortunately, perovskite can be synthesized from common chemicals for use in solar cells and other applications including light emitting diodes, catalyst electrodes, fuel cells, IC chips, lasers and sensors, to name a few.

    Perovskite as a solar cell material was initially discovered when it was used in lab-tested organic dye sensitized solar cells in the mid-2000s. While that solar cell effort was not successful, the perovskite compound was shown to be quite photo-reactive by itself. Fast forward 10 years and academia, national labs and corporations have taken the PSC efficiency from 2% to 25% in lab settings on approximately 30-cm x 30-cm forms. It’s a staggering advancement when looking at the history of other solar cell technologies that took 40 years to reach similar lab-scale efficiencies.

    Perovskite offers a large number of advantages in the solar cell/module realm:

    • It has wide bandgap, which results in greater tunability and more sunlight being converted to electricity.
    • Current lab efficiencies mirror those in silicon crystalline and other thin-film products, and tandem PSCs have a viable roadmap to reach greater than 33%.
    • Perovskite production is highly simplified via solutions processing and does not require high-cost machinery and facilities, which are required for semiconductor processing.
    • It is manufactured as a thin-film product resulting in 20-times less materials.
    • It requires no rare earth or supply-limited materials.
    • Perovskite is highly defect-tolerant which creates high manufacturing yields and ease when working in modules larger than 300 W (current thin-films require exceptional engineering, cost and risk to increase the manufacturing deposition area).
    • PSC-based modules can be shaped as traditional rectangular solar panels or be flexible, opening up new applications and markets.
    • Perovskite solar module manufacturing has a small environmental footprint depending on the manufacturing method employed.
    • Return on energy invested is measured in months vs. years with current solar module products.

    So what’s not to like here? When looking at perovskite solar cells, the number of misunderstandings are large.

    Stability

    Light-induced cell degradation and environmental stability are the continually cited villains. Early on, PSC stability was a large problem. But just as there were Rapid advancements in cell efficiency, there has also been similar quick advancements in stability.

    Perovskite Mineral Crystal

    Light-induced degradation is now well understood. A major hinderance was the choice of materials for the electron transfer layers on either side of the perovskite cell. Multiple labs and other entities have successfully substituted new materials and solved LID issues.

    On the weatherization front, just like with other thin-films and crystalline silicon solar cells, exposure to moisture, oxygen and other common environmental substances can cause Rapid module degradation in PSCs. Using standard “packaging” assembly schemes, the concerns about environmental degradation have largely been solved. This reminds me of the early days of copper indium gallium selenide (CIGS) cell development where the commonly cited bogeyman was intolerance to moisture. But now CIGS-based modules are widely deployed with no major environmental stability issues.

    Both light-induced and environmental PSC degradation parameters have exceeded 1,000-hour accelerated life cycle testing (the PV industry standard for new technologies) with some going as far as 10,000 hours.

    Other stability challenges related to mechanical durability, applied voltage heating and thermal influences and current-voltage behaviors are similarly understood with a number of fixes in the testing phase.

    Two whitepapers on light-induced and environmental stability cures can be found here and here. A more recent paper from Solliance partners TNO, imec and the Eindhoven University of Technology shows encapsulated perovskite solar modules using standard manufacturing methods successfully completed standard PV industry stability tests including light soak, damp-heat and thermal cycling.

    Lead

    Lead is used in the most common PSC cell structure, methylammonium lead halide. The amount of lead used is miniscule (average: 2 mg/Watt). For comparison, a common automobile battery contains 20 pounds (9,000,000 mg) on average.

    As lead is a heavy toxic metal substance, it needs to be tightly controlled in all parts of an application from manufacturing to end-of-life recycling, especially when it is deployed in a water soluble form such as a perovskite solar cell. The lead recycling stream is the largest and most complete of any material in the world and the PSC industry would use only 0.0007 grams of the total stream. Looking at it another way, PSC recycling would annually represent one-tenth of 1% of all lead recycled at a 1-terawatt recycling rate.

    Looking at current fossil fuel energy generation — where lead and other heavy metal particulates and toxic waste streams permeate our daily lives — the benefits of solar modules with only trace amounts of heavy metals far outweighs the overall risk.

    Scalability to meet zero emissions by 2050 mandate

    The topic of whether PSCs can scale fast enough to meet the United Nations’ 2050 zero emissions targets is a continual conversation. The established crystalline silicon solar module based industry is pointed to as mature and ready to scale. But silicon-based solar faces challenges to meet these energy needs.

    A perovskite cell using methylammonium lead triiodide thin films. Photo credit: Dennis Schroeder / NREL

    To meet 20% of all global energy for the U.N.’s emissions reduction scenario, the solar industry would need to install 300 to 500 GW annually (linear example) over the next 30 years. Current global PV module manufacturing at 100 GW per year presents an enormous supply chain challenge. The silicon module industry’s large capital intensity issues limits solar supply to 5% of all global energy. It’s a common misunderstanding that silicon modules’ low margins are principally because they are highly commoditized products. The capital intensity is preventing large and durable margins, contributing to supply chain scale-up financing challenges.

    The cost of solar energy is also still too expensive especially when looking at value deflation at the kilowatt-hour level. Looking at many studies, solar energy needs to have an installed cost lower than 0.30/W to solve this problem even when coupled with low cost energy storage.

    Perovskite solar with its unique simplified manufacturing attributes, raw material, performance and small environmental footprint makes it highly scalable — quickly. Depending on the business model, perovskite modules can be manufactured in facilities that cost 50% less than other solar factories and use less materials. The supply chain to support PSC manufacturing is also small, allowing for factories to be sited close to end markets.

    The PSC industry is in a state where its lab developments exceed prior solar panel commercialization launch points. The challenge is how quickly the simplified manufacturing can scale-up for a given production method to enter high-volume production.

    If you believe, as I do, that the acceleration of climate-related events will also accelerate global energy decarbonization timelines, the solar industry will need all existing and new arrows in the module sector quiver to meet 20% of all global energy, let alone 45% as some scenarios demand. Perovskite-based modules can be a timely addition to the solar energy industry’s push to zero energy emissions.

    Check back for the second article in this series, focusing on product design, raw materials and manufacturing.

    Leveraging a 23-year career in renewable energy with broad experience across the solar PV industry supply chain, Dave Buemi is Managing Director of Prescient Energy Consulting which provides strategic and tactical business consulting to renewable energy related entities throughout the low carbon energy transition ecosystem. He has held senior-level positions throughout the PV supply chain including technology commercialization, manufacturing, EPC and project development with companies that include Brightphase Energy, Daystar Technologies, Empower Energies, Gehrlicher Solar/MW, Suniva and Willdan Energy Solutions. He also played a key role developing the community energy model and commercialization of both advanced thin-film solar cell and solar tri-generation technologies. His work early in renewable energy overlapped with the U.S. Department of Defense renewable energy and microgrid strategy deployment where he provided strategy and knowledge in the warrior, forward deployed, home base and airborne sectors. A climate change and climate resiliency activist and consultant, Dave believes that urgent innovation throughout the PV industry ecosystem is key to meeting the Paris Climate Accords 2050 timeline while enabling a more profitable and stable industry for all stakeholders.

    Opinions expressed here are of the author only.

    Комментарии и мнения владельцев

    I AM AN IGNORANT RETIRED SOLICITOR BUT HAVE HOBBY READ SCIENCE JOURNALS FOR SOME 60 YEARS. QUERY: WOULD ENCAPSULATED PEROVSKITE CELLS BE IMPROVED BY USING A LIGHT CONDENSER THIN FILM FRESNEL LENS APPROACH ? I think that I have read that HEAT is not a problem – rather an opposite. Best regards

    An opt answer for to day’s Global mission of achieving 20 % power requirement through non polluting resources. Since the technology for manufacturing is understood the commercialization needs to be fastened solving the issues as we come across.I have been working in the areas of energy efficiency and alternate energy sources as ahead of an organization helping establishing small and medium sized manufacturing companies. Beeing a chemical engineer my self I can play a role.

    I believe some of the ‘new’ technologies of the mono crystalline solar PV cell manufacturing processes could be adaptable to perovskites. In adapting topologies like TOPcon cell manufacturing, could allow two differently tuned perovskite chemistries to be sprayed onto a mono crystalline cell structure with the proper ‘moisture sealing’ as part of the process. Tandem cells, what about tertiary cells in one manufacturing line process?

    Perovskite Solar Cell Market Analysis. 2030

    The global Perovskite solar cell market size was valued at 0.4 billion in 2020, and is forecasted to reach 6.6 billion by 2030, growing at a CAGR of 32.4% from 2021 to 2030. Perovskite solar cell (PSC) includes the perovskite-structured material as an active layer based on the solution processed by tin or halide. Perovskite materials offer excellent light absorption, charge-carrier mobilities, and lifetimes, resulting in high device efficiencies with opportunities to realize a low-cost, industry-scalable technology. Perovskite solar cells (PSCs) are the most emerging area of research among different new generation photovoltaic technologies, due to their super power conversion efficiency (PCE).

    Growing demand for solar cells due to their flexibility and light weight, rising number of applications in various industries, increasing environmental concerns about carbon emission reduction, and the prevalence of alternative energy sources are some of the factors that are expected to boost the growth of the perovskite solar cell market during the forecast period. Furthermore, increasing RD efforts as well as technical developments, can enhance different prospects, resulting in the expansion of the market throughout the forecast period. However, presence of toxic material as well as high cost of product may hamper the growth of market during the forecast period. These are some of the perovskite solar cell market trends observed globally. The Perovskite solar cell market is segmented on the basis of structure, product, method, end-use, and region. On the basis of structure, the global market is segmented into planar perovskite solar cells and mesoporous perovskite solar cells. The Product of Perovskite solar cell include rigid perovskite solar cells and flexible perovskite solar cells. On the basis of method, it is segmented into solution method, vapor-deposition method and vapor-assisted solution method. The end use of Perovskite solar cell includes aerospace, industrial automation, consumer electronics, energy and others. On the basis of region, the global market is studied across North America, Europe, Asia-Pacific, and LAMEA. Presently, Europe accounts for the largest share of the market, followed by North America. Major players operating in the global Perovskite solar cell industry include Oxford Photovoltaics, FrontMaterials Co. Ltd., Solaronix SA, Xiamen Weihua Solar Co. Ltd., Fraunhofer ISE, Dyesol, Saule Technologies, FlexLink Systems Inc., Polyera Corporation, and New Energy Technologies Inc.

