Next Article in Journal
Evaluation of the Efficiency of Photoelectrochemical Activity Enhancement for the Nanostructured LaFeO3 Photocathode by Surface Passivation and Co-Catalyst Deposition
Previous Article in Journal
The Use of Pyrolytic Char Derived from Waste Tires in the Removal of Malachite Green from Dyeing Wastewater
Previous Article in Special Issue
Simulation and Investigation of 26% Efficient and Robust Inverted Planar Perovskite Solar Cells Based on GA0.2FA0.78SnI3-1%EDAI2 Films
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Modification of SnO2 Electron Transport Layer in Perovskite Solar Cells

by
Helen Hejin Park
1,2
1
Advanced Materials Division, Korea Research Institute of Chemical Technology (KRICT), Daejeon 34114, Republic of Korea
2
Department of Advanced Materials and Chemical Engineering, University of Science and Technology (UST), Daejeon 34113, Republic of Korea
Nanomaterials 2022, 12(23), 4326; https://doi.org/10.3390/nano12234326
Submission received: 13 November 2022 / Revised: 28 November 2022 / Accepted: 2 December 2022 / Published: 5 December 2022
(This article belongs to the Special Issue New Advances for Halide Perovskite Materials and Applications)

Abstract

:
Rapid development of the device performance of organic-inorganic lead halide perovskite solar cells (PSCs) are emerging as a promising photovoltaic technology. Current world-record efficiency of PSCs is based on tin oxide (SnO2) electron transport layers (ETLs), which are capable of being processed at low temperatures and possess high carrier mobilities with appropriate energy- band alignment and high optical transmittance. Modification of SnO2 has been intensely investigated by various approaches to tailor its conductivity, band alignment, defects, morphology, and interface properties. This review article organizes recent developments of modifying SnO2 ETLs to PSC advancement using surface and bulk modifications, while concentrating on photovoltaic (PV) device performance and long-term stability. Future outlooks for SnO2 ETLs in PSC research and obstacles remaining for commercialization are also discussed.

1. Introduction

Solar energy is one of the most plentiful energy resources accessible to humankind. Among the various PV technologies, currently monocrystalline-silicon-solar cells dominate the PV market due to its high PCE of 26.1% and high working stability, but suffers from cost-intensive production cost of the highly purified monocrystalline silicon. The low production cost and rapid increase in unit-cell efficiency of organic-inorganic PSCs, which is currently 25.7% [1], enables it to compete with silicon-solar cells. Organometallic- halide perovskites are based on the chemical formula of ABX3, where A is organic or metal cations, such as methylammonium (CH3NH3+ (MA+)), formamidinium ((NH2)2CH+ (FA+)), Rb+, or Cs+, B is metal ions, such as Pb2+ or Sn2+, and X is halogen ions, such as I, Br, or Cl. Organometallic-halide perovskites possess features of an ideal absorber material, including high absorption coefficients (~10−4 cm−1), long carrier-diffusion lengths (>1 μm), ambipolar-charge-transport capabilities, and low exciton-binding energy (20–50 meV), [2].
A typical n-i-p PSC device structure consists of glass/transparent conducting oxide (TCO)/n-type ETL/perovskite absorber/p-type hole transport layer (HTL)/metal contact. The bottom TCO is usually fluorine-doped tin oxide (FTO) or indium tin oxide (ITO), and the top metal contact is typically gold (Au), silver (Ag), or aluminum (Al). For a typical p-i-n PSC device, the structure consists of glass/TCO/HTL/perovskite/ETL/top metal contact.
Among the various layers in PSCs, the ETL plays a key role in photovoltaic performance and the charge dynamics. Traditionally, mesoporous titanium dioxide (mp-TiO2) was used as the ETL in n-i-p structured devices with 2,2′,7,7′-tetrakis-(N,N-di-4-methoxyphenylamino)-9,9′-spirobifluorene (spiro-OMeTAD) or poly[bis(4-phenyl)(2,5,6-trimethylphenyl)amine] (PTAA) as the HTL. However, high-temperature sintering is required for the mp-TiO2 to enhance the crystallinity and remove organic material in the TiO2 paste. Furthermore, there is a charge barrier at the TiO2/perovskite interface, leading to inefficient charge transfer and a large charge accumulation at the interface. Other organic materials, including fullerene and its derivatives, have also been used as ETLs in PSCs [3,4]. However, such organic-conducting materials have unreliable stability against light, thermal, and environmental factors. Other low-temperature processed metal oxides, such as tin oxide (SnO2) [5], zinc oxide (ZnO) [6,7], tungsten oxide (WO3) [8], indium oxide (In2O3) [9], niobium oxide (Nb2O5) [10], cerium oxide (CeOx) [11], Zn2SO4 [12], BaSnO3 [13], SrTiO3 [14], and cadmium sulfide (CdS) [15] have also been explored. ZnO can be processed at low temperatures, but the -OH residue on the ZnO surface causes to decompose the perovskite layer leading to environmentally unstable devices. The high toxicity of cadmium becomes a concern regarding to CdS. In addition, the low bandgap of CdS (2.4 eV) causes current loss in the UV range.
Compared to the traditional mp-TiO2 ETL in PSCs, SnO2 can be fabricated at lower temperatures, which expands its possible applications to flexible substrates. SnO2 also has a more favorable band alignment in PSCs compared to TiO2, as the conduction-band minimum is lower than that of TiO2, which induces higher efficiency of charge injection. It also possesses good electrical conductivity and high optical transmittance [16,17]. Furthermore, the larger bandgap of SnO2 compared to TiO2 makes it less vulnerable to UV light. Thus, the device stability can be improved by suppressed UV-related photochemical reaction. Some drawbacks of SnO2 are the bulk and surface defects in the pristine films, such as surface-adsorbed hydroxyl groups, uncoordinated Sn4+, and oxygen vacancies, which deteriorate electronic properties and capture electrons, degrading the stability and efficiency of PSCs. Such drawbacks of SnO2 can be alleviated by modifying the bulk or surface of the film, which will be discussed in this review.
Suitable ETL materials should (i) have a favorable energy-band alignment with the absorber layer for efficient electron transfer from the perovskite to ETL, (ii) exhibit high transparency to minimize optical losses, (iii) possess a wide optical bandgap so that there will be no contribution as a second absorber layer and can also benefit the photochemical reaction, as the wide bandgap will prevent reaction with high-energy radiation, such as UV-light, which is key for the device stability, (iv) have a high conductivity for high fill factors and low series resistance, (v) exhibit appropriate hydrophobic behavior to tolerate long-term exposure to humidity, (vi) be environmentally friendly, and (vii) have low fabrications and materials costs [17,18,19,20,21].
The modification of SnO2 has been investigated by various institutions and has contributed to enhancing the PV performance and device stability of PSCs. In this review, the recent contributions of SnO2 ETLs in PSCs are organized based on photovoltaic performance and stability. Section 2 will cover modification of SnO2 ETLs in PSCs using elemental doping, insertion of metal-oxide layers, ionic compounds, carbon materials, and organic molecules. Section 2 will also discuss the long-term stability of the modified SnO2 ETL-based PSCs. Development in the device efficiency and stability are organized in Table 1, Table 2, Table 3, Table 4, Table 5, Table 6, Table 7, Table 8 and Table 9, respectively. Section 3 will summarize the SnO2 modification approaches and discuss future outlooks for commercialization.

2. Device Performance and Stability of SnO2 ETL-Based PSCs

2.1. Elemental Doping

Elemental doping of SnO2 is a straightforward method to effectively alter the conductivity, defect states, and energy level. SnO2 can be simply doped by various elements to tune its electrical and chemical properties. SnO2 is an n-type material. The tetravalent- Sn sites can be replaced by cations with low valence states, such as gallium (Ga3+), cobalt (Co3+), zinc (Zn2+), magnesium (Mg2+), and lithium (Li+) for p-type doping, or can be substituted by cations with high valence states, such as antimony (Sb5+), molybdenum (Mo5+), tantalum (Ta5+), and niobium (Nb5+) for n-type doping.
Li et al. reported a significant improvement in conductivity without declining the transmittance by doping SnO2 with Ta [22]. Ta-doped SnO2 improved the PCE from 19.5% to 20.9% by improved fill factor (FF) and short-circuit current density (JSC), as shown in the illuminated current-density (J-V) curves in Figure 1a. This is due to effective acceleration of electron collection and transfer, and reduction in recombination at the ETL/absorber interface, and shown in the steady state photoluminescence (PL) and time-resolved PL (TRPL) results in Figure 1b,c.
Doping can also be performed by non-metallic elements, such as fluorine (F), to replace the oxygen anion sites of SnO2. X. Gong et al. reported that gradual bilayer replacement of F into SnO2 can decrease the band offset and the ETL/perovskite interface, resulting in increased built-in electric field and enhancing the open-circuit voltage (VOC) [23] from 1.03 to 1.13 V. PSC devices resulted in enhancement of PCE from 16.3% to 20.2%, as shown in Figure 1d. Improved electron-extraction ability is suggested, based on the steady-state PL and TRPL results in Figure 1e,f. Encapsulated devices retained over 85% of its initial efficiency after stored in air at room temperature and 40–50% relative humidity for 300 h.
M. Park et al. reported a solution process to effectively dope SnO2 with Li at a low processing temperature [24]. Li-doping enhanced the conductivity of SnO2 and produced a reduction of the conduction-band energy, as shown in Figure 1g, facilitating the transfer of electrons and reduced the charge recombination. This resulted in improved VOC, FF, and JSC with a PCE increase of 15.3% to 18.2%, as shown in Figure 1h,i.
Niobium doping of SnO2 was reported by Ren et al. using a solution-processable low-temperature method [25]. The improvement in PV performance originates from the increased conductivity and improved surface morphology, which lead to enhanced electron extraction and inhibited charge recombination. Unencapsulated devices maintained 90% of its initial PCE after 288 h stored in air at room temperature.
Yttrium (Y3+) doping of SnO2 reported by Yang et al. promotes more homogeneous distribution and well-aligned SnO2 nanosheet arrays, which leads to improved electron transfer from the absorber to the ETL [26]. Enlarged bandgaps from the Y-doping and a higher conduction-band energy allows improved energy band alignment and reduced the charge recombination at the ETL/absorber interface. This improved the PCE of the PSC from 13.4% to 17.3% by increasing VOC, FF, and JSC.
Bai et al. reported that Sb:SnO2 nanocrystals was used to replace the undoped SnO2 ETL [27]. This shifted the Fermi-energy level upward, which improved the energy- band alignment and reduced charge recombination. Electron = recombination lifetime was longer, and VOC and FF increased with less photocurrent hysteresis. PCE values increased from 15.7% to 17.2%. Unsealed devices retained over 95% of its initial efficiency after 504 h stored in a desiccator at room temperature.
Gallium-doped SnO2 reported by Roose et al. observed decreased trap-state density in the ETL, leading to a reduced recombination rate [28]. VOC and FF increased from 1.00 to 1.07 V and 57.0% to 70.0%, respectively, resulting in a PCE enhancement of 12.5% to 17.0%. Unencapsulated devices maintained about 70% of its initial efficiency after 1000 h under continuous 1 SUN illumination in nitrogen (N2).
Xiong et al. reported Mg-doped SnO2 as the ETL in PSCs [29]. An optimum-Mg content resulted in uniform, smooth, and dense films with reduced free-electron density, which suppressed the charge recombination, and increased electron mobility, which enabled fast extraction of electrons, contributing to improved JSC. PCE improved from 8.2% to 15.2% by doping the ETL with Mg. Unencapsulated devices maintained over 90% of its initial PCE after 720 h of storage in air with <20% relative humidity.
Aluminum doped SnO2 reported by Chen et al. resulted in increased JSC and FF by using a low-temperature solution-processable method [30]. Doping SnO2 with Al enhanced the charge transport and electron extraction based on TRPL tests. PSC devices exhibited improved PCE of 9.0% to 12.1%.