    Get more information on this report : Request Sample Pages

    Perovskite solar cell Market, by Structure

    The planar perovskite solar cells segment accounted for the largest share in 2020, while the mesoporous perovskite solar cells segment is projected to grow at the highest CAGR of 32.8%.

    Get more information on this report : Request Sample Pages

    Perovskite solar cell Market, by Method

    The vapor-deposition method segment accounted for the largest share in 2020, while the vapor-assisted solution method segment is projected to grow at the highest CAGR of 32.9%.

    Get more information on this report : Request Sample Pages

    Perovskite solar cell Market, by Product

    The flexible perovskite solar cells segment accounted for the largest share in 2020, while the rigid perovskite solar cells segment is projected to grow at the highest CAGR of 32.7%.

    Get more information on this report : Request Sample Pages

    Perovskite solar cell Market, by End Use

    The energy segment accounted for the largest share of the Perovskite solar cell market in 2020.

    Get more information on this report : Request Sample Pages

    Perovskite solar cell Market, by Region

    The market is expected to grow during forecast period, owing to a growth in electronic advances in countries such as Japan, China, India, and South Korea. As a result, throughout the projected period, Asia Pacific is expected to lead the market.

    Get more information on this report : Request Sample Pages

    Key Benefits For Stakeholders

    • The report provides in-depth analysis of the global Perovskite solar cell market along with the current trends and future estimations.
    • This report highlights the key drivers, opportunities, and restraints of the market along with the impact analysis during the forecast period.
    • Porter’s five forces analysis helps to analyze the potential of the buyers suppliers and the competitive scenario of the global market for strategy building.
    • A comprehensive market analysis covers the factors that drive and restrain the global Perovskite solar cell market growth.
    • The qualitative data about market dynamics, trends, and developments is provided in the report.

    Impact Of Covid-19 On The Global Perovskite Solar Cell Market

    • COVID-19 has spread to almost 213 countries around the globe with the World Health Organization declaring it a public health emergency on March 11, 2020.
    • Some of the major economies suffering from the COVID-19 crises include Germany, France, Italy, Spain, the UK, and Norway.
    • Perovskite solar cell is primarily used in supplements, cosmetics, industrial, and personal care.
    • In many countries, the economy has dropped due to the halt of several industries, especially transport and supply chain. Demand for the product has been hindered as there is no development due to the lockdown.
    • The demand–supply gap, disruptions in raw material procurement, and price volatility are expected to hamper the growth of the industry during the COVID-19 pandemic.
    • Due to a scarcity of resources in various parts of the world, the COVID-19 epidemic has impacted negatively on the manufacturing and industrial industries. The industry’s top players are concerned about the market’s prospects and are rethinking their strategies to meet the challenge.

    Key Market Segments

    • By STRUCTURE
    • Planar perovskite solar cells
    • Mesoporous perovskite solar cells
    • Rigid Perovskite Solar Cells
    • Flexible Perovskite Solar Cells
    • Solution Method
    • Vapor-Deposition Method
    • Vapor-Assisted Solution Method
    • Aerospace
    • Industrial automation
    • Consumer Electronics
    • Energy
    • Others
    • North America
    • U.S.
    • Canada
    • Mexico
    • Germany
    • UK
    • France
    • Italy
    • Spain
    • Rest of Europe
    • China
    • Japan
    • India
    • South Korea
    • Australia
    • Rest of Asia-Pacific
    • Brazil
    • Saudi Arabia
    • South Africa
    • Rest of LAMEA

    Key Market Players

    • New Energy Technologies Inc
    • FlexLink Systems Inc.
    • Fraunhofer ISE
    • Polyera Corporation
    • Dyesol
    • Oxford Photovoltaics
    • Saule Technologies
    • FrontMaterials Co. Ltd.
    • Solaronix SA
    • Xiamen Weihua Solar Co. Ltd.

    1.1.Report description1.2.Key benefits for stakeholders1.3.Key market segments1.4.Research methodology

    1.4.1.Primary research1.4.2.Secondary research

    1.5.Analyst tools and models

    CHAPTER 2:EXECUTIVE SUMMARY

    2.1.Key findings of the study2.2.CXO perspective

    CHAPTER 3:MARKET LANDSCAPE

    3.1.Market definition and scope3.2.Key findings

    3.2.1.Top investment s

    3.3.Porter’s five forces analysis3.5.Market share analysis top player positioning, 2020

    3.5.1.Top player positioning, 2020

    3.6.1.1.Advantages of perovskite solar cells

    3.6.2.1.Presence of toxic material

    3.6.3.1.Increase in research and development activities

    3.7.Value chain3.8.Impact of key regulations on the global pervoskite solar cell market3.9.Impact of COVID-19 outburst on the Pervoskite solar cell market

    CHAPTER 4:PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE

    4.1.1.Market size and forecast

    4.2.Planar perovskite solar cells

    4.2.1.Key market trends, growth factors, and opportunities4.2.2.Market size and forecast, by region4.2.3.Market analysis, by country

    4.3.Mesoporous perovskite solar cells

    4.3.1.Key market trends, growth factors, and opportunities4.3.3.Market size and forecast, by region4.3.4.Market analysis, by country

    CHAPTER 5:PERVOSKITE SOLAR CELL MARKET, BY PRODUCT

    5.1.1.Market size and forecast

    5.2.Rigid Perovskite Solar Cells

    5.2.1.Key market trends, growth factors, and opportunities5.2.2.Market size and forecast, by region5.2.3.Market analysis, by country

    5.3.Flexible Perovskite Solar Cells

    5.3.1.Key market trends, growth factors, and opportunities5.3.3.Market size and forecast, by region5.3.4.Market analysis, by country

    CHAPTER 6:PERVOSKITE SOLAR CELL MARKET, BY METHOD

    6.1.1.Market size and forecast

    6.2.1.Key market trends, growth factors, and opportunities6.2.2.Market size and forecast, by region6.2.3.Market analysis, by country

    6.3.1.Key market trends, growth factors, and opportunities6.3.3.Market size and forecast, by region6.3.4.Market analysis, by country

    6.4.Vapor-Assisted Solution Method

    6.4.1.Key market trends, growth factors, and opportunities6.4.2.Market size and forecast, by region6.4.3.Market analysis, by country

    CHAPTER 7:PERVOSKITE SOLAR CELL MARKET, BY END-USER

    7.1.1.Market size and forecast

    7.2.1.Key market trends, growth factors, and opportunities7.2.2.Market size and forecast, by region7.2.3.Market analysis, by country

    7.3.1.Key market trends, growth factors, and opportunities7.3.2.Market size and forecast, by region7.3.3.Market analysis, by country

    7.4.1.Key market trends, growth factors, and opportunities7.4.2.Market size and forecast, by region7.4.3.Market analysis, by country

    7.5.1.Key market trends, growth factors, and opportunities7.5.2.Market size and forecast, by region7.5.3.Market analysis, by country

    7.6.1.Key market trends, growth factors, and opportunities7.6.2.Market size and forecast, by region7.6.3.Market analysis, by country

    CHAPTER 8:PERVOSKITE SOLAR CELL MARKET, BY REGION

    8.1.1.Market size and forecast

    8.2.1.Key market trends, growth factors, and opportunities8.2.2.Market size and forecast, by Structure8.2.3.Market size and forecast, by Product8.2.4.Market size and forecast, by Method8.2.5.Market size and forecast, by end-user8.2.6.Market share analysis, by country

    8.2.6.1.1.Market size and forecast, by Structure8.2.6.1.2.Market size and forecast, by Product8.2.6.1.3.Market size and forecast, by Method8.2.6.1.4.Market size and forecast, by end-user

    8.2.6.2.1.Market size and forecast, by Structure8.2.6.2.2.Market size and forecast, by Product8.2.6.2.3.Market size and forecast, by Method8.2.6.2.4.Market size and forecast, by end-user

    8.2.6.3.1.Market size and forecast, by Structure8.2.6.3.2.Market size and forecast, by Product8.2.6.3.3.Market size and forecast, by Method8.2.6.3.4.Market size and forecast, by end-user

    8.3.1.Key market trends, growth factors, and opportunities8.3.2.Market size and forecast, by Structure8.3.3.Market size and forecast, by Product8.3.4.Market size and forecast, by Method8.3.5.Market size and forecast, by end-user8.3.6.Market share analysis, by country

    8.3.6.1.1.Market size and forecast, by Structure8.3.6.1.2.Market size and forecast, by Product8.3.6.1.3.Market size and forecast, by Method8.3.6.1.4.Market size and forecast, by end-user

    8.3.6.2.1.Market size and forecast, by Structure8.3.6.2.2.Market size and forecast, by Product8.3.6.2.3.Market size and forecast, by Method8.3.6.2.4.Market size and forecast, by end-user

    8.3.6.3.1.Market size and forecast, by Structure8.3.6.3.2.Market size and forecast, by Product8.3.6.3.3.Market size and forecast, by Method8.3.6.3.4.Market size and forecast, by end-user

    8.3.6.4.1.Market size and forecast, by Structure8.3.6.4.2.Market size and forecast, by Product8.3.6.4.3.Market size and forecast, by Method8.3.6.4.4.Market size and forecast, by end-user