2.2. Metal Oxide

Binary layers of SnO2 with other metal oxides is another approach to modify the ETL in PSCs. This approach leads to tailoring the surface morphology and tunes the energy band alignment. Wang et al. investigated inserting indium oxide (In2O3) between ITO and SnO2, which resulted in reduced trap densities in the perovskite with improved charge transfer and band alignment [31], as shown in Figure 2a. This resulted in higher VOC and FF values with a PCE enhancement of 21.4% to 23.1%, as shown in Figure 2b. Unsealed devices retained 98% of its initial efficiency after being stored in N2 for 1920 h, as shown in Figure 2c, resulting in 91% of its initial PCE after 180 h of continuous 1 SUN illumination and about 80% of its initial PCE after 120 h of exposure to 75% relative humidity.
Europium (Eu) doped tungsten-oxide (WOx) nanorods were inserted between the perovskite and SnO2 layer by Chen et al. [32], which contributed to improved crystallinity of the perovskite film and enhanced conductivity and carrier mobility of both the SnO2 ETL and spiro-OMeTAD HTL. Due to the energy level of the Eu:WOx nanorods, as shown in Figure 2d, the electron and hole extraction were remarkably boosted at the ETL/absorber and absorber/HTL interface, improving the PV performance, as shown in Figure 2e. Unencapsulated devices retained over 90% of its initial PCE after 500 h of continuous 1 SUN illumination at 16–25 °C and 20–30% relative humidity, and over 90% of its initial PCE after 2000 h of exposure to ambient air, as shown in Figure 2f.
Song et al. investigated inserting an anodized-TiO2 between the FTO and SnO2 ETL, which lead to a defect-free physical contact and improved electron extraction [33], as illustrated in Figure 2g. Such bi-layered ETLs resulted in a large change in free energy and moderate electron mobility, as illustrated in Figure 2h. This enhanced the VOC from 1.14 to 1.20 V, which led to an enhanced PCE of 19.0% to 21.1%, as shown in Figure 2i. Similarly, Lee et al. investigated combining a compact TiO2 between the TCO and SnO2 ETL, which improved charge collection due to better hole-blocking ability of the TiO2 underlayer [34]. This resulted in increasing the PCE from 16.4% to 19.8%.
Unencapsulated devices maintained over 95% of its initial PCE after 1200 h of storage in air with 20% relative humidity.
Dagar et al. studied inserting a thin magnesium-oxide (MgO) overlayer on top of SnO2, which led to more uniform films and reduced interfacial-carrier recombination [35]. This resulted in better stability and enhanced device performance from a PCE of 15.2% to 19.0%. Unsealed devices maintained 67% of its initial PCE after 2568 h in 30% relative humidity.
A bilayer of lead oxide (PbO) doped-SnO2 with undoped SnO2 was investigated by Bi et al. [36], which improved the shunt resistance and enhanced the FF from 72.9% to 75.5% and PCE from 17.0% to 18.8%. Devices maintained over 90% of its initial efficiency after 1080 h at room temperature with 15% relative humidity.
Ultrathin MgO was also inserted between FTO and SnO2 ETL in PSCs by Ma et al., which resulted in enhanced electron transporting and hole-blocking properties [37]. Due to MgO passivation, less FTO-surface defects were observed along with a smoother surface and suppressed carrier recombination. The PCE increased from 16.4% to 18.2%.
Hou et al. studied chemical-bath deposition of a SnO2/TiO2 bilayer [38], which facilitates charge separation achieving effective extraction and transport of electrons. Higher electron mobility and suppressed recombination was observed due to the reduced energy barriers and gradual-energy levels, which lead to a PCE enhancement of 12.0% to 18.1%.
Yan et al. reported a bi-layered ETL of SnO2 and zinc oxide (ZnO), which showed superior electron extraction and a lower charge recombination rate [39]. This resulted in a VOC enhancement of 1.06 to 1.23 V, and a PCE enhancement of 11.9% to 14.6%. Unencapsulated devices retained 80% of its initial efficiency after 300 h in N2 at 85 °C.

2.3. Ionic Compounds

2.3.1. Surface Modification by Ionic Compounds

Adding ionic compounds into the SnO2 solution or applying it to the surface the SnO2 are some other cost-effective approaches to modifying the SnO2 ETL. Compared to organic molecules and carbon materials, ionic compounds are usually more stable and lower in cost.
Zhuang et al. looked into the modification of SnO2 by using rubidium fluoride (Rb) by using two different methods: (i) adding RbF into the SnO2 solution and (ii) inserting RbF at the SnO2/perovskite interface [40]. Adding RbF to the SnO2 bulk resulted in improved electron mobility, while adding RbF to the surface of SnO2 resulted in inhibited ion migration and reduced carrier recombination due to the Rb+ cations escaping into the bulk perovskite. This led to increased VOC with PCE of over 23%, as shown in Figure 3a with negligible hysteresis. A stronger steady-state PL intensity of the RbF-treated SnO2 corresponds to improved perovskite-film quality, as shown in Figure 3b. Based on the TRPL results in Figure 3c, a shorter fast decay (τ1) indicates enhanced electron extraction, while the longer slow decay (τ2) corresponds to decreased defects/traps and improved perovskite-film quality. Unencapsulated PSCs maintained about 75% of its initial PCE after 200 h of exposure to white LED light illumination.
Chen et al. studied the combined effect of doping planar-SnO2 (p-SnO2) with different concentrations of RbF and depositing RbF onto the mesoporous SnO2 (m-SnO2) layer [41], resulting in improved PV performance, as shown in Figure 3d. RbF modification increased the conductivity of SnO2, as shown in Figure 3e, and passivated interfacial traps through F-Sn bonds. Rb+ diffused into the perovskite which passivated the perovskite and suppressed ion migration. Unencapsulated devices maintained 90% of its initial PCE after 300 h at the maximum power point (MPP) under 1 SUN illumination, as shown in Figure 3f.
Chen et al. reported surface modification of the SnO2 by using 4-imidazoleacetic acid hydrochloride (ImAcHCl) [42], as illustrated in Figure 3g. The chloride anion in ImAcHCl improves the crystallinity of the perovskite layer. Modifying SnO2 with ImAcHCl shifts the conduction and valence bands up, as shown in Figure 3h, suppresses carrier recombination, and enhances carrier lifetime. This results in the enhancement of PCE from 19.5% to 21.0%, as shown in Figure 3i. Unencapsulated-PSC devices retained 95% of its initial PCE after 840 h at room temperature with 46–60% relative humidity. Unencapsulated PSCs based on PTAA maintained 90% of its initial PCE after 40 h of exposure to 85 °C in N2.
Potassium-hydroxide (KOH) modification of SnO2 surfaces was investigated by Bu et al., which resulted in suppressed hysteresis and enhanced PV performance with a PCE increase from 19.3% to 20.5% [43]. Potassium cations were shown to passivate the ETL/perovskite interface, facilitate grain growth of the perovskite, and improve stability.
Cesium-carbonate (Cs2CO3) post-treatment, studied by Li et al., improves the electrical properties of SnO2 and passivates the ETL/perovskite interface [44]. Such Cs2CO3 modification improves the surface wettability of the ETL and reduces the roughness, resulting in a perovskite film with larger grains. The Cs2CO3 post-treatment lowers the work function, reducing electron-hole recombination, and enhancing the electron transfer. Devices retained 91% of its initial efficiency after stored in 35–45% relative humidity for 340 h.
Alkali-metal cations, such as lithium chloride (LiCl), sodium chloride (NaCl), potassium chloride (KCl), rubidium chloride (RbCl), and cesium chloride (CsCl), were used to modify the SnO2 surface [45]. Such modification increases the mobility and reduces the trap density of states of SnO2. Efficient defect passivation suppresses the recombination at the ETL/absorber interface. Devices based on NaCl-treated SnO2 maintained over 90% of its initial PCE after 960 h.

2.3.2. Bulk Incorporation of Ionic Compounds

Cobalt chloride hexahydrate (CoCl2·6H2O) was incorporated into SnO2 by Wang et al. which shows better band alignment, as shown in Figure 4a; it enhanced charge extraction and suppressed interfacial recombination [46]. This enhanced the VOC up to 1.20 V for a perovskite layer with a bandgap of 1.54 eV. PSC devices with a PCE of 23.8% was achieved, as shown in Figure 4b, with enhanced stability maintaining 84% of initial efficiencies after 200 h of continuous light exposure, as shown in Figure 4c.
The contact between the SnO2 and perovskite was improved by introducing a biological polymer, heparin potassium, to the SnO2 [47]. Such bulk incorporation regulated the arrangement of SnO2 nanocrystals and induced vertically aligned crystal growth of the perovskite, as illustrated in Figure 4d. This improved the PV performance, as shown in Figure 4e. Due to the strengthened interface binding, device-operational stability was enhanced resulting in maintaining 97% of the initial efficiency after 1000 h under 1 SUN illumination at the MPP, as shown in Figure 4f.
Introducing KCl to the SnO2 ETL passivated both the grain boundaries of the perovskite and the defects at the ETL/absorber interface [48]. The Cl and K+ ions passivate the ETL/absorber contact, while the K+ ions in the ETL diffuse into the perovskite layer and passivate the grain boundaries, resulting in enhanced VOC from 1.08 to 1.14 V and increased PCE from 20.2% to 22.2%, as shown in Figure 4g. The stronger steady-state PL intensity with the presence of KCl demonstrates suppressed recombination of the perovskite, as shown in Figure 4h. Based on TRPL results in Figure 4i, the decrease in τ1 suggests faster electron transfer with the incorporation of KCl, and the increase in τ2 suggests slower recombination in the perovskite film grown on KCl-incorporated SnO2. Unsealed devices maintained 88% of its initial PCE after 120 h of continuous 1 SUN illumination.
Girard’s Reagent T (GRT) was introduced to the SnO2 nanoparticles by Bi et al. [49]. The carbonyl group in GRT is anticipated to prevent agglomeration of the SnO2 nanoparticles. The quaternary-ammonium-chloride salt in GRT is expected to facilitate the crystal growth of the perovskite, and the quaternary-ammonium cation and chloride anion can passivate the defects at the ETL/absorber interface. Such GRT incorporation into SnO2 resulted in better electrical properties of SnO2, promoted vertical growth of the perovskite, and reduced interfacial defects at the ETL/perovskite interface, resulting in a PCE enhancement of 19.8% to 21.6%. Unencapsulated devices retained over 99% of its initial efficiency after 720 h at 60 °C, and 59% after 672 h under 1 SUN illumination.
Liu et al. introduced ammonium chloride (NH4Cl) into SnO2, which resulted in PSC devices with negligible hysteresis and improvement in PCE from 18.7% to 21.4% [50]. Such improvement is due to the increased electron mobility, and improved band alignment and passivation of the ETL/perovskite interface, which also improved the device’s stability. Unencapsulated devices stored in N2 retained over 95% after 1000 h.
Sun et al. incorporated potassium sodium tartrate (PSTA) into the SnO2 colloidal dispersion [51]. PTSA contains mobile-alkali-metal cations leading to improved uniformity and conductivity, and less defects in SnO2, which improves the crystallinity of the perovskite film. Sodium and potassium cations can diffuse into the perovskite and passivate the defects at the grain boundaries and surface. This results in PCE enhancement of 18.3% to 21.1% with reduced hysteresis. Device stability improves with unencapsulated devices retaining over 95% of its initial efficiency after 1440 h of exposure to ambient air with 45% relative humidity at 25 °C.
Phosphoric acid was introduced into SnO2 by Jiang et al. to eliminate dangling bonds on the SnO2 surface and improve the carrier-collection efficiency [52]. Electron mobility increased by 3 times and surface-trap states reduced, decreasing the electron- transport barriers of SnO2. Attributed to the enhanced electron-collection efficacy, the PCE increased from 19.7% to 21.0%.
Tetramethylammonium hydroxide (TMAH) was incorporated into SnO2 by Huang et al. at low temperatures of 100–150 °C [53]. Such modification of SnO2 attributed to higher conductivity of SnO2 and also effectively passivated the grain boundaries of the perovskite film. Improved charge transport between the perovskite and ETL resulted in enhanced efficiencies from 18.1% to 20.5%. Encapsulated devices in 15% relative humidity maintained 97% of its initial PCE after 360 h.
Modification of SnO2 by introducing cesium fluoride (CsF) into the ETL was investigated by Akin et al., which led to improved optoelectronic properties and rapid extraction of photogenerated electrons [54]. By combining the modification of SnO2 and inserting zwitterion molecules at the perovskite/HTL interface, a high VOC value of 1.23 V and a PCE value of 20.5% were achieved. Device-operational stability was also improved retaining over 90% of its initial PCE after 800 h under continuous 1 SUN illumination at the MPP.
Ammonium sulfide [(NH4)2S] is incorporated in SnO2 reported by Ai et al., which passivated the surface defects by terminating the Sn-S dangling bonds [55]. The conductivity and electron mobility of SnO2 are increased, enhancing the electron collection and lowering electron-hole recombination rate. The Sn-S-Pb anchors the perovskite crystals at the ETL/absorber interface, which enhances the stability and electron extraction of the PSC. PCE values increase from 18.7% to 20.0% by using this method.

2.4. Carbon Materials

2.4.1. Surface Modification by Carbon Materials

Insertion of highly conductive carbon material at the ETL/absorber interface can facilitate electron transfer. A smooth ETL surface can also be enabled to impact the growth and nucleation of the perovskite layer on top. In addition, defects at the ETL/absorber interface can be passivated by the carbon material. For example, Wang et al. investigated the application of novel fulleropyrrolidine (NMBF-X, X = H or Cl) monomers and dimers, as shown in Figure 5a, in between the perovskite and ETL [56]. The chlorinated- fullerene dimers resulted in the most efficient PCE of 22.3% with minimal hysteresis, as shown in Figure 5b,c, which stems from the passivation of the NMBF-Cl dimer with the SnO2 and perovskite simultaneously. After 1000 h, unencapsulated PSC devices exposed to air at 25–35 °C with 45–60% relative humidity retained over 95% of its initial PCE.
Polystyrene (PS) was inserted between the ETL and absorber layer to release residual stress in the absorber during annealing, which attributes to reduced interface defects, less recombination, and lower ion-migration tendencies [57]. PS was also applied on top of the perovskite film for inner-encapsulation, as shown in Figure 5d, which improves the long-term device stability maintaining over 90% or its initial efficiency after 72 h at the MPP under continuous 1 SUN illumination, as shown in Figure 5e. Devices with the PS- inner encapsulation maintained over 90% of its initial efficiency after 2800 h in air, as shown in Figure 5f.
Adding graphene quantum dots (GQDs) to the SnO2 surface, as shown in Figure 5g, will improve the conductivity and fill the electron traps in SnO2, which improves the electron extraction and reduces the recombination at the ETL/absorber interface [58]. PSC devices with GQD-treated SnO2 results in a VOC enhancement of 1.10 to 1.13 V and a PCE enhancement of 17.9% to 20.3% with little hysteresis, as shown in Figure 5h,i. Unencapsulated devices maintain over 95% of its initial efficiency in N2 after 720 h and then an additional 1440 h in air (20–30% relative humidity).
Zheng et al. reported modifying the SnO2 ETL surface with [6,6]-phenyl-C61-butyric acid methyl ester (PCBM) [59]. By simultaneously inserting PbS above the HTL and spiro-OMeTAD, a PCE of 19.6% was demonstrated with long-term storage stabilities of about 100% retained after exposure to air for 1000 h.
Ke et al. also reported fullerene, PCBM, and modification of the SnO2 ETL surface, which resulted in an increase in PCE from 16.5% to 19.1% [60]. PCBM promotes enhanced electron transfer, suppressed carrier recombination at the ETL/absorber interface, and passivates both the ETL/absorber interface and the perovskite-grain boundaries.
Plasma-enhanced atomic layer deposition (PEALD) was reported to use a low deposition temperature of 100 °C and still attained high efficiencies by surface modification with passivation by C60-self-assembled monolayer (C60-SAM) [61]. The VOC improved from 1.07 to 1.13 V and FF increased from 75.5% to 79.1%, resulting in an efficiency enhancement from 17.2% to 19.0%. Unencapsulated devices maintained over 98% of its initial PCE after 480 h in N2 with room-light exposure and a relative humidity below 10%.
A comparative study between passivation of SnO2 by [6,6]-phenyl-C61-butyric acid (PCBA) and PCBM was performed by Wang et al. [62]. Passivation of fullerene-molecules at the SnO2 surface efficiently decreases the defects and improves the conductivity through enhanced electron mobility. The PCBA modification showed higher passivation efficiency, resulting in a PCE increase from 15.4% to 18.6%.

2.4.2. Bulk Incorporation of Carbon Materials

Highly conductive carbon-based materials incorporated into the SnO2 can enhance the conductivity of the ETL. For example, polymeric carbon nitrides (cPCNs) were introduced into the SnO2 nanocrystals by Li et al., which led to electron mobilities three times higher than that of the undoped SnO2 [63]. Such modification of SnO2 led to less wettability, reduced grain boundaries from the suppressed heterogeneous nucleation of the perovskite, and improved the band alignment, as shown in Figure 6a. This resulted in negligible hysteresis with a PCE enhancement of 21.3% to 23.2%, as shown in Figure 6b. Unencapsulated devices retained 95% of its initial PCE after 2880 h in N2, as shown in Figure 6c.
Niu et al. studied the effect of incorporating Nb2C MXenes into SnO2, which led to increased grain growth and increased lattice-spacing facets of SnO2 [64]. A decrease in the steady-state PL, shown in Figure 6d, demonstrates enhanced photogenerated carrier generation and reduced recombination. Superior crystallinity and effective carrier transport resulted in improved PSC PCE from 19.0% to 22.9%, as shown in Figure 6e. Unencapsulated devices retained 98% of its initial PCE after 960 h in 40–60% relative humidity at 25 °C, as shown in Figure 6f.
Red-carbon quantum dots were used to dope solution-processed SnO2, which resulted in increasing the electron mobility by 20 times [65]. The enhanced electron mobility also showed to help passivate the traps and defects at the ETL/perovskite interface and promote growth of highly crystalline perovskite. Efficiencies improved from 19.2% to 22.8% through improvement of JSC, VOC, and FF, as shown in Figure 6g. Unencapsulated devices maintained 96% of its initial PCE after 1000 h in air with 40–60% relative humidity at 25 °C, as shown in Figure 6h. Reduced steady-state PL intensity of the perovskites grown on the SnO2 doped with red-carbon quantum dots suggest improved photogenerated-carrier generation and suppressed recombination, as shown in Figure 6i.
A water-soluble nonionic polymer, polyacrylamide (PAM), is introduced into SnO2 be Dong et al., which improves the electron mobility, wettability, and uniformity of SnO2 [66]. Defects in the perovskite is also reduced and grain size is increased from the PAM addition into SnO2. Band alignment at the ETL/perovskite interface is also improved, resulting in a PCE enhancement from 20.2% to 22.6%. Unencapsulated devices maintained 90% of its initial efficiency after 1080 h of exposure to 45%–55% humidity.
Graphitic carbon nitride (g-C3N4) quantum dots were added to SnO2 and applied to PSC devices by Chen et al. [67]. The oxygen-vacancy-reduced trap centers were effectively eliminated and bulk and interface-electron transport were promoted. The high conductivity and suitable energy band alignment of g-C3N4-treated SnO2 led to a PCE improvement of 20.2% to 22.1%. Unencapsulated-PSC devices retained 90% of its initial efficiency after 1500 h in air with 60% relative humidity at 25 °C, and 80% of its initial efficiency after 75 h in air with 60% relative humidity at 85 °C.
Zhang et al. incorporated graphdiyne into SnO2, which led to improved electron mobility [68]. The enhanced hydrophobicity inhibits heterogeneous-perovskite nucleation, attributing to the reduced grain boundaries and less defect density. The ETL/absorber interface is also improved from passivation of the Pb-I anti-site defects. This results in a PCE increase of 19.2% to 21.1%.
Zhao et al. introduced naphthalene-diimide graphene into nanocrystal SnO2 [69]. Such modification increases the surface hydrophobicity of SnO2 and forms a van-der-Waals interaction at the ETL/perovskite interface. Enhanced FF values from 74.6% to 82.1% are attributed to the enhanced electron mobility and electron-extraction efficiency, and reduced carrier recombination. PCE improves from 19.0% to 20.2%.
Carbon nanodots by a hydrothermal process were incorporated into SnO2 by Wang et al. using a simple solution process [70]. Incorporation of carbon nanodots into SnO2 reduced the density of trap-states and increased the electron mobility of SnO2, resulting in a PCE of 20.0% with negligible hysteresis. As carbon materials exhibit high conductivity, incorporation of a carbon material into the ETL can enhance its charge transport and conductivity. Carbon materials can also passivate defects, improving the stability and efficiency of PSCs. Unencapsulated PSCs maintained 90% of its initial efficiency after 200 h of UV exposure in air with 20–30% relative humidity at 20 °C.

2.5. Organic Molecules

2.5.1. Surface Modification by Organic Molecules

Modification of SnO2 by organic molecules can passivate defects at the ETL surface and improve the electrical properties and transport at the interface. The ETL and absorber layer can be chemically bridged by the organic molecule enhancing the interfacial-electron transfer. For example, Lou et al. investigated the introduction of π-conjugated n-type small organic molecules (BTAC4 and Y6) onto the surface of KCl-doped SnO2 ETL [71], as illustrated in Figure 7a, which yields less trap states, suppressed carrier recombination, and improved electron transport and extraction. Band alignment improves with the modification, as shown in Figure 7b. Applying BTAC4 and Y6 to the surface of SnO2 results in an enhanced PCE from 21.2% to 23.1% and 22.1%, respectively, as shown in Figure 7c. Unencapsulated devices maintained about 90% of its initial efficiency after 768 h in air with 35% relative humidity.
An iodine-terminated SAM, 3-iodopropyl trimethoxysilane [Si(OCH3)3(CH2)3I, I-SAM] was applied to the SnO2 ETL surface in PSCs, which increased the adhesion toughness at the ETL/absorber interface-enhancing mechanical reliability [72], as shown in Figure 7d. This is attributed to the higher toughness and decreased hydroxyl groups at the interface. Without the SAMs treatment, irreversible morphological degradation, such as voids and delamination, was observed at the ETL/perovskite interface for operational stability tested-devices. Treatment with I-SAM on the SnO2 enhanced the PCE from 20.2% to 21.4% with diminished hysteresis, as shown in Figure 7e. Long-term working stability was also improved, retaining over 90% of its initial efficiency in N2 with continuous 1 SUN illumination at the MPP for unencapsulated devices for 1200 h, as shown in Figure 7f.
Triphenylphosphine oxide (TPPO) is an air-robust and cost-effective molecule for n-type doping of SnO2 [73], as shown in Figure 7g. Surface modification by TPPO enhanced the conductivity and lowered the work function of SnO2, as shown in Figure 7h,i. The VOC improved from 1.08 to 1.11 V and PCE improved from 19.0% to 20.7% attributed to the lower recombination rate and faster electron extraction.
Thiophene-based interlayers were adopted to the SnO2 surface to reduce the energy loss by optimizing the surface-electronic states of SnO2 and improving the perovskite- film quality [74]. Surface modification of SnO2 by thiophene-3-acetic acid improved the conductivity and lowered the work function of SnO2. Ion-defect states at the ETL/perovskite interfaces were passivated by bonding of the under-coordinated Pb2+ of MAPbI3 with the sulfur atoms of the thiophene rings with a lone pair of electrons. VOC improved from 1.07 to 1.12 V, and FF improved from 73.5% to 80.1%, resulting in a PCE improvement of 17.5% to 20.6%. This resulted in improved device stability retaining over 90% or its initial efficiency after 1440 h in N2 and over 80% of its initial efficiency after 130 h in air with 70% relative humidity at 85 °C.
Aminosulfonic acid (+H3N-SO3, SA) is introduced to the surface of SnO2 [75], and a chemical bridge is formed between the ETL and perovskite through the coordination bond to SnO2 via -SO3 anions and electrostatic interactions with the perovskite via -NH3+ cations. Better surface wettability of the SA-treated SnO2 led to larger grain size of perovskite films. Attributed to the passivated-contact defects, VOC improved from 1.11 to 1.15 V, while barrier-free charge transferred led to improved FF and JSC with reduced hysteresis, resulting in a PCE improvement of 18.2% to 20.4%. Unencapsulated devices maintained over 80% of its initial PCE after 1000 h in air of 25–35% relative humidity and over 75% of its initial PCE after 500 h in N2 at 60 °C.
The organic molecule p-amino benzenesulfonic acid (ABSA) was introduced to the surface of SnO2 by inactivating the under-coordinated Sn ions [76]. This decreased the energy-band barrier on the surface of SnO2, and increased the conductivity and lessens- carrier recombination. This results in a PCE enhancement of 18.0% to 20.3%. Unsealed PSC devices retain 57% of its initial efficiency after 720 h in N2.
A plant-photosynthesis promoter, choline chloride, was introduced to the surface of SnO2 by a simple molecular self-assembly method [77]. Such modification reduces the oxygen vacancies of SnO2, while the Cl ions form strong Pb-Cl bonds with the uncoordinated Pb ions in the MAPbI3. This passivates the defects at the ETL/absorber interface and reduces carrier recombination, improving the VOC from 1.07 to 1.15 V. PCE is improved from 16.8% to 18.9%.
Zuo et al. investigated various SAMs, such as 4-pyridinescarboxylic acid (PA-SAM), 4-cyanobenzoic acid (CBA-SAM), and benzoic acid (BA-SAM), and applied them to the surface of SnO2 [78]. Proper interfacial interactions were shown to decrease trap- state density and enhance interfacial-charge transfer. Among the various SAMs, application of PA-SAM resulted in the highest efficiency enhancement of 17.2% to 18.8%, with a VOC enhancement from 1.06 to 1.10 V. This is due to improved electronic coupling and suppressed interfacial traps. Charge transfer at the ETL/absorber interface improved from the lowered work function.
Wang et al. reported interfacial-sulfur functionalization anchoring of SnO2 by using potassium hexylxanthate to modify the surface of SnO2 [79]. This approach effectively passivated the charge traps and suppressed carrier recombination at the interface by sulfur functionalization of the SnO2 surface. Functionalized-sulfur atoms can also coordinate with the under-coordinated Pb2+ ions at the interface. Such strategy resulted in improved device efficiencies of 16.6% to 18.4%. Unsealed devices maintained about 90% of its initial PCE after 1680 h at room temperature.