    8.3.6.5.1.Market size and forecast, by Structure8.3.6.5.2.Market size and forecast, by Product8.3.6.5.3.Market size and forecast, by Method8.3.6.5.4.Market size and forecast, by end-user

    8.3.6.6.1.Market size and forecast, by Structure8.3.6.6.2.Market size and forecast, by Product8.3.6.6.3.Market size and forecast, by Method8.3.7.Market size and forecast, by end-user

    8.4.1.Key market trends, growth factors, and opportunities8.4.2.Market size and forecast, by Structure8.4.3.Market size and forecast, by Product8.4.4.Market size and forecast, by Method8.4.5.Market size and forecast, by end-user8.4.6.Market share analysis, by country

    8.4.6.1.1.Market size and forecast, by Structure8.4.6.1.2.Market size and forecast, by Product8.4.6.1.3.Market size and forecast, by Method8.4.6.1.4.Market size and forecast, by end-user

    8.4.6.2.1.Market size and forecast, by Structure8.4.6.2.2.Market size and forecast, by Product8.4.6.2.3.Market size and forecast, by Method8.4.6.2.4.Market size and forecast, by end-user

    8.4.6.3.1.Market size and forecast, by Structure8.4.6.3.2.Market size and forecast, by Product8.4.6.3.3.Market size and forecast, by Method8.4.6.3.4.Market size and forecast, by end-user

    8.4.6.4.1.Market size and forecast, by Structure8.4.6.4.2.Market size and forecast, by Product8.4.6.4.3.Market size and forecast, by Method8.4.6.4.4.Market size and forecast, by end-user

    8.4.6.5.1.Market size and forecast, by Structure8.4.6.5.2.Market size and forecast, by Product8.4.6.5.3.Market size and forecast, by Method8.4.6.5.4.Market size and forecast, by end-user

    8.4.6.6.1.Market size and forecast, by Structure8.4.6.6.2.Market size and forecast, by Product8.4.6.6.3.Market size and forecast, by Method8.4.6.6.4.Market size and forecast, by end-user

    8.5.1.Key market trends, growth factors, and opportunities8.5.2.Market size and forecast, by Structure8.5.3.Market size and forecast, by Product8.5.4.Market size and forecast, by Method8.5.5.Market size and forecast, by end-user8.5.6.Market share analysis, by country

    8.5.6.1.1.Market size and forecast, by Structure8.5.6.1.2.Market size and forecast, by Product8.5.6.1.3.Market size and forecast, by Method8.5.6.1.4.Market size and forecast, by end-user

    8.5.6.2.1.Market size and forecast, by Structure8.5.6.2.2.Market size and forecast, by Product8.5.6.2.3.Market size and forecast, by Method8.5.6.2.4.Market size and forecast, by end-user

    8.5.6.3.1.Market size and forecast, by Structure8.5.6.3.2.Market size and forecast, by Product8.5.6.3.3.Market size and forecast, by Method8.5.6.3.4.Market size and forecast, by end-user

    8.5.6.4.1.Market size and forecast, by Structure8.5.6.4.2.Market size and forecast, by Product8.5.6.4.3.Market size and forecast, by Method8.5.6.4.4.Market size and forecast, by end-user

    CHAPTER 9:COMPETITIVE LANDSCAPE

    9.1.Introduction9.2.Product mapping of top 10 players9.3.competitive Heatmap9.4.Key development

    9.4.1.Business Expansion9.4.2.New Product9.4.3.Acquisition

    CHAPTER 10:COMPANY PROFILES

    10.1.1.company overview10.1.2.company snapshot10.1.3.Operating business segments10.1.4.Product portfolio10.1.5.Business performance10.1.6.Key strategic moves and developments

    10.2.1.company overview10.2.2.company snapshot10.2.3.Operating business segments10.2.4.Product portfolio10.2.5.Business performance10.2.6.Key strategic moves and developments

    10.3.1.company overview10.3.2.company snapshot10.3.3.Operating business segments10.3.4.Product portfolio10.3.5.Business performance10.3.6.Key strategic moves and developments

    10.4.Xiamen Weihua Solar Co. Ltd.

    10.4.1.company overview10.4.2.company snapshot10.4.3.Operating business segments10.4.4.Product portfolio10.4.5.Business performance10.4.6.Key strategic moves and developments

    10.5.1.company overview10.5.2.company snapshot10.5.3.Operating business segments10.5.4.Product portfolio10.5.5.Business performance10.5.6.Key strategic moves and developments

    10.6.1.company overview10.6.2.company snapshot10.6.3.Product portfolio10.6.4.Key strategic moves and developments

    10.7.1.company overview10.7.2.company snapshot10.7.3.Operating business segments10.7.4.Product portfolio10.7.5.Business performance10.7.6.Key strategic moves and developments

    10.8.1.company overview10.8.2.company snapshot10.8.3.Operating business segments10.8.4.Product portfolio10.8.5.Business performance

    10.9.1.company overview10.9.2.company snapshot10.9.3.Operating business segments10.9.4.Product portfolio10.9.5.Business performance10.9.6.Key strategic moves and developments

    10.10.New Energy Technologies Inc

    10.10.1.company overview10.10.2.company snapshot10.10.3.Operating business segments10.10.4.Product portfolio10.10.5.Business performance10.10.6.Key strategic moves and developments

    TABLE 01.GLOBAL PERVOSKITE SOLAR CELL, BY STRUCTURE, 2020–2030 (KILOTONS)TABLE 02.GLOBAL PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020–2030(MILLION)TABLE 03.PLANAR PEROVSKITE SOLAR CELLS PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (KILOTONS)TABLE 04.PLANAR PEROVSKITE SOLAR CELLS PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (MILLION)TABLE 05.MESOPOROUS PEROVSKITE SOLAR CELLS PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (KILOTONS)TABLE 06.MESOPOROUS PEROVSKITE SOLAR CELLS PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (MILLION)TABLE 07.GLOBAL PERVOSKITE SOLAR CELL, BY PRODUCT, 2020–2030 (KILOTONS)TABLE 08.GLOBAL PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020–2030(MILLION)TABLE 09.RIGID PEROVSKITE SOLAR CELLS PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (KILOTONS)TABLE 10.RIGID PEROVSKITE SOLAR CELLS PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (MILLION)TABLE 11.FLEXIBLE PEROVSKITE SOLAR CELLS PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (KILOTONS)TABLE 12.FLEXIBLE PEROVSKITE SOLAR CELLS PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (MILLION)TABLE 13.GLOBAL PERVOSKITE SOLAR CELL, BY METHOD, 2020–2030 (KILOTONS)TABLE 14.GLOBAL PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020–2030(MILLION)TABLE 15.SOLUTION METHOD PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (KILOTONS)TABLE 16.SOLUTION METHOD PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (MILLION)TABLE 17.VAPOR-DEPOSITION METHOD PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (KILOTONS)TABLE 18.VAPOR-DEPOSITION METHOD PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (MILLION)TABLE 19.VAPOR-ASSISTED SOLUTION METHOD PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (KILOTONS)TABLE 20.VAPOR-ASSISTED SOLUTION METHOD PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (MILLION)TABLE 21.GLOBAL PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020–2030 (KILOTONS)TABLE 22.GLOBAL PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020–2030 (MILLION)TABLE 23.PERVOSKITE SOLAR CELL MARKET, FOR AEROSPACE, BY REGION, 2020-2030 (KILOTONS)TABLE 24.PERVOSKITE SOLAR CELL MARKET,FOR AEROSPACE, BY REGION, 2020-2030 (MILLION)TABLE 25.PERVOSKITE SOLAR CELL MARKET, FOR INDUSTRIAL AUTOMATION, BY REGION, 2020-2030 (KILOTONS)TABLE 26.PERVOSKITE SOLAR CELL MARKET,FOR INDUSTRIAL AUTOMATION, BY REGION, 2020-2030 (MILLION)TABLE 27.PERVOSKITE SOLAR CELL MARKET, FOR CONSUMER ELECTRONICS, BY REGION, 2020-2030 (KILOTONS)TABLE 28.PERVOSKITE SOLAR CELL MARKET,FOR CONSUMER ELECTRONICS, BY REGION, 2020-2030 (MILLION)TABLE 29.PERVOSKITE SOLAR CELL MARKET, FOR ENERGY, BY REGION, 2020-2030 (KILOTONS)TABLE 30.PERVOSKITE SOLAR CELL MARKET,FOR ENERGY, BY REGION, 2020-2030 (MILLION)TABLE 31.PERVOSKITE SOLAR CELL MARKET, FOR OTHER END-USER, BY REGION, 2020-2030 (KILOTONS)TABLE 32.PERVOSKITE SOLAR CELL MARKET,FOR OTHERS, BY REGION, 2020-2030 (MILLION)TABLE 33.PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (KILOTONS)TABLE 34.PERVOSKITE SOLAR CELL MARKET, BY REGION, 2020-2030 (MILLION)TABLE 35.NORTH AMERICA PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 36.NORTH AMERICA PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 37.NORTH AMERICA PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 38.NORTH AMERICA PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 39.NORTH AMERICA PERVOSKITE SOLAR CELL MARKET, BY COUNTRY, 2020-2030 (MILLION)TABLE 40.U.S. PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 41.U.S. PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 42.U.S. PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 43.U.S. PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 44.CANADA PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 45.CANADA PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 46.CANADA PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 47.CANADA PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 48.MEXICO PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 49.MEXICO PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 50.MEXICO PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 51.MEXICO PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 52.EUROPE PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 53.EUROPE PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 54.EUROPE PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 55.EUROPE PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 56.EUROPE PERVOSKITE SOLAR CELL MARKET, BY COUNTRY, 2020-2030 (MILLION)TABLE 57.GERMANY PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 58.GERMANY PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 59.GERMANY PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 60.GERMANY PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 61.UK PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 62.UK PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 63.UK PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 64.UK PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 65.FRANCE PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 66.FRANCE PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 67.FRANCE PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 68.FRANCE PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 69.ITALY PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 70.ITALYPERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 71.ITALY PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 72.ITALY PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 73.SPAIN PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 74.ITALY PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 75.ITALY PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 76.ITALY PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 77.REST OF EUROPE PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 78.REST OF EUROPE PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 79.REST OF EUROPE PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 80.REST OF EUROPE PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 81.ASIA-PACIFIC PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 82.ASIA-PACIFIC PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 83.ASIA-PACIFIC PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 84.ASIA-PACIFIC PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 85.ASIA-PACIFIC PERVOSKITE SOLAR CELL MARKET, BY COUNTRY, 2020-2030 (MILLION)TABLE 86.CHINA PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 87.CHINA PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 88.CHINA PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 89.CHINA PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 90.JAPAN PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 91.JAPAN PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 92.NORTH AMERICA PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 93.CHINA PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 94.INDIA PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 95.INDIA PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 96.INDIA PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 97.INDIA PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 98.SOUTH KOREA PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 99.SOUTH KOREA PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 100.SOUTH KOREA PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 101.SOUTH KOREA PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 102.AUSTRALIA PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 103.AUSTRALIA PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 104.AUSTRALIA PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 105.AUSTRALIA PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 106.REST OF ASIA-PACIFIC PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 107.REST OF ASIA-PACIFIC PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 108.REST OF ASIA-PACIFIC PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 109.REST OF ASIA-PACIFIC PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 110.LAMEA PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 111.LAMEA PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 112.LAMEA PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 113.LAMEA PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 114.LAMEA PERVOSKITE SOLAR CELL MARKET, BY COUNTRY, 2020-2030 (MILLION)TABLE 115.BRAZIL PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 116.BRAZIL PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 117.BRAZIL PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 118.BRAZIL PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 119.SAUDI ARABIA PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 120.SAUDI ARABIA PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 121.SAUDI ARABIA PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 122.SAUDI ARABIA PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 123.SOUTH AFRICA PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 124.SOUTH AFRICA PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 125.SOUTH AFRICA PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 126.SOUTH AFRICA PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 127.REST OF LAMEA PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020-2030 (MILLION)TABLE 128.REST OF LAMEA PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020-2030 (MILLION)TABLE 129.REST OF LAMEA PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020-2030 (MILLION)TABLE 130.REST OF LAMEA PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020-2030 (MILLION)TABLE 131.KEY BUSINESS EXPANSIONTABLE 132.KEY NEW PRODUCTTABLE 133.KEY ACQUISITIONTABLE 134.ARCELORMITTAL S.A.: COMPANY SNAPSHOTTABLE 135.ARCELORMITTAL S.A.: OPERATING SEGMENTSTABLE 136.ARCELORMITTAL S.A.: PRODUCT PORTFOLIOTABLE 137.OVERALL FINANCIAL STATUS (MILLION)TABLE 138.ARCELORMITTAL S.A.: KEY STRATEGIC MOVES AND DEVELOPMENTSTABLE 139.BORAL LIMITED: COMPANY SNAPSHOTTABLE 140.BORAL LTD.: OPERATING SEGMENTSTABLE 141.BORAL LTD.: PRODUCT PORTFOLIOTABLE 142.OVERALL FINANCIAL STATUS (MILLION)TABLE 143.BORAL LTD.: KEY STRATEGIC MOVES AND DEVELOPMENTSTABLE 144.BASF SE: COMPANY SNAPSHOTTABLE 145.BASF SE: OPERATING SEGMENTSTABLE 146.BASF SE: PRODUCT PORTFOLIOTABLE 147.OVERALL FINANCIAL STATUS (MILLION)TABLE 148.CEMEX S.A.B. DE C.V.: COMPANY SNAPSHOTTABLE 149.CEMEX S.A.B. DE C.V.: OPERATING SEGMENTSTABLE 150.CEMEX S.A.B. DE C.V.: PRODUCT PORTFOLIOTABLE 151.OVERALL FINANCIAL STATUS (MILLION)TABLE 152.CEMEX S.A.B. DE C.V.: KEY STRATEGIC MOVES AND DEVELOPMENTSTABLE 153.CHARAH SOLUTIONS, INC.: COMPANY SNAPSHOTTABLE 154.BASF SE: OPERATING SEGMENTTABLE 155.BASF SE: PRODUCT PORTFOLIOTABLE 156.OVERALL FINANCIAL STATUS (MILLION)TABLE 157.CHARAH SOLUTIONS INC.: KEY STRATEGIC MOVES AND DEVELOPMENTSTABLE 158.CR MINERALS LLC: COMPANY SNAPSHOTTABLE 159.CR MINERALS LLC: PRODUCT PORTFOLIOTABLE 160.CR MINERALS LLC: KEY STRATEGIC MOVES AND DEVELOPMENTSTABLE 161.FERROGLOBE PLC: COMPANY SNAPSHOTTABLE 162.FERROGLOBE PLC.: OPERATING SEGMENTSTABLE 163.FERROGLOBE PLC: PRODUCT PORTFOLIOTABLE 164.OVERALL FINANCIAL STATUS (MILLION)TABLE 165.FERROGLOBE PLC: KEY STRATEGIC MOVES AND DEVELOPMENTSTABLE 166.LAFARGEHOLCIM: COMPANY SNAPSHOTTABLE 167.LAFARGEHOLCIM: OPERATING SEGMENTSTABLE 168.LAFARGEHOLCIM: PRODUCT PORTFOLIOTABLE 169.OVERALL FINANCIAL STATUS (MILLION)TABLE 170.LAFARGEHOLCIM: KEY STRATEGIC MOVES AND DEVELOPMENTSTABLE 171.SIKA AG: COMPANY SNAPSHOTTABLE 172.SIKA AG: OPERATING SEGMENTSTABLE 173.SIKA AG: PRODUCT PORTFOLIOTABLE 174.OVERALL FINANCIAL STATUS (MILLION)TABLE 175.SIKA AG: KEY STRATEGIC MOVES AND DEVELOPMENTSTABLE 176.TATA STEEL LTD.: COMPANY SNAPSHOTTABLE 177.TATA STEEL LTD.: OPERATING SEGMENTSTABLE 178.TATA STEEL LTD.: PRODUCT PORTFOLIOTABLE 179.OVERALL FINANCIAL STATUS (MILLION)TABLE 180.TATA STEEL LTD.: KEY STRATEGIC MOVES AND DEVELOPMENTS