2.5.2. Bulk Incorporation of Organic Molecules

Incorporation of organic molecules possessing versatile functional groups into SnO2 improves the dispersion of colloids, enhances the electrical properties, and passivates the defects in SnO2. For example, Xiong et al. introduced poly(ethylene glycol) diacrylate (PEGDA), as shown in Figure 8a, into the SnO2 dispersion to prevent aggregations [80], which resulted in more uniform film and well-matched band-energy alignment with the perovskite. PEGDA-modified SnO2 also attributed to passivating the defects at the ETL/absorber interface, as shown in Figure 8b,c. This showed a PCE improvement from 21.8% to 23.3%, with a VOC improvement of 1.09 to 1.14 V. Unencapsulated-PSC devices maintained over 90% of its initial PCE after 850 h in N2 under 1 SUN illumination and 98% of its initial efficiency after 1000 h in air with 30–35% humidity.
Luan et al. incorporated 2,2,2-trifluoroethanol (TFE) into the SnO2 ETL [81], as shown in Figure 8d, which showed enhanced electron mobility and optimized energy- band alignment. The modified SnO2 exhibits a very smooth surface, which attributed to the less trap density at the ETL/absorber interface and inside the perovskite film, leading to lessened carrier recombination. With the addition of oxygen-plasma treatments PCE values of 21.7% were achieved with FF over 80%, as shown in Figure 8e. Unencapsulated devices maintained over 90% of its initial efficiency after 720 h in 30–40% relative humidity, as shown in Figure 8f.
Ethylene diamine tetraacetic acid (EDTA) was incorporated into the SnO2 ETL by Yang et al. [82]. Electron transfer is facilitated due to the optimized energy-band alignment, as shown in Figure 8g, and enhanced electron mobility of the modified SnO2. Perovskite film grown on the modified SnO2 also exhibited larger grain size and lower trap density. This led to a PCE enhancement of 18.9% to 21.6%, as shown in Figure 8h. Unencapsulated devices maintained 86% of its initial efficiency after 120 h in continuous 1 SUN illumination, as shown in Figure 8i. The large grain size of the perovskite repressed perovskite degradation at the grain boundaries.
Polyethylene glycol (PEG) introduced into the SnO2 prevented nanoparticle agglomeration, resulting in a dense and uniform film [83]. Such modification of SnO2 improved the wettability and enabled pinhole-free perovskite films, demonstrating a PCE enhancement of 19.2% to 20.8%. Unsealed devices retained over 97% of its initial efficiency after 2160 h in air with 30–80% relative humidity at 28–35 °C.
Polyethylenimine (PEIE) was added into SnO2 by a low-temperature solution process [84]. Such doping of SnO2 resulted in optimized band alignment, larger built-in potential, improved electron transport and extraction, and mitigated charge recombination. Unencapsulated devices retained 82% of its initial efficiency after1,680 h in 40% relative humidity.

3. Conclusions and Future Directions

In summary, recent progress in modifying the SnO2-ETL bulk and surface properties are discussed. SnO2 has been considered the most promising alternative to TiO2, as it has a high electron mobility and conductivity. SnO2 also possesses a suitable band structure owning a deep conduction band allowing enhanced electron extraction at the ETL/absorber interface. The wide bandgap of SnO2 allows most of the light to be absorbed by the perovskite-absorber layer and suppresses UV-related photochemical reaction, improving the device stability. SnO2 can also be processed at a lower temperature than TiO2 allowing for flexible applications to be possible. Despite the many advantages of SnO2, there are drawbacks, such as defects in the surface and bulk of the pristine films, deteriorating its electronic properties. Such drawbacks can be alleviated by the various surface-modification and bulk-incorporation methods discussed in this review.
Surface-modification and bulk-incorporation methods are discussed as strategies, including elemental doping, metal-oxides bilayers, incorporation of ionic compounds, carbon materials, and organic molecules. Among the five main aspects of modifying the ETL, bulk and surface modification using ionic compounds are generally lower in cost and more stable than carbon materials and organic molecules. Bilayer-metal-oxide approaches are simple and straightforward; however, an additional fabrication step and additional material cost is required. Generally, surface-modification approaches require an additional fabrication step, so bulk incorporation may be considered more fabrication-friendly. Various modification at the surface or the bulk can lead to improved morphology, conductivity, and band alignment of SnO2, resulting in improved electron-transport capabilities and reduced carrier recombination. Modification of the surface properties of SnO2 can also lead to improved quality and crystallinity of the perovskite absorber and reduced interfacial defects at the ETL/absorber interface, attributing to improved PV device performance and stability.
Although such modification approaches of SnO2 have demonstrated enhanced device performance and stability, there still requires improvement in the operational stability of the PSCs at the MPP under continuous 1 SUN illumination. There have been many reports on the long-term stability of devices exposed to elevated temperatures and high relative humidity conditions. However, there are a few reports on the MPP tracking under continuous 1 SUN illumination which are working conditions of the PSCs, and is an important evaluation of the long-term operational stability of PSCs. Such long-term working-stability evaluation methods will help create a better understanding of the mechanisms for improved working stability and will be critical for future directions for commercialization. Another important factor to consider for future directions will be the scalability of the modification approaches for the SnO2 ETL, since developing methods compatible with large-area substrates will be essential for commercialization.
Perovskite-tandem applications, with lower bandgap absorbers, such as silicon, Cu(In,Ga)Se2, and tin-related absorbers [85,86,87,88,89,90,91], below wider bandgap perovskite-based solar cells, are also future steps to commercialization [92]. Thus, depending on the bottom- solar cell, there may be limitations in fabrication methods or temperature process of the ETL in the tandem configurations. Especially, tandem devices with flexible substrates will have a limitation on the process temperature of the layers in the top solar cell.

Funding

This work was supported financially by the Korea Research Institute of Chemical Technology (KRICT), Republic of Korea (SS2222-20), from the Korea Institute of Energy Technology Evaluation and Planning (KETEP) and the Ministry of Trade Industry and Energy (MOTIE), Republic of Korea (no. 20203040010320), and was also supported by the National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT (2022K1A4A8A02079724).

Conflicts of Interest

The author declares no conflict of interest.