    FIGURE 01.KEY MARKET SEGMENTSFIGURE 02.EXECUTIVE SUMMARY, BY SEGMENTFIGURE 03.EXECUTIVE SUMMARY, BY COUNTRYFIGURE 04.TOP INVESTMENT S, BY COUNTRYFIGURE 05.HIGH BARGAINING POWER OF SUPPLIERSFIGURE 06.MODERATE THREAT OF NEW ENTRANTSFIGURE 07.MODERATE THREAT OF SUBSTITUTESFIGURE 08.MODERATE INTENSITY OF RIVALRYFIGURE 09.HIGH BARGAINING POWER OF BUYERSFIGURE 10.TOP PLAYER POSITIONING, 2020FIGURE 11.PERVOSKITE SOLAR CELL DYNAMICSFIGURE 12.PERVOSKITE SOLAR CELL MARKET: VALUE CHAINFIGURE 13.GLOBAL PERVOSKITE SOLAR CELL MARKET, BY STRUCTURE, 2020–2030 (MILLION)FIGURE 14.COMPARATIVE SHARE ANALYSIS OF PLANAR PEROVSKITE SOLAR CELLS PERVOSKITE SOLAR CELL MARKET, BY COUNTRY, 2020 2030 (MILLION)FIGURE 15.COMPARATIVE SHARE ANALYSIS OF MESOPOROUS PEROVSKITE SOLAR CELLS PERVOSKITE SOLAR CELL MARKET, BY COUNTRY, 2020 2030 (MILLION)FIGURE 16.GLOBAL PERVOSKITE SOLAR CELL MARKET, BY PRODUCT, 2020–2030 (MILLION)FIGURE 17.COMPARATIVE SHARE ANALYSIS OF RIGID PEROVSKITE SOLAR CELLS PERVOSKITE SOLAR CELL MARKET, BY COUNTRY, 2020 2030 (MILLION)FIGURE 18.COMPARATIVE SHARE ANALYSIS OF FLEXIBLE PEROVSKITE SOLAR CELLS MARKET, BY COUNTRY, 2020 2030 (MILLION)FIGURE 19.GLOBAL PERVOSKITE SOLAR CELL MARKET, BY METHOD, 2020–2030 (MILLION)FIGURE 20.COMPARATIVE SHARE ANALYSIS OF SOLUTION METHOD PERVOSKITE SOLAR CELL MARKET, BY COUNTRY, 2020 2030 (MILLION)FIGURE 21.COMPARATIVE SHARE ANALYSIS OF VAPOR-DEPOSITION METHOD PERVOSKITE SOLAR CELL MARKET, BY COUNTRY, 2020 2030 (MILLION)FIGURE 22.COMPARATIVE SHARE ANALYSIS OF VAPOR-ASSISTED SOLUTION METHOD PERVOSKITE SOLAR CELL MARKET, BY COUNTRY, 2020 2030 (MILLION)FIGURE 23.GLOBAL PERVOSKITE SOLAR CELL MARKET, BY END-USER, 2020–2030 (MILLION)FIGURE 24.COMPARATIVE SHARE ANALYSIS OF PERVOSKITE SOLAR CELL MARKET, FOR AEROSPACE. BY COUNTRY, 2020 2030 (MILLION)FIGURE 25.COMPARATIVE SHARE ANALYSIS OF PERVOSKITE SOLAR CELL MARKET, FOR INDUSTRIAL AUTOMATION. BY COUNTRY, 2020 2030 (MILLION)FIGURE 26.COMPARATIVE SHARE ANALYSIS OF PERVOSKITE SOLAR CELL MARKET, FOR CONSUMER ELECTRONICS. BY COUNTRY, 2020 2030 (MILLION)FIGURE 27.COMPARATIVE SHARE ANALYSIS OF PERVOSKITE SOLAR CELL MARKET, FOR ENERGY, BY COUNTRY, 2020 2030 (MILLION)FIGURE 28.COMPARATIVE SHARE ANALYSIS OF PERVOSKITE SOLAR CELL MARKET, FOR OTHERS. BY COUNTRY, 2020 2030 (MILLION)FIGURE 29.U.S. PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 30.CANADA PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 31.MEXICO PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 32.GERMANY PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 33.UK PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 34.FRANCE PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 35.ITALY PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 36.SPAIN PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 37.REST OF EUROPE PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 38.CHINA PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 39.JAPAN PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 40.INDIA PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 41.SOUTH KOREA PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 42.AUSTRALIA PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 43.REST OF ASIA-PACIFIC PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 44.BRAZIL PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 45.SAUDI ARABIA PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 46.SOUTH AFRICA PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 47.REST OF LAMEA PERVOSKITE SOLAR CELL MARKET REVENUE, 2020-2030 (MILLION)FIGURE 48.PRODUCT MAPPING OF TOP 10 PLAYERS SOURCE: COMPANY WEBSITE,FIGURE 49.COMPETITIVE HEATMAPFIGURE 50.ARCELORMITTAL S.A.: NET SALES, 2018–2020 (MILLION)FIGURE 51.ARCELORMITTAL S.A.: REVENUE SHARE, BY SEGMENT, 2020 (%)FIGURE 52.BASF SE: REVENUE SHARE BY REGION, 2020 (%)FIGURE 53.BORAL LTD.: NET SALES, 2018–2020 (MILLION)FIGURE 54.BORAL LIMITED: REVENUE SHARE BY SEGMENT, 2020 (%)FIGURE 55.BASF SE: NET SALES, 2018–2020 (MILLION)FIGURE 56.BASF SE: REVENUE SHARE, BY SEGMENT, 2020 (%)FIGURE 57.BASF SE: REVENUE SHARE BY REGION, 2020 (%)FIGURE 58.CEMEX S.A.B. DE C.V.: NET SALES, 2018–2020 (MILLION)FIGURE 59.CEMEX S.A.B. DE C.V.: REVENUE SHARE, BY SEGMENT, 2020 (%)FIGURE 60.CEMEX S.A.B. DE C.V.: REVENUE SHARE BY REGION, 2020 (%)FIGURE 61.CHARAH SOLUTIONS, INC.: REVENUE SHARE BY REGION, 2020 (%)FIGURE 62.FERROGLOBE PLC: NET SALES, 2018–2020 (MILLION)FIGURE 63.FERROGLOBE PLC.: REVENUE SHARE, BY SEGMENT, 2020 (%)FIGURE 64.BASF SE: REVENUE SHARE BY REGION, 2020 (%)FIGURE 65.LAFARGEHOLCIM: NET SALES, 2018–2020 (MILLION)FIGURE 66.LAFARGEHOLCIM: REVENUE SHARE, BY SEGMENT, 2020 (%)FIGURE 67.LAFARGEHOLCIM: REVENUE SHARE BY REGION, 2020 (%)FIGURE 68.SIKA AG: NET SALES, 2018–2020 (MILLION)FIGURE 69.SIKA AG: REVENUE SHARE, BY REGION, 2020 (%)FIGURE 70.TATA STEEL LTD.: NET SALES, 2018–2020 (MILLION)FIGURE 71.ARKEMA S.A.: REVENUE SHARE, BY SEGMENT, 2020 (%)FIGURE 72.BASF SE: REVENUE SHARE BY REGION, 2020 (%)

    The Main Progress of Perovskite Solar Cells in 2020–2021

    Perovskite solar cells (PSCs) emerging as a promising photovoltaic technology with high efficiency and low manufacturing cost have attracted the attention from all over the world. Both the efficiency and stability of PSCs have increased steadily in recent years, and the research on reducing lead leakage and developing eco-friendly lead-free perovskites pushes forward the commercialization of PSCs step by step. This review summarizes the main progress of PSCs in 2020 and 2021 from the aspects of efficiency, stability, perovskite-based tandem devices, and lead-free PSCs. over, a brief discussion on the development of PSC modules and its challenges toward practical application is provided.

    Introduction

    Perovskite solar cells (PSCs) have become a promising thin-film photovoltaic (PV) technology due to the high light-absorption coefficient, long carrier diffusion length, and solution processibility of metal halide perovskite materials [1,2,3,4,5]. Currently, the highest power conversion efficiency (PCE) of PSCs has reached 25.5% [6], exceeding the record efficiency of copper indium gallium selenium (CIGS) solar cells and approaching that of crystalline-Si solar cells. over, the scale-up deposition techniques for perovskites and charge transport layers promote the development of large-area perovskite solar modules, including doctor-blade coating, slot-die coating, screen printing, and spray deposition strategies. Recently, a certified efficiency of about 18% has been reported for an over 800 cm 2 PSC sub-module [6], indicating large potential for practical use. On the other hand, perovskite-based tandem solar cells with a theoretical PCE beyond the limit of single-junction PSCs also make a huge progress owing to the reduction of defect density and increase of carrier diffusion length in both wide-bandgap and narrow-bandgap perovskite absorbers [7,8,9]. By now, the highest certified efficiency of the perovskite-based tandems has increased to over 29% [10].

    Besides the efficiency progress, the long-term stability of PSCs against damp, light, and heat also improved significantly in recent years, which could be attributed to the construction of diffusion barrier against ion migration, additive engineering, design of chemically inert carbon-based electrodes, and development of cell encapsulation technique to reduce the lead leakage from a broken PSC module [11,12,13,14,15,16,17,18]. It was demonstrated that the printable PSCs had passed the most popular international standards of IEC61215:2016 for mature PV technology [19].

    With the continuous progress of PSCs toward commercialization, exploiting eco-friendly lead-free perovskite materials has also become a hot research topic in this field in view of the toxicity of Pb element in Pb-containing PSCs giving rise to the concern of environmental pollution [20,21,22,23]. So far, the highest certified efficiency of lead-free PSCs has reached 11.22%, enabled by minimizing the defect density in tin halide perovskite films via a template-growth deposition method [24].

    In this review, we summarize the representative works on PSCs published by worldwide research groups in 2020–2021 from the aspects of efficiency, stability, perovskite-based tandem solar cells, and the development of lead-free PSCs. In addition, we point out the future challenges on realizing the commercialization of PSCs, and then give a brief outlook on the possible research topics at the next stage.

    Efficiency

    PSCs are usually composed of perovskite absorbers, charge transport layers and counter electrodes. The energy loss in the bulk and interface of perovskites layers, and the charge extraction and transportation process in device play a critical role in determining the efficiency of PSCs. Therefore, improving the crystal quality of perovskite films, suppressing the non-radiative recombination at interface, and rational design of the charge transport layers are expected to further improve the device efficiency.

    2.1 Perovskite Absorbers

    In 2020–2021, the research activities on perovskite layer mainly FOCUS on stabilizing the formamidinium lead iodide (FAPbI3) perovskite phase with wide absorption range and long carrier lifetime to boost the short-circuit current density (JSC) and open-circuit voltage (VOC) of PSCs [25, 26]. Besides, exploiting the single-crystal device with very low defect density has also been reported for realizing high-performance PSCs.

    To maximize the photon-absorption at the UV–vis region, Kim et al. used the inherent bandgap (1.47 eV) of α-phase FAPbI3 to increase the photocurrent of PSCs [27]. They replaced the FA cations by a trace amount (0.03 molar fraction) of cesium (Cs) and methylenediammonium (MDA) cations to stabilize the α-phase FAPbI3 without changing its inherent bandgap, but improved the UV–vis absorption of perovskite layer. These effects contribute to a high JSC of 26.23 mA cm −2 for PSCs, very close to the theoretical current limit of about 27 mA cm −2 for 1.47 eV-bandgap semiconductors [7]. over, Fig. 1a shows that co-doping of 0.03 mol Cs and MDA cations also relaxes the lattice strain of FAPbI3 by over 70% compared to the sample only treated with MDA cation reported in their previous study [28], which prolongs the carrier recombination lifetime and enables a VOC increase of 30 mV (Fig. 1b), leading to a high certified PCE reaching 24.4%.

    solar, cell

    Besides the cation-doping strategies, substitution of X-site halide anions could also significantly affect the optoelectronic properties of FAPbI3 perovskites. Jeong et al. introduced an anion engineering strategy that employs the pseudo-halide anion formate (HCOO − ) to fill the halide vacancy defects located at grain boundaries and surface of perovskite films [29] and to enhance the crystallinity of FAPbI3. It is found that the doping of 2% formate anions could enlarge the grain size to about 2 μm, increasing the crystal orientation along (100) and (200) directions that are better for carrier transport, and suppressing the formation of non-photoactive δ-FAPbI3 phase. over, the theoretical calculation revealed that formate anions had a larger binding affinity toward iodide vacancy sites compared to other anions like Cl −. Br −. and BF4 − owing to the fact that every carboxylate group can form two Pb–O coordination bonds with the lead cations (Fig. 1c, d). As a result, the FAPbI3-based PSCs with pseudo-halide treatment attained a record PCE of 25.6% (certified 25.2%) and a VOC of 1.19 V.

    Another representative work for the improvement in perovskite layers is the fabrication of uniaxial-oriented perovskite layer with millimeter-sized grains via a methylamine gas-assisted crystallization method [30]. In this study, the methylammonium lead iodide (MAPbI3) perovskite film was formed by controlling the evaporation rate of MA gas molecules from liquid intermediate phase of MAPbI3·xMA in a closed system. The results demonstrated that a slow release rate of MA molecules from the liquid intermediate phase significantly reduced the supersaturation, leading to a low nucleation density that provided more time for growing large MAPbI3 grains during thermal annealing process. Consequently, the perovskite grains with about 1 mm-size can be obtained, resulting in a very low trap density of 9.7 × 10 13 cm −3. approaching that of the MAPbI3 single crystal [31], and the PSCs with millimeter-sized MAPbI3 perovskite grains exhibited a promising PCE of 21.36%.

    2.2 Interface Engineering

    The interfacial properties between perovskite and charge transport layers play an important role in the charge recombination mechanism of PSCs, which greatly influence the VOC and fill factor (FF) of the solar cells [32,33,34]. Therefore, many methods have been developed to reduce the non-radiative recombination loss at the interface and optimize the series resistance of the passivation layers for efficiency improvement.

    To suppress carrier recombination at perovskite surface and grain boundaries, Zheng et al. used a trace amount of surface-anchoring alkylamine ligands (AALs) with different chain lengths for the modification of perovskite surface [35]. They found that the oleylamine ligand with a long alkyl chain showed a more obvious passivation effect for improving carrier recombination lifetime and the VOC of PSCs (Fig. 2a), as compared to other organic AALs with shorter alkyl chains, which could be attributed to the electron-tunneling effect in the insulating AALs layers. Such a tunneling effect allows the movement of electrons from perovskite conduction Band to the lowest unoccupied molecular orbital (LUMO) of the fullerene-C60 layer, but effectively blocks the hole injection from valence Band to the highest occupied molecular orbital (LUMO) of C60 (Fig. 2b). This passivation technique based on the insulating tunneling layer is also widely applied in silicon solar cells [36]. Based on this oleylamine ligand-anchoring strategy, the inverted-structure PSCs achieved a strong VOC improvement up to 110 mV, resulting in a record certified efficiency of 22.3%.

    Although the additional passivation layer is beneficial for the VOC improvement of PSCs, it may undesirably increase the series resistance of devices and thus leads to the loss of FF. This detrimental effect will become more obvious in the case of large-area devices. In order to balance the improvement of VOC and FF, Peng et al. designed a nanoscale localized contact at the polymer-passivated perovskite/TiO2 interface to realize a higher FF for 1 cm 2 PSCs [37]. They deposited a nanopatterned electron-selective TiO2 layer via atomic-layer deposition (ALD) to form nanorod-like charge transport channels through the passivated interface (the device structure is shown in Fig. 2c), which provided both effective contact passivation and excellent charge extraction ability for reducing the device series resistance. As a result, they achieved a high FF of 83.9% and a high VOC of 1.20 V by employing an ALD TiO2 nanopattern with 300 nm spacing at the polymer passivation interface (Fig. 2d), resulting in a record certified PCE of 21.6% for an over 1 cm 2 cell.

    2.3 Charge Transport Layer Design

    Rational design of the charge transport layer also boosts the PV parameters of PSCs significantly. Jeong et al. designed two fluorinated isomeric analogs of the well-known hole transport molecule spiro-OMeTAD to modify the energy-level alignment, hydrophobicity, and hole extraction ability in PSCs [38]. They found that the fluorinated group at the meta-position on the benzene ring (spiro-mF) could lower the HOMO position from −4.97 to −5.19 eV and thus provided a more suitable energy-level alignment with the valence Band maximum (VBM) of FAPbI3 perovskite absorber (-5.40 eV) to reduce the interfacial energy loss, resulting in a VOC improvement of the device. over, the F atoms could induce a denser solid-state molecule packing through non-covalent intramolecular interaction, which further improves the electronic contact between spiro-mF and perovskite surface, producing better transport and extraction ability of holes, resulting in a slightly increase of JSC and FF. These effects boost the efficiency from 23.44% to 24.84% for PSCs.

    On the other side, chemical bath deposition was applied to produce a high-quality electron-selective SnO2 layer for limiting excess charge carrier recombination in PSCs [39]. Unlike the conventional deposition using SnOx nanoparticle dispersion, chemical bath deposition enables the uniform and complete coverage of SnO2 film on the substrate via a rational control of the reaction time and pH value of the SnCl2 precursor solution. As can be seen in Fig. 3, at the stage A-i, with a low pH value and short reaction time, the as-deposited SnO2 film showed many pinholes. After increasing the pH value to stage A-ii, a low O-vacancy SnO2 film (targeted sample) with ideal film coverage, thickness, and chemical composition could be achieved. When further increasing the pH and reaction time to stages A-iii and B, the density of O-vacancy increased significantly and the additional Sn6O4(OH)4 and SnO phases were present, lowering the electron transport ability and inducing charge recombination in SnO2 films. By further optimizing the Br concentration in perovskite films, a certified PCE of 25.2% was achieved, corresponding to 80.5% of the theoretical efficiency limit.

    Stability

    Stability issues are the major bottleneck for the commercialization of PSCs. The chemical components of metal halide perovskite are bonded through weak interactions such as ionic interaction, hydrogen bonding, and van der Waals forces, which results in the soft-material nature of perovskite semiconductors. Irreversible decomposition of organic species and the ion migration occur easily in PSCs under moisture penetration, continuous light soaking, thermal stress, and external electric field [11, 40,41,42], which cause damage to both the perovskite and charge transport layers. On the other hand, the leakage of Pb during long-term operation of PSCs under bad weather conditions could further cause environmental and public health risk, which should also be considered as an important stability issue to be addressed for future commercialization. In this part, we discuss the recent research progress on PSC stability from the viewpoint of additive engineering, heterostructure stabilization, and the cell encapsulation technology for reducing lead leakage.

    3.1 Additive Engineering

    Introducing the additive molecules that could form extra chemical interactions with the volatile components is an effective way to suppress ion migration and irreversible decomposition of perovskite films. Mei et al. [19] introduced bifunctional 5-ammoniumvaleric acid iodide (5-AVAI) to inhibit the methylammonium iodide (MAI) loss on the surface and the crystal reconstruction of MAPbI3 grains. 5-AVAI could form hydrogen bonds with the iodine ions through amino groups (−NH3 ), while their carboxyl groups (-COOH) form a strong hydrogen bond with another 5-AVAI in the adjacent grains, which results in a cross-linking of the perovskite surface and interface and a large enhancement of the thermal stability. Besides, non-volatile 5-AVAI at the interfaces limits the volatilization of organic components, which inhibits the forward process of decomposition reaction of MAPbI3. It can also stabilize the metal oxide matrix (ZrO2 and TiO2) via the anchoring effect between.COOH and metal cations, thus restricting ion migration of MA and I − at the grain boundaries and the interfaces. After encapsulation with a hot-melt polymer film, the printable PSCs treated with 5-AVAI successfully passed the main items of IEC61215:2016 qualification tests, especially working for more than 9000 h at a maximum power point of 55 ± 5 °C without obvious decay.

    Furthermore, an organic ionic salt called 1-butyl-1-methylpiperidinium tetrafluoroborate ([BMP] [BF4] − ) was developed to suppress the light-induced phase segregation and degradation in thermally stable CsFA-based lead-halide perovskites [43]. The results indicated that small amounts of [BMP] [BF4] − additives could penetrate the entire volume of perovskite film and passivate both the bulk and surface defect states via ionic interaction, leading to a large photovoltage enhancement for the PSCs. During the aging process, [BMP] [BF4] − prevented the segregation of Br-rich FAPbBr3 phase in mixed-halide perovskite film and suppress the corrosion of silver electrode (one of the key reasons for device degradation) induced by I2 generation under light and heat. I2 is mainly formed by the combination of interstitial I − and holes, or two neutral iodine atoms [44]. In the presence of [BMP] [BF4] −. the ion diffusion channels at the grain surface were effectively reduced, resulting in the inhibition of I2 formation. Consequently, under full-spectrum simulated sunlight in ambient atmosphere, the unencapsulated and encapsulated PSCs retained 80% and 95% of their post-burn-in efficiencies for 1010 h (60 °C) and 1200 h (85 °C), respectively.

    Other additives such as lead chloride, organic dye molecules, phosphorus-based Lewis acid, and cyano derivatives have been reported to suppress ion migration and passivate surface ionic defects in perovskite film, resulting in the improvement of long-term stability [45,46,47].

    3.2 Heterostructure Stabilization

    The heterostructures formed between perovskite and charge transport layers play a significant role in the long-term stability of the whole device, so much attention has been paid to stabilizing the interface of each functional layer in PSCs. Liu et al. employed a holistic strategy to stabilize the front and back interface of perovskite layer under operational condition [48].They used ethylenediaminetetraacetic acid dipotassium salt (EDTAK) to treat the electron-selective SnO2 surface for stabilization of the front heterostructure (Fig. 4a). The Lewis acid–base reaction between the alkylamine group of EDTAK and Pb 2 cations efficiently passivated the vacancy defects in the perovskite film. Besides, incorporation of EDTAK also shifted the conduction Band minimum (CBM) of SnO2 from −3.69 to −3.95 eV, resulting in a better energy-level alignment for charge extraction. The perovskite/spiro-OMeTAD heterostructure was stabilized by the surface ethylammonium iodide (EAI) modification to form an EAMA-based perovskite capping layer with better ambient stability. Regarding the spiro-OMeTAD/electrode interface, a small amount of P3HT was incorporated to inhibit the degradation caused by inward migration of gold into the perovskite layer and also enhance the moisture stability of spiro-OMeTAD. By employing such a holistic stabilization approach, the PSC modules without encapsulation achieved an efficiency of 16.6% with a designated area of 22.4 cm 2. and the encapsulated solar modules with parylene-coated cover glass retained approximately 86% of the initial performance after continuous operation for 2000 h under AM1.5G light illumination (Fig. 4b).