References

  1. NREL, E.C. Available online: http://www.nrel.gov/pv/ (accessed on 12 November 2022).
  2. Hwang, T.; Lee, B.; Kim, J.; Lee, S.; Gil, B.; Yun, A.J.; Park, B. From Nanostructural Evolution to Dynamic Interplay of Constituents: Perspectives for Perovskite Solar Cells. Adv. Mater. 2018, 30, 1–22. [Google Scholar] [CrossRef] [PubMed]
  3. Yip, H.L.; Jen, A.K.Y. Recent Advances in Solution-Processed Interfacial Materials for Efficient and Stable Polymer Solar Cells. Energy Environ. Sci. 2012, 5, 5994–6011. [Google Scholar] [CrossRef]
  4. Nakamura, T.; Yakumaru, S.; Truong, M.A.; Kim, K.; Liu, J.; Hu, S.; Otsuka, K.; Hashimoto, R.; Murdey, R.; Sasamori, T.; et al. Sn(IV)-free tin perovskite films realized by in situ Sn(0) nanoparticle treatment of the precursor solution. Nat. Commun. 2020, 11, 3008. [Google Scholar] [CrossRef] [PubMed]
  5. Jiang, Q.; Zhang, L.; Wang, H.; Yang, X.; Meng, J.; Liu, H.; Yin, Z.; Wu, J.; Zhang, X.; You, J. Enhanced Electron Extraction using SnO2 for High-Efficiency Planar-Structure HC(NH2)2 PbI3-Based Perovskite Solar Cells. Nat. Energy 2016, 2, 16177. [Google Scholar] [CrossRef]
  6. Chakravarthi, N.; Park, H.-Y.; Kranthiraja, K.; Kim, H.; Shin, J.; Song, M.; Jin, S.-H. Substituent position engineering of phosphine oxide functionalized triazine-based cathode interfacial materials for flexible organic and perovskite solar cells. Org. Electron. 2018, 54, 54–63. [Google Scholar] [CrossRef]
  7. 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. [Google Scholar] [CrossRef]
  8. Wang, K.; Shi, Y.; Dong, Q.; Li, Y.; Wang, S.; Yu, X.; Wu, M.; Ma, T. Low-Temperature and Solution-Processed Amorphous WOX as Electron-Selective Layer for Perovskite Solar Cells. J. Phys. Chem. Lett. 2015, 6, 755–759. [Google Scholar] [CrossRef]
  9. Qin, M.; Ma, J.; Ke, W.; Qin, P.; Lei, H.; Tao, H.; Zheng, X.; Xiong, L.; Liu, Q.; Chen, Z.; et al. Perovskite Solar Cells Based on Low-Temperature Processed Indium Oxide Electron Selective Layers. ACS Appl. Mater. Interfaces 2016, 8, 8460–8466. [Google Scholar] [CrossRef]
  10. Kogo, A.; Numata, Y.; Ikegami, M.; Miyasaka, T. Nb2O5 Blocking Layer for High Open-circuitVoltage Perovskite Solar Cells. Chem. Lett. 2015, 44, 829–830. [Google Scholar] [CrossRef]
  11. Wang, X.; Deng, L.-L.; Wang, L.-Y.; Dai, S.-M.; Xing, Z.; Zhan, X.-X.; Lu, X.-Z.; Xie, S.-Y.; Huang, R.-B.; Zheng, L.-S. Cerium oxide standing out as an electron transport layer for efficient and stable perovskite solar cells processed at low temperature. J. Mater. Chem. A 2016, 5, 1706–1712. [Google Scholar] [CrossRef]
  12. Yeom, E.J.; Shin, S.S.; Yang, W.S.; Lee, S.J.; Yin, W.; Kim, D.; Noh, J.H.; Ahn, T.K.; Seok, S.I. Controllable synthesis of single crystalline Sn-based oxides and their application in perovskite solar cells. J. Mater. Chem. A 2016, 5, 79–86. [Google Scholar] [CrossRef]
  13. Shin, S.S.; Yeom, E.J.; Yang, W.S.; Hur, S.; Kim, M.G.; Im, J.; Seo, J.; Noh, J.H.; Seok, S.I. Colloidally Prepared La-Doped BaSnO3 Electrodes for Efficient, Photostable Perovskite Solar Cells. Science 2017, 356, 167–171. [Google Scholar] [CrossRef] [PubMed]
  14. Bera, A.; Wu, K.; Sheikh, A.; Alarousu, E.; Mohammed, O.F.; Wu, T. Perovskite Oxide SrTiO3 as an Efficient Electron Transporter for Hybrid Perovskite Solar Cells. J. Phys. Chem. C 2014, 118, 28494–28501. [Google Scholar] [CrossRef]
  15. Dunlap-Shohl, W.A.; Younts, R.; Gautam, B.; Gundogdu, K.; Mitzi, D.B. Effects of Cd Diffusion and Doping in High-Performance Perovskite Solar Cells Using CdS as Electron Transport Layer. J. Phys. Chem. C 2016, 120, 16437–16445. [Google Scholar] [CrossRef]
  16. Yun, A.J.; Kim, J.; Hwang, T.; Park, B. Origins of Efficient Perovskite Solar Cells with Low-Temperature Processed SnO2 Electron Transport Layer. ACS Appl. Energy Mater. 2019, 2, 3554–35600. [Google Scholar] [CrossRef]
  17. Xiong, L.; Guo, Y.; Wen, J.; Liu, H.; Yang, G.; Qin, P.; Fang, G. Review on the Application of SnO2 in Perovskite Solar Cells. Adv. Funct. Mater. 2018, 28, 1802757. [Google Scholar] [CrossRef]
  18. Liu, D.; Wang, Y.; Xu, H.; Zheng, H.; Zhang, T.; Zhang, P.; Wang, F.; Wu, J.; Wang, Z.; Chen, Z.; et al. SnO2-Based Perovskite Solar Cells: Configuration Design and Performance Improvement. Sol. RRL 2019, 3, 1800292. [Google Scholar] [CrossRef]
  19. Deng, K.; Chen, Q.; Li, L. Modification Engineering in SnO 2 Electron Transport Layer toward Perovskite Solar Cells: Efficiency and Stability. Adv. Funct. Mater. 2020, 30, 2004209. [Google Scholar] [CrossRef]
  20. Huang, S.; Li, P.; Wang, J.; Huang, J.C.-C.; Xue, Q.; Fu, N. Modification of SnO2 electron transport Layer: Brilliant strategies to make perovskite solar cells stronger. Chem. Eng. J. 2022, 439, 135687. [Google Scholar] [CrossRef]
  21. Jiang, Q.; Zhang, X.; You, J. SnO2: A Wonderful Electron Transport Layer for Perovskite Solar Cells. Small 2018, 14, e1801154. [Google Scholar] [CrossRef]
  22. Liu, Q.; Zhang, X.; Li, C.; Lu, H.; Weng, Z.; Pan, Y.; Chen, W.; Hang, X.-C.; Sun, Z.; Zhan, Y. Effect of tantalum doping on SnO2 electron transport layer via low temperature process for perovskite solar cells. Appl. Phys. Lett. 2019, 115, 143903. [Google Scholar] [CrossRef]
  23. Gong, X.; Sun, Q.; Liu, S.; Liao, P.; Shen, Y.; Graetzel, C.; Zakeeruddin, S.M.; Grätzel, M.; Wang, M. Highly Efficient Perovskite Solar Cells with Gradient Bilayer Electron Transport Materials. Nano Lett. 2018, 18, 3969–3977. [Google Scholar] [CrossRef] [PubMed]
  24. Park, M.; Kim, J.Y.; Son, H.J.; Lee, C.H.; Jang, S.S.; Ko, M.J. Low-Temperature Solution-Processed Li-doped SnO2 as an Effective Electron Transporting Layer for High-Performance Flexible and Wearable Perovskite Solar Cells. Nano Energy 2016, 26, 208–215. [Google Scholar] [CrossRef]
  25. Ren, X.; Yang, D.; Yang, Z.; Feng, J.; Zhu, X.; Niu, J.; Liu, Y.; Zhao, W.; Liu, S.F. Solution-Processed Nb:SnO2 Electron Transport Layer for Efficient Planar Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2017, 9, 2421–2429. [Google Scholar] [CrossRef] [PubMed]
  26. Yang, G.; Lei, H.; Tao, H.; Zheng, X.; Ma, J.; Liu, Q.; Ke, W.; Chen, Z.; Xiong, L.; Qin, P.; et al. Reducing Hysteresis and Enhancing Performance of Perovskite Solar Cells Using Low-Temperature Processed Y-Doped SnO2 Nanosheets as Electron Selective Layers. Small 2017, 13, 1601769. [Google Scholar] [CrossRef] [PubMed]
  27. Bai, Y.; Fang, Y.; Deng, Y.; Wang, Q.; Zhao, J.; Zheng, X.; Zhang, Y.; Huang, J. Low Temperature Solution-Processed Sb:SnO2 Nanocrystals for Efficient Planar Perovskite Solar Cells. ChemSusChem 2016, 9, 2686–2691. [Google Scholar] [CrossRef] [Green Version]
  28. Roose, B.; Johansen, C.M.; Dupraz, K.; Jaouen, T.; Aebi, P.; Steiner, U.; Abate, A. A Ga-doped SnO2 mesoporous contact for UV stable highly efficient perovskite solar cells. J. Mater. Chem. A 2017, 6, 1850–1857. [Google Scholar] [CrossRef] [Green Version]
  29. Xiong, L.; Qin, M.; Yang, G.; Guo, Y.; Lei, H.; Liu, Q.; Ke, W.; Tao, H.; Qin, P.; Li, S.; et al. Performance enhancement of high temperature SnO2-based planar perovskite solar cells: Electrical characterization and understanding of the mechanism. J. Mater. Chem. A 2016, 4, 8374–8383. [Google Scholar] [CrossRef]
  30. Chen, H.; Liu, D.; Wang, Y.; Wang, C.; Zhang, T.; Zhang, P.; Sarvari, H.; Chen, Z.; Li, S. Enhanced Performance of Planar Perovskite Solar Cells Using Low-Temperature Solution-Processed Al-Doped SnO2 as Electron Transport Layers. Nanoscale Res. Lett. 2017, 12, 238. [Google Scholar] [CrossRef] [Green Version]
  31. Wang, P.; Li, R.; Chen, B.; Hou, F.; Zhang, J.; Zhao, Y.; Zhang, X. Gradient Energy Alignment Engineering for Planar Perovskite Solar Cells with Efficiency Over 23%. Adv. Mater. 2020, 32, e1905766. [Google Scholar] [CrossRef]
  32. Chen, X.; Xu, W.; Shi, Z.; Pan, G.; Zhu, J.; Hu, J.; Li, X.; Shan, C.; Song, H. Europium ions doped WOx nanorods for dual interfacial modification facilitating high efficiency and stability of perovskite solar cells. Nano Energy 2020, 80, 105564. [Google Scholar] [CrossRef]
  33. Song, S.; Kang, G.; Pyeon, L.; Lim, C.; Lee, G.-Y.; Park, T.; Choi, J. Systematically Optimized Bilayered Electron Transport Layer for Highly Efficient Planar Perovskite Solar Cells (η = 21.1%). ACS Energy Lett. 2017, 2, 2667–2673. [Google Scholar] [CrossRef]
  34. Lee, Y.; Paek, S.; Cho, K.T.; Oveisi, E.; Gao, P.; Lee, S.; Park, J.S.; Zhang, Y.; Humphry-Baker, R.; Asiri, A.M.; et al. Enhanced Charge Collection with Passivation of the Tin Oxide layer in Planar Perovskite Solar Cells. J. Mater. Chem. A 2017, 5, 12729–12734. [Google Scholar] [CrossRef]
  35. Dagar, J.; Castro-Hermosa, S.; Lucarelli, G.; Cacialli, F.; Brown, T.M. Highly Efficient Perovskite Solar Cells for Light Harvesting under Indoor Illumination via Solution Processed SnO2/MgO Composite Electron Transport Layers. Nano Energy 2018, 49, 290–299. [Google Scholar] [CrossRef] [Green Version]
  36. Bi, Z.; Zhang, S.; Thandapani, M.; Zhu, Y.; Zheng, Y.; Liem, N.Q.; Xiao, X.; Xu, G.; Guerrero, A.; Xu, X. High Shunt Resistance SnO2-PbO Electron Transport Layer for Perovskite Solar Cells Used in Low Lighting Applications. Adv. Sustain. Syst. 2021, 5, 2100120. [Google Scholar] [CrossRef]
  37. Ma, J.; Yang, G.; Qin, M.; Zheng, X.; Lei, H.; Chen, C.; Chen, Z.; Guo, Y.; Han, H.; Zhao, X.; et al. MgO Nanoparticle Modified Anode for Highly Efficient SnO2-Based Planar Perovskite Solar Cells. Adv. Sci. 2017, 4, 1700031. [Google Scholar] [CrossRef] [PubMed]
  38. Hou, Y.; Chen, X.; Yang, S.; Li, C.; Zhao, H.; Yang, H.G. A Band-Edge Potential Gradient Heterostructure to Enhance Electron Extraction Efficiency of the Electron Transport Layer in High-Performance Perovskite Solar Cells. Adv. Funct. Mater. 2017, 27, 1700878. [Google Scholar] [CrossRef]
  39. Yan, L.; Xue, Q.; Liu, M.; Zhu, Z.; Tian, J.; Li, Z.; Chen, Z.; Chen, Z.; Yan, H.; Yip, H.-L.; et al. Interface Engineering for All-Inorganic CsPbI2 Br Perovskite Solar Cells with Efficiency over 14%. Adv. Mater. 2018, 30, e1802509. [Google Scholar] [CrossRef]
  40. Zhuang, J.; Mao, P.; Luan, Y.; Chen, N.; Cao, X.; Niu, G.; Jia, F.; Wang, F.; Cao, S.; Wang, J. Rubidium Fluoride Modified SnO2 for Planar n-i-p Perovskite Solar Cells. Adv. Funct. Mater. 2021, 31, 2010385. [Google Scholar] [CrossRef]
  41. Chen, Q.; Peng, C.; Du, L.; Hou, T.; Yu, W.; Chen, D.; Shu, H.; Huang, D.; Zhou, X.; Zhang, J.; et al. Synergy of mesoporous SnO2 and RbF modification for high-efficiency and stable perovskite solar cells. J. Energy Chem. 2021, 66, 250–259. [Google Scholar] [CrossRef]