    Our previous work employed CsFA-based perovskite with high thermal stability as the absorber to construct a scalable integrated heterostructure for PSC modules [49]. Under the operational condition, the iodide ions can easily migrate from perovskite to the hole transport layer (HTL) and change its semiconducting properties from p‐type to n‐type, causing deterioration of hole extraction in PSCs. To stabilize this heterostructure, a scalable bridge‐jointed graphene oxide (BJ-GO) layer was deposited on the perovskite surface to block iodide migration. During this process, 3‐aminopropyl triethoxysilane (APTES) was used to connect the small-size GO nanosheets by forming C-N covalent bonds with GO, resulting in a full‐coverage hydrophobic blocking layer on perovskite surface for inhibiting ion migration, passivating surface under-coordinated Pb 2. and protecting perovskite layer from the damage of water in ambient environment. By employing the BJ-GO layer in the PSCs with undoped HTL, we achieved a high PCE of 16.21% for a 36 cm 2 solar module. The encapsulated module retained over 91% of its initial efficiency after the damp heat test at 85 °C and 85% relative humidity for 1000 h, while maintaining 90% of the initial value for 1000 h under standard operational condition at 60 °C.

    3.3 Cell Encapsulation

    In 2020–2021, researches on PSC encapsulation mainly focused on reducing the Pb leakage from a broken device. During the long-term operation process, Pb 2 showing weak chemical interaction with other perovskite components could also escape from perovskite layer after water penetration, resulting in additional environmental problems. For a typical lead perovskite absorber with a thickness of 550 nm, the unit-area lead concentration is estimated to be about 0.75 g m −2. which is more than 100 times higher than that of the commonly used Pb-containing paints (0.007 g m −2 ) [50].

    The physical encapsulation with Pb 2.absorption materials can effectively suppress the irreversible outflow of Pb from damaged devices into the underground water or soils at severe weather conditions [51]. Li et al. deposited the lead-absorbing materials on both front and back sides of PSCs to reduce the leakage of lead [23]. On the front glass side, they used a transparent Pb-absorbing molecule with phosphonic acid groups that showed a large binding energy with Pb 2 cations to absorb lead in water when water seeps into the device. On the back side, they coated a polymer film mixed with lead-coordination agents between the metal electrode and the encapsulation layer. As a consequence, the lead-absorption effects on both sides of PSCs could sequestrate more than 96% of lead leakage in water induced by severe device damage while retain the structural integrity of solar cells.

    Furthermore, a low-cost and chemically robust cation-exchange resin (CER)-based method was developed to prevent lead leakage from broken PSC modules [52]. The results showed that CER exhibited both high adsorption capacity and high adsorption rate of Pb 2 in water due to the high binding energy between sulfonate-terminated groups and the common divalent metal ions like Pb 2. Ca 2. and Mg 2 in the mesoporous polymer matrix (Fig. 5a). Additionally, mixing the CER with carbon electrode and integrating them on the front glass of PSC modules have a negligible detrimental impact on the device PCE while reduces lead leakage from mini-modules in water by 62-fold to 14.3 ppb as compared to that of the device without CER (Fig. 5b). The theoretical calculation indicates that the CER treatment could further reduce the lead leakage of large-area perovskite solar panels to lower than 7.0 ppb even under the worst condition that all the PSC sub-modules are broken.

    Perovskite-Based Tandem Solar Cells

    The tunable bandgap of ABX3 perovskite absorber from 1.2 eV to 3.0 eV benefits the design of silicon-perovskite or perovskite-perovskite tandem devices [34] that enables the theoretical efficiency beyond the Shockley–Queisser limit of single-junction solar cells. For silicon-perovskite tandems, exploiting the ideal wide-bandgap perovskite materials (1.6–1.7 eV) with suitable spectrum response matching the 1.1 eV-bandgap silicon absorber is key to obtaining high efficiency [53]. By contrast, suppressing the generation of Sn 4 defects and increasing the carrier diffusion length in narrow-bandgap mixed Sn–Pb perovskites (1.1–1.2 eV) are currently the major tasks for efficient all-perovskite tandem devices.

    4.1 Silicon-Perovskite Tandem Structure

    In the past two years, the efficiency of silicon-perovskite tandems boosted rapidly from about 25% to over 29% due to the reduced VOC deficit of wide-bandgap perovskite top cell enabled by compositional engineering for a stable phase and the selection of charge transport layers [54,55,56]. Xu et al. [57] reported a triple-halide alloys containing chlorine, bromine, and iodine to tailor the carrier lifetime and suppress the light-induced phase segregation in wide-bandgap perovskite films. They found that direct incorporation of large amounts of Cl (more than 15%) into the double-halide perovskite could produce a uniform halide distribution throughout the film with an ideal bandgap (1.67 eV) matching the spectrum of the bottom 1.12-eV Si absorber. This effect increased the photocarrier mobility by two times and efficiently suppressed the phase separation that was previously found in most of the mixed-halide perovskite compositions [58, 59]. By extending the double-halide to triple-halide components, a distinct efficiency enhancement for opaque single-junction PSCs from 18.15% to 20.42% was demonstrated, with a VOC increasing over 100 mV. over, they obtained a PCE of 27.04% in 1-cm 2 two-terminal monolithic tandems via integrating the optimized perovskite top cells with the silicon bottom cells.

    Besides the modification of perovskite layer, a novel self-assembled, methyl-substituted carbazole monolayer (Me-4PACz, the chemical structure is shown in Fig. 6a) was developed as the HTL to minimize the non-radiative recombination loss at the hole-selective contact in wide-bandgap PSCs [10], resulting in a record efficiency of 29.15% for Si-perovskite tandem devices. The carbazole unit provided fast hole extraction from the wide-bandgap perovskite layer and the methyl substitution was considered to passivate the defects on perovskite surface, minimizing the FF loss induced by series resistance (transport loss) and the interfacial non-radiative recombination in PSCs (Fig. 6b), as investigated by intensity-dependent absolute photoluminescence measurements. By using the Me-4PACz hole-selective layer, the FF of single-junction PSCs greatly increased from 79.8% (control cell based on PTAA) to 84.0%, and the device VOC also showed a large improvement from 1.19 V to 1.25 V.

    Textured crystalline-Si subcells have also been reported for Si-perovskite tandem devices to reduce the device reflectance and increase the photon usage rate. Hou et al. [60] directly deposited the solution-processed micrometer-thick perovskite top cell on a fully textured Si-heterojunction bottom cell to fabricate tandem devices. They addressed the carrier-extraction issue in micrometer-thick perovskite layer by increasing the depletion width at the bases of silicon bottom cell and employing a self-assembled 1-butanethiol passivation layer on the perovskite surface. This strategy not only increased the carrier diffusion length but also stabilized the wide-bandgap perovskite phase, enabling a certified efficiency of 25.7%. Similarly, Chen et al. used a nitrogen-assisted blading process to deposit hole transport layer and high-quality planarizing perovskite absorber that fully covered the rough silicon pyramids [61]. over, a textured light-scattering layer was added to the perovskite top cell to reduce reflectance at the front surface, leading to an efficiency of 26.0% for textured Si-perovskite tandem devices.

    4.2 All-Perovskite Tandem Structure

    Compared to the Si-perovskite tandems, all-perovskite tandems exhibit additional advantages of low materials and fabrication costs because the bottom and top cells can be fabricated by the same preparation process without using specific equipment for other types of solar cells. The initial two-terminal all-perovskite tandems used MAPbBr3 and MAPbI3 as the wide-bandgap (2.30 eV) and narrow-bandgap (1.55 eV) absorbers, respectively, giving an efficiency of 10.8% [62].

    After a few years of development, the efficiency of all-perovskite tandems has increased to over 24% via tailoring the bandgap alignment in devices. The mixed Sn–Pb perovskites with a minimized bandgap of about 1.2 eV are a promising candidate as the narrow-bandgap absorbers [63]. However, the spontaneous oxidation of Sn 2 to Sn 4 and Sn vacancy cause a high defect density in Sn–Pb perovskites and thus a shorter carrier diffusion length compared to that of the full-Pb counterpart, which lowers the charge transport efficiency in the thick narrow-bandgap absorber (over 1 μm). To overcome this challenge, Lin et al. used metallic Sn to reduce the Sn 4 impurities in perovskite precursors via a comproportionation reaction (Sn Sn 4 → Sn 2 ) [64], reducing the defect density from 1.4 × 10 16 to 5.4 × 10 15 cm −3 and thereby prolonging the diffusion length from 0.75 to 2.99 μm in MA0.3FA0.7Pb0.5Sn0.5I3 perovskites. After integrating this narrow-bandgap Sn–Pb PSCs (1.22 eV) with the wide-bandgap PSCs (1.77 eV), they obtained a high PCE of 24.8% for the tandem solar cells.

    To further reduce the defect density in mixed Sn–Pb perovskite, Lin et al. employed a strongly reductive surface-anchoring molecules, formamidine sulfinic acid (FSA), to reduce the Sn 4 defects and passivate the surface ionic vacancy [9]. As shown in Fig. 6c, the sulfinic acid group could react with oxygen molecules and thus suppress the Sn 2 oxidation. Meanwhile, the O atoms of sulfinic group serving as an electron donor could form coordination bond with the under-coordinated Sn 2 or Pb 2 cations to passivate the surface halide vacancy. On the other hand, the formamidine group of FSA showing similar structure with FA cation could also passivate the surface A-site vacancy defects. Such synergistic effect of the surface-anchoring FSA improved the carrier recombination lifetime in Sn–Pb perovskites by three times. Via employing this narrow-bandgap absorber into the monolithic tandem structure (Fig. 6d), they obtained a VOC enhancement of about 60 mV and a large PCE improvement from 22.7% to 24.7% for a 1.05-cm 2 all-perovskite tandem (Fig. 6e).

    Furthermore, the solution-processed all-perovskite triple-junction tandem device with a PCE of over 20% was also reported [65]. This triple-junction solar cell contained a 1.99−eV perovskite at the front cell, a 1.60−eV perovskite at the middle cell, and a 1.22−eV perovskite at the back cell (Fig. 6f). By developing compatible interconnecting layers with those solution-processed perovskite absorbers, a high VOC of 2.8 V can be achieved, much higher than the value of all-perovskite double-junction solar cells.

    4.3 Lead-Free PSCs

    With the continuous movement of PSCs to commercialization, the toxicity of Pb element in perovskite absorber arouses the concern of environmental problems [66]. In recent years, a growing number of studies have aimed at developing the eco-friendly lead-free PSCs to directly avoid the use of lead in metal halide perovskite layers. So far, a number of lead-free perovskites based on tin (Sn), antimony (Sb), bismuth (Bi), titanium (Ti), germanium (Ge), and copper (Cu) have been exploited for the application of solar cells [67,68,69,70,71].

    Among all the lead-free perovskite materials, the most promising candidate is tin perovskite. Sn has a similar outer electronic structure (ns 2 np 2 ) and ionic radius to Pb, enabling the complete replacement of lead in perovskite lattice without causing notable phase segregation. Furthermore, tin perovskites demonstrate some additional advantages: (1) ideal bandgap close to Shockley–Queisser limit (1.3 ~ 1.4 eV), (2) low exciton binding energy (29 meV for MASnI3 and 62 meV for MAPbI3), and (3) high charge carrier mobility (μe (electron mobility) = 2000 cm 2 V −1 s −1 for MASnI3 and 60 cm 2 V −1 s −1 for MAPbI3) [72, 73].

    After a few years of development, the efficiency of tin PSCs underwent a Rapid growth from 6 to 11%–13% owing to the careful design of antioxidant additives, mediating the crystallization rate, design of suitable A-site cations such as phenethylammonium (PEA) and pentafluorophen-oxyethylammonium (FOE), and constructing the oriented low-dimensional perovskite structure [22, 74, 75]. Currently, researchers mainly FOCUS on the defect passivation and selection of suitable electron transport layers to minimize the VOC loss in tin PSCs.

    Nishimura et al. combined the bulk ethylammonium (EA) doping and surface edamine passivation to reduce the defect density of FASnI3 perovskite films by as much as 1 order of magnitude [21], which resulted in a high device VOC of 0.84 V. Furthermore, our group developed a template-growth technique assisted by an n-propylammonium iodide (PAI) salt to reduce the bulk defect density in the solution-processed FASnI3 films [24]. The PAI post-treatment increased the electron diffusion length from 70 to 180 nm and thereby reduced the recombination loss in FASnI3 absorbers due to the increased crystal orientation along (h00) directions (Fig. 7a), which enabled a significant VOC enhancement of 200 mV and a record certified PCE of 11.22% for tin PSCs (Fig. 7b). On the other side, Jiang et al. utilized indene-C60 bis-adduct (ICBA) with a high LUMO position as the electron transport layer to reduce the energy-level offset in tin PSCs [76]. They found that a shallower LUMO of −3.74 eV in ICBA than that (−3.91 eV) of traditional [6, 6]-phenyl-C61-butyric acid methyl ester (PCBM) can upshift the quasi-Fermi level of electrons, approaching the CBM (−3.69 eV) of FA0.85PEA0.15SnI3 perovskite (PEA represents cation) and thus improving the quasi-Fermi level splitting and the maximum attainable VOC. Consequently, the ICBA-based tin PSCs obtained an ultra-high open-circuit voltage of 0.94 V, about 300 mV increasing compared to that of the PCBM-based cell, which enables a high PCE of 12.4%. The detailed PV parameters of tin PSCs reported in 2020–2021 are summarized in Table 1.

    Furthermore, recent study showed that organic solvents used in the mass production of PSC modules such as N,N-dimethylformamide (DMF), N,N-dimethylacetamide (DMAC), N-methyl-2-pyrrolidone (NMP), and gamma-butyrolactone (GBL) are toxic to the human reproductive systems [80]. Therefore, development of green solvent systems or the solvent-free deposition technology for fabricating large-area perovskite film will be an important research topic in the future [81, 82].

    Besides the efficiency, more and more attention has been paid to the long-term stability of PSC modules. In 2020, Microquanta company announced that its mass-produced PSC module has passed the strict stability test according to the International Electro technical Commission (IEC) standards, the 20 cm 2 perovskite module underwent 3000 h damp heat test without degradation, and showed an efficiency loss less than 2% after UV preconditioning test, associated with a product lifetime of over 25 years. Also, Utmo Light Ltd. reported that their PSC mini-module passed the IEC test with a stabilized efficiency of over 20%, which is a milestone for PSCs toward the practical use. Recently, Okinawa Institute of Science and Technology Graduate University (OIST, Japan) reported over 1100-h operational lifetime for a 10 × 10 cm 2 solar module [83]. Although many research groups and companies claimed that their devices have passed IEC standard test, there are still some stability issues needed to be addressed at the next stage, one important thing is the cell encapsulation technology. A well-designed encapsulation of PSCs should not only block the outside moisture and oxygen effectively, but also suppress the escape of volatile products and lead from the decomposed organic hybrid perovskite layer [84]. In this regard, we believe that a growing number of studies will move to exploit such multifunctional encapsulation materials in the future.

    So far, a series of stability tests custom-made for PSCs have been proposed [85], and we therefore suggest a standardized stability test for PSCs and encourage the researchers to measure the module stability at an authorized third-party test center to increase the credibility of their results.

    Development of highly efficient lead-free PSCs is also an alternative choice to extend their application range in the PV markets, especially for the indoor power generation like wearable power sources that have a strict limit on lead content [86,87,88]. It was demonstrated that tin PSCs could be the next generation of PSCs for realizing over 20% efficiency and the strategies for their mass production have been investigated [20]. over, developing reducing solvent system for Sn 2 precursor, constructing antioxidant capping layer during the crystallization of tin perovskite film, and the investigation of compatible device encapsulation approaches also led to a large stability improvement of tin PSCs [89, 90]. In our opinion, the application of tin PSCs for indoor PV products could be prior to lead PSCs when their efficiencies reach over 15%.

    Acknowledgements

    This work was supported by the National Natural Science Foundation of China (Grant Nos. 11834011 and 12074245). Y. B. Q. acknowledges the support from the Energy Materials and Surface Sciences Unit of the Okinawa Institute of Science and Technology Graduate University.

    Author information

    Authors and Affiliations

    • State Key Laboratory of Metal Matrix Composites, School of Material Science and Engineering, Shanghai Jiao Tong University, Shanghai, 200240, People’s Republic of China Tianhao Wu, Zhenzhen Qin, Yanbo Wang Liyuan Han
    • Key Laboratory for Advanced Materials and Feringa Nobel Prize Scientist Joint Research Centre, Shanghai Key Laboratory of Functional Materials Chemistry, Joint International Research Laboratory of Precision Chemistry and Molecular Engineering, School of Chemistry and Molecular Engineering, East China University of Science and Technology, Meilong Road 130, Shanghai, 200237, People’s Republic of China Yongzhen Wu
    • Wuhan National Laboratory for Optoelectronics, Huazhong University of Science and Technology, Luoyu Road 1037, Wuhan, 430074, People’s Republic of China Wei Chen
    • College of Materials Science and Engineering, Nanjing University of Science and Technology, Nanjing, 210094, People’s Republic of China Shufang Zhang
    • Beijing Key Laboratory of Novel Thin-Film Solar Cells and State Key Laboratory of Alternate Electrical Power System With Renewable Energy Sources, North China Electric Power University, Beijing, 102206, People’s Republic of China Molang Cai Songyuan Dai
    • Department of Microelectronic Science and Engineering, Ningbo University, Zhejiang, 315211, People’s Republic of China Jing Zhang
    • College of Chemical Engineering, Jiangsu Key Lab of Biomass-Based Green Fuels and Chemicals, Nanjing Forestry University, Nanjing, 210037, People’s Republic of China Jian Liu
    • College of Chemistry and Molecular Engineering, Qingdao University of Science and Technology, Qingdao, 266042, People’s Republic of China Zhongmin Zhou
    • Special Division of Environmental and Energy Science, Komaba Organization for Educational Excellence (KOMEX), College of Arts and Sciences, University of Tokyo, Tokyo, 153-8902, Japan Xiao Liu, Hiroshi Segawa Liyuan Han
    • National Laboratory of Solid State Microstructures, Jiangsu Key Laboratory of Artificial Functional Materials, College of Engineering and Applied Sciences, Nanjing University, Nanjing, 210093, People’s Republic of China Hairen Tan
    • College of Information Science and Technology, Jinan University, Guangzhou, 510632, People’s Republic of China Qunwei Tang
    • School of Physics and Electronic Science, Engineering Research Center of Nanophotonics and Advanced Instrument, Ministry of Education, East China Normal University, Shanghai, 200062, People’s Republic of China Junfeng Fang
    • Laboratory of Advanced Optoelectronic Materials, College of Chemistry, Chemical Engineering and Materials Science, Soochow University, Suzhou, 215123, People’s Republic of China Yaowen Li
    • Center for Excellence in Nanoscience, Key Laboratory of Nanosystem and Hierarchical Fabrication, National Center for Nanoscience and Technology, Beijing, 100190, People’s Republic of China Liming Ding
    • School of Physical Science and Technology, ShanghaiTech University, 100 Haike Road, Shanghai, 201210, People’s Republic of China Zhijun Ning
    • Energy Materials and Surface Sciences Unit (EMSSU), Okinawa Institute of Science and Technology Graduate University (OIST), Okinawa, 904-0495, Japan Yabing Qi
    • School of Materials Science and Engineering, Henan Institute of Advanced Technology, Zhengzhou University, Zhengzhou, 450001, People’s Republic of China Yiqiang Zhang

    Leave a Reply

    Your email address will not be published. Required fields are marked *