  42. Chen, J.; Zhao, X.; Kim, S.; Park, N. Multifunctional Chemical Linker Imidazoleacetic Acid Hydrochloride for 21% Efficient and Stable Planar Perovskite Solar Cells. Adv. Mater. 2019, 31, 1902902. [Google Scholar] [CrossRef] [PubMed]
  43. Bu, T.; Li, J.; Zheng, F.; Chen, W.; Wen, X.; Ku, Z.; Peng, Y.; Zhong, J.; Cheng, Y.-B.; Huang, F. Universal Passivation Strategy to Slot-Die Printed SnO2 for Hysteresis-Free Efficient Flexible Perovskite Solar Module. Nat. Commun. 2018, 9, 4609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Li, T.; Rui, Y.; Wang, X.; Shi, J.; Wang, Y.; Yang, J.; Zhang, Q. Grain Size and Interface Modification via Cesium Carbonate Post-Treatment for Efficient SnO2-Based Planar Perovskite Solar Cells. ACS Appl. Energy Mater. 2021, 4, 7002–7011. [Google Scholar] [CrossRef]
  45. Huang, Y.; Li, S.; Wu, C.; Wang, S.; Wang, C.; Ma, R. Interfacial modification of various alkali metal cations in perovskite solar cells and their influence on photovoltaic performance. New J. Chem. 2020, 44, 8902–8909. [Google Scholar] [CrossRef]
  46. Wang, P.; Chen, B.; Li, R.; Wang, S.; Ren, N.; Li, Y.; Mazumdar, S.; Shi, B.; Zhao, Y.; Zhang, X. Cobalt Chloride Hexahydrate Assisted in Reducing Energy Loss in Perovskite Solar Cells with Record Open-Circuit Voltage of 1.20 V. ACS Energy Lett. 2021, 6, 2121–2128. [Google Scholar] [CrossRef]
  47. You, S.; Zeng, H.; Ku, Z.; Wang, X.; Wang, Z.; Rong, Y.; Zhao, Y.; Zheng, X.; Luo, L.; Li, L.; et al. Multifunctional Polymer-Regulated SnO 2 Nanocrystals Enhance Interface Contact for Efficient and Stable Planar Perovskite Solar Cells. Adv. Mater. 2020, 32, 2003990. [Google Scholar] [CrossRef]
  48. Zhu, P.; Gu, S.; Luo, X.; Gao, Y.; Li, S.; Zhu, J.; Tan, H. Simultaneous Contact and Grain-Boundary Passivation in Planar Perovskite Solar Cells Using SnO 2 -KCl Composite Electron Transport Layer. Adv. Energy Mater. 2019, 10, 1903083. [Google Scholar] [CrossRef]
  49. Bi, H.; Zuo, X.; Liu, B.; He, D.; Bai, L.; Wang, W.; Li, X.; Xiao, Z.; Sun, K.; Song, Q.; et al. Multifunctional Organic Ammonium Salt-Modified SnO2 Nanoparticles toward Efficient and Stable Planar Perovskite Solar Cells. J. Mater. Chem. A 2021, 9, 3940. [Google Scholar] [CrossRef]
  50. Liu, Z.; Deng, K.; Hu, J.; Li, L. Coagulated SnO2 Colloids for High-Performance Planar Perovskite Solar Cells with Negligible Hysteresis and Improved Stability. Angew. Chem. 2019, 131, 11621–11628. [Google Scholar]
  51. Sun, Y.; Pang, Z.; Quan, Y.; Han, D.; Zhang, X.; Ge, X.; Wang, F.; Sun, Y.; Yang, J.; Yang, L. A Synchronous Defect Passivation Strategy for Constructing High-Performance and Stable Planar Perovskite Solar Cells. Chem. Eng. J. 2021, 413, 127387. [Google Scholar] [CrossRef]
  52. Jiang, E.; Ai, Y.; Yan, J.; Li, N.; Lin, L.; Wang, Z.; Shou, C.; Yan, B.; Zeng, Y.; Sheng, J.; et al. Phosphate-Passivated SnO2 Electron Transport Layer for High-Performance Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2019, 11, 36727–36734. [Google Scholar] [CrossRef] [PubMed]
  53. Huang, C.; Lin, P.; Fu, N.; Sun, K.; Ye, M.; Liu, C.; Zhou, X.; Shu, L.; Hao, X.; Xu, B.; et al. Ionic liquid modified SnO2 nanocrystals as a robust electron transporting layer for efficient planar perovskite solar cells. J. Mater. Chem. A 2018, 6, 22086–22095. [Google Scholar] [CrossRef]
  54. Akin, S. Boosting the efficiency and stability of perovskite solar cells through facile molecular engineering approaches. Sol. Energy 2020, 199, 136–142. [Google Scholar] [CrossRef]
  55. Ai, Y.; Liu, W.; Shou, C.; Yan, J.; Li, N.; Yang, Z.; Song, W.; Yan, B.; Sheng, J.; Ye, J. SnO2 surface defects tuned by (NH4)2S for high-efficiency perovskite solar cells. Sol. Energy 2019, 194, 541–547. [Google Scholar] [CrossRef]
  56. Wang, H.; Li, F.; Wang, P.; Sun, R.; Ma, W.; Chen, M.; Miao, W.; Liu, D.; Wang, T. Chlorinated Fullerene Dimers for Interfacial Engineering Toward Stable Planar Perovskite Solar Cells with 22.3% Efficiency. Adv. Energy Mater. 2020, 10, 2000615. [Google Scholar] [CrossRef]
  57. Wu, J.; Cui, Y.; Yu, B.; Liu, K.; Li, Y.; Li, H.; Shi, J.; Wu, H.; Luo, Y.; Li, D.; et al. A Simple Way to Simultaneously Release the Interface Stress and Realize the Inner Encapsulation for Highly Efficient and Stable Perovskite Solar Cells. Adv. Funct. Mater. 2019, 29, 1905336. [Google Scholar] [CrossRef]
  58. Xie, J.; Huang, K.; Yu, X.; Yang, Z.; Xiao, K.; Qiang, Y.; Zhu, X.; Xu, L.; Wang, P.; Cui, C.; et al. Enhanced Electronic Properties of SnO2 via Electron Transfer from Graphene Quantum Dots for Efficient Perovskite Solar Cells. ACS Nano 2017, 11, 9176–9182. [Google Scholar] [CrossRef]
  59. Zheng, X.; Lei, H.; Yang, G.; Ke, W.; Chen, Z.; Chen, C.; Ma, J.; Guo, Q.; Yao, F.; Zhang, Q.; et al. Enhancing efficiency and stability of perovskite solar cells via a high mobility p-type PbS buffer layer. Nano Energy 2017, 38, 1–11. [Google Scholar] [CrossRef]
  60. Ke, W.; Zhao, D.; Xiao, C.; Wang, C.; Cimaroli, A.J.; Grice, C.R.; Yang, M.; Li, Z.; Jiang, C.-S.; Al-Jassim, M.; et al. Cooperative tin oxide fullerene electron selective layers for high-performance planar perovskite solar cells. J. Mater. Chem. A 2016, 4, 14276–14283. [Google Scholar] [CrossRef]
  61. Wang, C.; Zhao, D.; Grice, C.R.; Liao, W.; Yu, Y.; Cimaroli, A.; Shrestha, N.; Roland, P.J.; Chen, J.; Yu, Z.; et al. Low-temperature plasma-enhanced atomic layer deposition of tin oxide electron selective layers for highly efficient planar perovskite solar cells. J. Mater. Chem. A 2016, 4, 12080–12087. [Google Scholar] [CrossRef]
  62. Wang, J.; Datta, K.; Weijtens, C.H.L.; Wienk, M.M.; Janssen, R.A.J. Insights into Fullerene Passivation of SnO 2 Electron Transport Layers in Perovskite Solar Cells. Adv. Funct. Mater. 2019, 29, 1905883. [Google Scholar] [CrossRef]
  63. Li, Z.; Gao, Y.; Zhang, Z.; Xiong, Q.; Deng, L.; Li, X.; Zhou, Q.; Fang, Y.; Gao, P. cPCN-Regulated SnO2 Composites Enables Perovskite Solar Cell with Efficiency Beyond 23%. Nano-Micro Lett. 2021, 13, 101. [Google Scholar] [CrossRef] [PubMed]
  64. Niu, Y.; Tian, C.; Gao, J.; Fan, F.; Zhang, Y.; Mi, Y.; Ouyang, X.; Li, L.; Li, J.; Chen, S.; et al. Nb2C MXenes modified SnO2 as high quality electron transfer layer for efficient and stability perovskite solar cells. Nano Energy 2021, 89, 106455. [Google Scholar] [CrossRef]
  65. Hui, W.; Yang, Y.; Xu, Q.; Gu, H.; Feng, S.; Su, Z.; Zhang, M.; Wang, J.; Li, X.; Fang, J.; et al. Red-Carbon-Quantum-Dot-Doped SnO2 Composite with Enhanced Electron Mobility for Efficient and Stable Perovskite Solar Cells. Adv. Mater. 2020, 32, 1906374. [Google Scholar] [CrossRef] [PubMed]
  66. Dong, H.; Wang, J.; Li, X.; Liu, W.; Xia, T.; Yao, D.; Zhang, L.; Zuo, C.; Ding, L.; Long, F. Modifying SnO2 with Polyacrylamide to Enhance the Performance of Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2022, 14, 34143–34150. [Google Scholar] [CrossRef] [PubMed]
  67. Chen, J.; Dong, H.; Zhang, L.; Li, J.; Jia, F.; Jiao, B.; Xu, J.; Hou, X.; Liu, J.; Wu, Z. Graphitic carbon nitride doped SnO2 enabling efficient perovskite solar cells with PCEs exceeding 22%. J. Mater. Chem. A 2020, 8, 2644–2653. [Google Scholar] [CrossRef]
  68. Zhang, S.; Si, H.; Fan, W.; Shi, M.; Li, M.; Xu, C.; Zhang, Z.; Liao, Q.; Sattar, A.; Kang, Z.; et al. Graphdiyne: Bridging SnO2 and Perovskite in Planar Solar Cells. Angew. Chem. 2020, 132, 11670–11679. [Google Scholar] [CrossRef]
  69. Zhao, X.; Tao, L.; Li, H.; Huang, W.; Sun, P.; Liu, J.; Liu, S.; Sun, Q.; Cui, Z.; Sun, L.; et al. Efficient Planar Perovskite Solar Cells with Improved Fill Factor via Interface Engineering with Graphene. Nano Lett. 2018, 18, 2442–2449. [Google Scholar] [CrossRef] [PubMed]
  70. Wang, S.; Zhu, Y.; Liu, B.; Wang, C.; Ma, R. Introduction of carbon nanodots into SnO2 electron transport layer for efficient and UV stable planar perovskite solar cells. J. Mater. Chem. A 2019, 7, 5353–5362. [Google Scholar] [CrossRef]
  71. Lou, Q.; Han, Y.; Liu, C.; Zheng, K.; Zhang, J.; Chen, X.; Du, Q.; Chen, C.; Ge, Z. π-Conjugated Small Molecules Modified SnO2 Layer for Perovskite Solar Cells with over 23% Efficiency. Adv. Energy Mater. 2021, 11, 2101416. [Google Scholar] [CrossRef]
  72. Dai, Z.; Yadavalli, S.K.; Chen, M.; Abbaspourtamijani, A.; Qi, Y.; Padture, N.P. Interfacial toughening with self-assembled monolayers enhances perovskite solar cell reliability. Science 2021, 372, 618–622. [Google Scholar] [CrossRef]
  73. Tu, B.; Shao, Y.; Chen, W.; Wu, Y.; Li, X.; He, Y.; Li, J.; Liu, F.; Zhang, Z.; Lin, Y.; et al. Novel Molecular Doping Mechanism for n-Doping of SnO2 via Triphenylphosphine Oxide and Its Effect on Perovskite Solar Cells. Adv. Mater. 2019, 31, 1805944. [Google Scholar] [CrossRef]
  74. Zhang, J.; Yu, H. Reduced Energy Loss Enabled by Thiophene-Based Interlayers for High Performance and Stable Perovskite Solar Cells. J. Mater. Chem. A 2021, 9, 4138. [Google Scholar] [CrossRef]
  75. Xia, H.; Li, X.; Zhou, J.; Wang, B.; Chu, Y.; Li, Y.; Wu, G.; Zhang, D.; Xue, B.; Zhang, X.; et al. Interfacial Chemical Bridge Constructed by Zwitterionic Sulfamic Acid for Efficient and Stable Perovskite Solar Cells. ACS Appl. Energy Mater. 2020, 3, 3186–3192. [Google Scholar] [CrossRef]
  76. Sun, Y.; Zhang, J.; Yu, H.; Wang, J.; Huang, C.; Huang, J. Mechanism of bifunctional p-amino benzenesulfonic acid modified interface in perovskite solar cells. Chem. Eng. J. 2021, 420, 129579. [Google Scholar] [CrossRef]
  77. Yan, J.; Lin, Z.; Cai, Q.; Wen, X.; Mu, C. Choline Chloride-Modified SnO2 Achieving High Output Voltage in MAPbI3 Perovskite Solar Cells. ACS Appl. Energy Mater. 2020, 3, 3504–3511. [Google Scholar] [CrossRef]
  78. Zuo, L.; Chen, Q.; De Marco, N.; Hsieh, Y.-T.; Chen, H.; Sun, P.; Chang, S.-Y.; Zhao, H.; Dong, S.; Yang, Y. Tailoring the Interfacial Chemical Interaction for High-Efficiency Perovskite Solar Cells. Nano Lett. 2016, 17, 269–275. [Google Scholar] [CrossRef] [PubMed]
  79. Wang, Z.; Kamarudin, M.A.; Huey, N.C.; Yang, F.; Pandey, M.; Kapil, G.; Ma, T.; Hayase, S. Interfacial Sulfur Functionalization Anchoring SnO2 and CH3NH3PbI3 for Enhanced Stability and Trap Passivation in Perovskite Solar Cells. ChemSusChem 2018, 11, 3941–3948. [Google Scholar] [CrossRef]
  80. Xiong, Z.; Lan, L.; Wang, Y.; Lu, C.; Qin, S.; Chen, S.; Zhou, L.; Zhu, C.; Li, S.; Meng, L.; et al. Multifunctional Polymer Framework Modified SnO2 Enabling a Photostable α-FAPbI3 Perovskite Solar Cell with Efficiency Exceeding 23%. ACS Energy Lett. 2021, 6, 3824–3830. [Google Scholar] [CrossRef]
  81. Luan, Y.; Yi, X.; Mao, P.; Wei, Y.; Zhuang, J.; Chen, N.; Lin, T.; Li, C.; Wang, J. High-Performance Planar Perovskite Solar Cells with Negligible Hysteresis Using 2,2,2-Trifluoroethanol-Incorporated SnO2. iScience 2019, 16, 433–441. [Google Scholar] [CrossRef]
  82. Yang, D.; Yang, R.; Wang, K.; Wu, C.; Zhu, X.; Feng, J.; Ren, X.; Fang, G.; Priya, S.; Liu, S.F. High Efficiency Planar-Type Perovskite Solar Cells with Negligible Hysteresis using EDTA-Complexed SnO2. Nat. Commun. 2018, 9, 3239. [Google Scholar] [CrossRef] [PubMed]
  83. Wei, J.; Guo, F.; Wang, X.; Xu, K.; Lei, M.; Liang, Y.; Zhao, Y.; Xu, D. SnO2-in-Polymer Matrix for High-Efficiency Perovskite Solar Cells with Improved Reproducibility and Stability. Adv. Mater. 2018, 30, 1805153. [Google Scholar] [CrossRef] [PubMed]
  84. Huang, X.; Du, J.; Guo, X.; Lin, Z.; Ma, J.; Su, J.; Feng, L.; Zhang, C.; Zhang, J.; Chang, J.; et al. Polyelectrolyte-Doped SnO2 as a Tunable Electron Transport Layer for High-Efficiency and Stable Perovskite Solar Cells. Sol. RRL 2020, 4, 1900336. [Google Scholar] [CrossRef]
  85. Hwang, S.; Larina, L.; Lee, H.; Kim, S.; Choi, K.S.; Jeon, C.; Ahn, B.T.; Shin, B. Wet Pretreatment-Induced Modification of Cu(In,Ga)Se2/Cd-Free ZnTiO Buffer Interface. ACS Appl. Mater. Interfaces 2018, 10, 20920–20928. [Google Scholar] [CrossRef] [PubMed]
  86. Park, H.H.; Heasley, R.; Sun, L.; Steinmann, V.; Jaramillo, R.; Hartman, K.; Chakraborty, R.; Sinsermsuksakul, P.; Chua, D.; Buonassisi, T.; et al. Co-optimization of SnS absorber and Zn(O,S) buffer materials for improved solar cells. Prog. Photovolt. Res. Appl. 2015, 23, 901–908. [Google Scholar] [CrossRef] [Green Version]
  87. Kim, D.; Jung, H.J.; Park, I.J.; Larson, B.W.; Dunfield, S.P.; Xiao, C.; Kim, J.; Tong, J.; Boonmongkolras, P.; Ji, S.G.; et al. Efficient, stable silicon tandem cells enabled by anion-engineered wide-bandgap perovskites. Science 2020, 368, 155–160. [Google Scholar] [CrossRef] [PubMed]
  88. Park, H.H.; Kim, J.; Kim, G.; Jung, H.; Kim, S.; Moon, C.S.; Lee, S.J.; Shin, S.S.; Hao, X.; Yun, J.S.; et al. Transparent Electrodes Consisting of a Surface-Treated Buffer Layer Based on Tungsten Oxide for Semitransparent Perovskite Solar Cells and Four-Terminal Tandem Applications. Small Methods 2020, 4, 2000074. [Google Scholar] [CrossRef]
  89. Jung, H.; Kim, G.; Jang, G.S.; Lim, J.; Kim, M.; Moon, C.S.; Hao, X.; Jeon, N.J.; Yun, J.S.; Park, H.H.; et al. Transparent Electrodes with Enhanced Infrared Transmittance for Semitransparent and Four-Terminal Tandem Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2021, 13, 30497–30503. [Google Scholar] [CrossRef]
  90. Park, H.H. Transparent Electrode Techniques for Semitransparent and Tandem Perovskite Solar Cells. Electron. Mater. Lett. 2021, 17, 18–32. [Google Scholar] [CrossRef]
  91. Eom, T.; Kim, S.; Agbenyeke, R.E.; Jung, H.; Shin, S.M.; Lee, Y.K.; Kim, C.G.; Chung, T.; Jeon, N.J.; Park, H.H.; et al. Copper Oxide Buffer Layers by Pulsed-Chemical Vapor Deposition for Semitransparent Perovskite Solar Cells. Adv. Mater. Interfaces 2020, 8, 2001482. [Google Scholar] [CrossRef]
  92. Park, H.H.; Song, S.; Seo, J. Perspective: Approaches for layers above the absorber in perovskite solar cells for semitransparent and tandem applications. Mater. Today Energy 2021, 21, 100729. [Google Scholar] [CrossRef]
Figure 1. (a) Illuminated current density vs. voltage (J-V) curves under reverse and forward voltage scanning of PSC devices, (b) photoluminescence (PL) spectra, and (c) time-resolved PL (TRPL) spectra of perovskite films on undoped and Ta-doped SnO2 ETLs. Reproduced from [22] with permission from AIP Publishing, 2019. (d) Illuminated J-V curves under 1 SUN under reverse and forward voltage scanning of PSC devices, (e) photoluminescence (PL) spectra, and (f) TRPL spectra of perovskite films on undoped and bilayer F-doped SnO2 ETLs. Reproduced from [23] with permission from American Chemical Society (ACS) Publications, 2018. (g) Energy diagram of FTO, ETLs, and perovskite. Illuminated J-V curve of PSC based on (h) undoped SnO2 and (i) Li-doped SnO2 under 1 SUN. Reproduced from [24] with permission from Elsevier, 2016.
Figure 1. (a) Illuminated current density vs. voltage (J-V) curves under reverse and forward voltage scanning of PSC devices, (b) photoluminescence (PL) spectra, and (c) time-resolved PL (TRPL) spectra of perovskite films on undoped and Ta-doped SnO2 ETLs. Reproduced from [22] with permission from AIP Publishing, 2019. (d) Illuminated J-V curves under 1 SUN under reverse and forward voltage scanning of PSC devices, (e) photoluminescence (PL) spectra, and (f) TRPL spectra of perovskite films on undoped and bilayer F-doped SnO2 ETLs. Reproduced from [23] with permission from American Chemical Society (ACS) Publications, 2018. (g) Energy diagram of FTO, ETLs, and perovskite. Illuminated J-V curve of PSC based on (h) undoped SnO2 and (i) Li-doped SnO2 under 1 SUN. Reproduced from [24] with permission from Elsevier, 2016.
Nanomaterials 12 04326 g001
Figure 2. (a) Energy-band diagram of PSC based on In2O3/SnO2. (b) Illuminated J-V curves under 1 SUN for forward- and reverse-voltage scans of In2O3, SnO2, and In2O3/SnO2-based PSC devices. (c) Long-term stability measurements of devices with different ETLs. Reproduced from [31] with permission from Wiley, 2020 (d) Energy-band diagram of each layer in PSCs. (e) Illuminated J-V curves under 1 SUN under reverse- and forward-voltage scanning of PSC devices with different ETLs. (f) Long-term stability measurements of PSCs with different ETLs in the ambient. Reproduced from [32] with permission from Elsevier, 2021. (g) Schematic illustration of bilayered TiO2/SnO2 ETL. (h) Illustration of TiO2/SnO2 ETLs with large change in free energy (ΔG) and moderate electron mobility (μe). (i) Illuminated J-V curves under 1 SUN for various ETLs. Reproduced from [33] with permission from ACS Publications, 2017.
Figure 2. (a) Energy-band diagram of PSC based on In2O3/SnO2. (b) Illuminated J-V curves under 1 SUN for forward- and reverse-voltage scans of In2O3, SnO2, and In2O3/SnO2-based PSC devices. (c) Long-term stability measurements of devices with different ETLs. Reproduced from [31] with permission from Wiley, 2020 (d) Energy-band diagram of each layer in PSCs. (e) Illuminated J-V curves under 1 SUN under reverse- and forward-voltage scanning of PSC devices with different ETLs. (f) Long-term stability measurements of PSCs with different ETLs in the ambient. Reproduced from [32] with permission from Elsevier, 2021. (g) Schematic illustration of bilayered TiO2/SnO2 ETL. (h) Illustration of TiO2/SnO2 ETLs with large change in free energy (ΔG) and moderate electron mobility (μe). (i) Illuminated J-V curves under 1 SUN for various ETLs. Reproduced from [33] with permission from ACS Publications, 2017.
Nanomaterials 12 04326 g002
Figure 3. (a) Illuminated J-V curves under 1 SUN for forward- and reverse-voltage scanning of SnO2 and SnO2/RbF-based PSC devices. (b) Steady-state PL and (c) TRPL spectra of perovskite films on various ETLs. Reproduced from [40] with permission from Wiley, 2021 (d) Illuminated J-V curves under 1 SUN for PSCs based on various ETLs. (e) Current vs. voltage (I-V) characteristic curves for various ETLs with an ITO/ETL/Ag structure. (f) Long-term maximum power point tracking (MPPT) stability measurements of devices with different ETLs under continuous 1 SUN illumination (simulated by LED light). Reproduced from [41] with permission from Elsevier, 2022. (g) Schematic illustration of formation of ImAcHCl-modified SnO2 ETL. (h) Energy-band diagram of each layer in PSCs. (i) Illuminated J-V curves under 1 SUN for various ETLs. Reproduced from [42] with permission from Wiley, 2019.
Figure 3. (a) Illuminated J-V curves under 1 SUN for forward- and reverse-voltage scanning of SnO2 and SnO2/RbF-based PSC devices. (b) Steady-state PL and (c) TRPL spectra of perovskite films on various ETLs. Reproduced from [40] with permission from Wiley, 2021 (d) Illuminated J-V curves under 1 SUN for PSCs based on various ETLs. (e) Current vs. voltage (I-V) characteristic curves for various ETLs with an ITO/ETL/Ag structure. (f) Long-term maximum power point tracking (MPPT) stability measurements of devices with different ETLs under continuous 1 SUN illumination (simulated by LED light). Reproduced from [41] with permission from Elsevier, 2022. (g) Schematic illustration of formation of ImAcHCl-modified SnO2 ETL. (h) Energy-band diagram of each layer in PSCs. (i) Illuminated J-V curves under 1 SUN for various ETLs. Reproduced from [42] with permission from Wiley, 2019.
Nanomaterials 12 04326 g003
Figure 4. (a) Energy-band levels of each layer in the PSC. (b) Illuminated J-V curves under 1 SUN for reverse- and forward-voltage scanning of SnO2 and CoCl2:SnO2-based PSC devices. (c) Stability measurements of devices with various ETLs in N2 atmosphere under continuous irradiation. Reproduced from [46] with permission from ACS Publications, 2021. (d) Illustration of incorporation of heparin potassium (HP) into the SnO2 nanocrystal dispersions, resulting in arrangements of ETL nanocrystals, and crystal growth of the perovskite films with and without heparin potassium. (e) Illuminated J-V curves under 1 SUN for reverse- and forward-voltage scans of PSCs with and without HP incorporation into SnO2. (f) Long-term MPPT-stability measurements of devices with different ETLs for 1000 h under continuous 1 SUN illumination. Reproduced from [47] with permission from Wiley, 2020. (g) Illuminated J-V curves under 1 SUN for various ETLs. (h) Steady state PL and (i) TRPL spectra for various ETLs. Reproduced from [48] with permission from Wiley, 2020.
Figure 4. (a) Energy-band levels of each layer in the PSC. (b) Illuminated J-V curves under 1 SUN for reverse- and forward-voltage scanning of SnO2 and CoCl2:SnO2-based PSC devices. (c) Stability measurements of devices with various ETLs in N2 atmosphere under continuous irradiation. Reproduced from [46] with permission from ACS Publications, 2021. (d) Illustration of incorporation of heparin potassium (HP) into the SnO2 nanocrystal dispersions, resulting in arrangements of ETL nanocrystals, and crystal growth of the perovskite films with and without heparin potassium. (e) Illuminated J-V curves under 1 SUN for reverse- and forward-voltage scans of PSCs with and without HP incorporation into SnO2. (f) Long-term MPPT-stability measurements of devices with different ETLs for 1000 h under continuous 1 SUN illumination. Reproduced from [47] with permission from Wiley, 2020. (g) Illuminated J-V curves under 1 SUN for various ETLs. (h) Steady state PL and (i) TRPL spectra for various ETLs. Reproduced from [48] with permission from Wiley, 2020.
Nanomaterials 12 04326 g004aNanomaterials 12 04326 g004b
Figure 5. (a) Chemical structures of fullerene dimers and monomers. Illuminated J-V curves under 1 SUN for reverse- and forward-voltage scans of (b) SnO2 and SnO2/PCBM-based, and (c) SnO2/NMBF-Cl and SnO2/NMBF-H monomer and dimer-based PSC devices. Reproduced from [56] with permission from Wiley, 2020. (d) Schematic of PSC-device stack based on polystyrene (PS) modified SnO2 ETL. (e) Stability tests at the MPP under 1 SUN illumination. (f) Long-term stability measurements of devices with inner-PS encapsulation stored in air. Reproduced from [57] with permission from Wiley, 2019. (g) Cross-sectional scanning electron microscopy (SEM) image of device with SnO2/GQDs ETL. (h) Energy-band levels of device based on SnO2/GQDs before and after illumination. (i) Illuminated J-V curves under 1 SUN for various ETLs. Reproduced from [58] with permission from ACS Publishing, 2017.
Figure 5. (a) Chemical structures of fullerene dimers and monomers. Illuminated J-V curves under 1 SUN for reverse- and forward-voltage scans of (b) SnO2 and SnO2/PCBM-based, and (c) SnO2/NMBF-Cl and SnO2/NMBF-H monomer and dimer-based PSC devices. Reproduced from [56] with permission from Wiley, 2020. (d) Schematic of PSC-device stack based on polystyrene (PS) modified SnO2 ETL. (e) Stability tests at the MPP under 1 SUN illumination. (f) Long-term stability measurements of devices with inner-PS encapsulation stored in air. Reproduced from [57] with permission from Wiley, 2019. (g) Cross-sectional scanning electron microscopy (SEM) image of device with SnO2/GQDs ETL. (h) Energy-band levels of device based on SnO2/GQDs before and after illumination. (i) Illuminated J-V curves under 1 SUN for various ETLs. Reproduced from [58] with permission from ACS Publishing, 2017.
Nanomaterials 12 04326 g005
Figure 6. (a) Energy-band level of each layer in the PSC. (b) Illuminated J-V curves under 1 SUN for SnO2 and cPCN:SnO2-based devices. (c) Long-term stability measurements of unencapsulated devices in N2. Reproduced from [63] with permission from Springer, 2021. (d) PL spectra of perovskite film grown on various ETLs. (e) Illuminated J-V curves under 1 SUN for SnO2 and Nb2C:SnO2-based devices. (f) Long-term stability measurements of devices with different ETLs. Reproduced from [64] with permission from Elsevier, 2021. (g) Illuminated J-V curves under 1 SUN for reverse and forward scans of SnO2 and carbon quantum dot:SnO2-based devices. (h) Long-term stability measurements of unencapsulated devices with different ETLs in ambient dark environment at 25 °C with 40–60% relative humidity. (i) Steady-state PL spectra of perovskite on varying ETLs. Reproduced from [65] with permission from Wiley, 2020.
Figure 6. (a) Energy-band level of each layer in the PSC. (b) Illuminated J-V curves under 1 SUN for SnO2 and cPCN:SnO2-based devices. (c) Long-term stability measurements of unencapsulated devices in N2. Reproduced from [63] with permission from Springer, 2021. (d) PL spectra of perovskite film grown on various ETLs. (e) Illuminated J-V curves under 1 SUN for SnO2 and Nb2C:SnO2-based devices. (f) Long-term stability measurements of devices with different ETLs. Reproduced from [64] with permission from Elsevier, 2021. (g) Illuminated J-V curves under 1 SUN for reverse and forward scans of SnO2 and carbon quantum dot:SnO2-based devices. (h) Long-term stability measurements of unencapsulated devices with different ETLs in ambient dark environment at 25 °C with 40–60% relative humidity. (i) Steady-state PL spectra of perovskite on varying ETLs. Reproduced from [65] with permission from Wiley, 2020.
Nanomaterials 12 04326 g006aNanomaterials 12 04326 g006b
Figure 7. (a) Illustration of mechanism of oxygen-vacancy-defect passivation of Y6 and BTAC4. (b) Energy-level diagrams of each layer in the PSC. (c) Illuminated J-V curves under 1 SUN for reverse and forward scans of SnO2, SnO2/Y6, and SnO2/BTAC4-based devices. Reproduced from [71] with permission from Wiley, 2021. (d) Schematic of toughness testing and toughness results with various ETLs. Inset shows illustration of idealized I-SAM. (e) Illuminated J-V curves under 1 SUN for reverse and forward scans of SnO2, SnO2/H-SAM, and SnO2/I-SAM based devices. (f) Long-term stability measurements of unencapsulated devices with various ETLs at the MPP in N2 atmosphere under continuous 1 SUN illumination. Reproduced from [72] with permission from American Association for the Advancement of Science (AAAS), 2021. (g) The relaxed model of a TPPO molecule absorbed on SnO2 (110) surface. Energy-level band diagrams of ETL/perovskite heterojunction (h) without and (i) with TPPO-surface treatment of SnO2. Reproduced from [73] with permission from Wiley, 2019.
Figure 7. (a) Illustration of mechanism of oxygen-vacancy-defect passivation of Y6 and BTAC4. (b) Energy-level diagrams of each layer in the PSC. (c) Illuminated J-V curves under 1 SUN for reverse and forward scans of SnO2, SnO2/Y6, and SnO2/BTAC4-based devices. Reproduced from [71] with permission from Wiley, 2021. (d) Schematic of toughness testing and toughness results with various ETLs. Inset shows illustration of idealized I-SAM. (e) Illuminated J-V curves under 1 SUN for reverse and forward scans of SnO2, SnO2/H-SAM, and SnO2/I-SAM based devices. (f) Long-term stability measurements of unencapsulated devices with various ETLs at the MPP in N2 atmosphere under continuous 1 SUN illumination. Reproduced from [72] with permission from American Association for the Advancement of Science (AAAS), 2021. (g) The relaxed model of a TPPO molecule absorbed on SnO2 (110) surface. Energy-level band diagrams of ETL/perovskite heterojunction (h) without and (i) with TPPO-surface treatment of SnO2. Reproduced from [73] with permission from Wiley, 2019.
Nanomaterials 12 04326 g007
Figure 8. (a) Molecular structure of PEGDA. Cross-sectional SEM images of the perovskite on (b) undoped SnO2 and (c) SnO2 doped with PEGDA. Reproduced from [80] with permission from ACS Publications, 2021. (d) Cross-sectional SEM image of PSC device with SnO2 doped with 2,2,2-trifluoroethanol (T-SnO2). (e) Illuminated J-V curves under 1 SUN for forward and reverse scans of SnO2 and T-SnO2 based devices. (f) Long-term stability measurements of unencapsulated devices in air with 30–40% relative humidity for SnO2, T-SnO2, and oxygen plasma-treated T-SnO2 (p-T-SnO2) as ETLs. Reproduced from [81] with permission from Cell Press, 2019. (g) Fermi levels of EDTA, SnO2, and EDTA:SnO2 relative to the conduction band of the perovskite layer. (h) Illuminated J-V curves of PSC devices with various ETLs. (i) Long-term stability measurements of unencapsulated devices with different ETLs under 1 SUN illumination. Reproduced from [82] with permission from Springer Nature, 2018.
Figure 8. (a) Molecular structure of PEGDA. Cross-sectional SEM images of the perovskite on (b) undoped SnO2 and (c) SnO2 doped with PEGDA. Reproduced from [80] with permission from ACS Publications, 2021. (d) Cross-sectional SEM image of PSC device with SnO2 doped with 2,2,2-trifluoroethanol (T-SnO2). (e) Illuminated J-V curves under 1 SUN for forward and reverse scans of SnO2 and T-SnO2 based devices. (f) Long-term stability measurements of unencapsulated devices in air with 30–40% relative humidity for SnO2, T-SnO2, and oxygen plasma-treated T-SnO2 (p-T-SnO2) as ETLs. Reproduced from [81] with permission from Cell Press, 2019. (g) Fermi levels of EDTA, SnO2, and EDTA:SnO2 relative to the conduction band of the perovskite layer. (h) Illuminated J-V curves of PSC devices with various ETLs. (i) Long-term stability measurements of unencapsulated devices with different ETLs under 1 SUN illumination. Reproduced from [82] with permission from Springer Nature, 2018.
Nanomaterials 12 04326 g008aNanomaterials 12 04326 g008b
Table 1. Summary of perovskite-solar cells based on elemental doping of SnO2 ETL.
Table 1. Summary of perovskite-solar cells based on elemental doping of SnO2 ETL.
ETLMethodDevice StackJSC
(mA/cm2)
VOC
(V)
FF
(%)
η
(%)
Institute, Year [Ref.]
Ta:SnO2Chemical Bath DepositionITO/ETL/MAPbI3/Spiro-OMeTAD/Au21.7 → 22.81.16 → 1.1677.7 → 78.619.5 → 20.8Fudan, 2019 [22]
Bilayer F:SnO2Spin CoatingFTO/ETL/(FAPbI3)0.85(MAPbBr3)0.15/Spiro-OMeTAD/Au21.7 → 22.91.03 → 1.1372.3 → 78.116.3 → 20.2Huazhong UST, 2018 [23]
Li:SnO2Spin CoatingITO/ETL/MAPbI3/Spiro-OMeTAD/Au22.0 → 23.31.08 → 1.1164.2 → 70.715.3 → 18.2KIST, 2016 [24]
Nb:SnO2Spin CoatingFTO/ETL/(FAPbI3)0.85(MAPbBr3)0.15/Spiro-OMeTAD/Au21.7 → 22.41.06 → 1.0865.9 → 72.715.1 → 17.6Shaanxi Normal U., 2017 [25]
Y:SnO2Hydrothermal GrowthFTO/ETL/MAPbI3/Spiro-OMeTAD/Au19.3 → 22.61.05 → 1.0866.0 → 71.013.4 → 17.3Wuhan U. & Toledo, 2017 [26]
Sb:SnO2Spin CoatingITO/ETL/MAPbI3/Spiro-OMeTAD/Au22.3 → 22.61.01 → 1.0669.6 → 72.015.7 → 17.2UNL, 2016 [27]
Ga:SnO2Spin CoatingAZO/ETL/CsFAMAPb(Br,I)3/Spiro-OMeTAD/Au22.0 → 22.81.00 → 1.0757.0 → 70.012.5 → 17.0Adolphe Merkle Inst. & HZB, 2018 [28]
Mg:SnO2Spin CoatingFTO/ETL/MAPbI3/Spiro-OMeTAD/Au17.4 → 21.40.94 → 1.0050.0 → 70.8 8.2 → 15.2Wuhan U., 2016 [29]
Al:SnO2Spin CoatingFTO/ETL/MAPbI3/Spiro-OMeTAD/Au16.8 → 19.41.00 → 1.0353.0 → 58.0 9.0 → 12.1UESTC, 2017 [30]
Table 2. Summary of perovskite-solar cells based on metal oxide modified SnO2 ETL.
Table 2. Summary of perovskite-solar cells based on metal oxide modified SnO2 ETL.
ETLMethodDevice StackJSC
(mA/cm2)
VOC
(V)
FF
(%)
η
(%)
Institute, Year [Ref.]
In2O3/SnO2Spin CoatingITO/ETL/MAFAPbICl/spiro-OMeTAD/Au24.3 → 24.51.13 → 1.1678.1 → 81.221.4 → 23.1Nankai, 2020 [31]
SnO2/Eu:WOxSpin CoatingFTO/ETL/CsFAMAPbIBr/spiro-OMeTAD/Eu:WOx/Au23.3 → 24.01.11 → 1.1673.4 → 79.419.0 → 22.1Jilin, 2021 [32]
TiO2/SnO2Potentiostatic Anodization, Spin CoatingFTO/ETL/MAFAPbBrI/spiro-OMeTAD/Ag22.1 → 22.91.14 → 1.2075.4 → 76.419.0 → 21.1Toronto, POSTECH, 2017 [33]
TiO2/SnO2Spray Pyrolysis, Spin CoatingFTO/ETL/MAPbI3/PTAA/Au21.5 → 22.61.10 → 1.1369.0 → 78.016.4 → 19.8EPFL, 2017 [34]
SnO2/MgOSpin CoatingITO/ETL/MAPbI3/spiro-OMeTAD/Au21.3 → 22.11.10 → 1.1364.9 → 75.715.2 → 19.0RTV, 2018 [35]
PbO:SnO2/ SnO2Spin CoatingFTO/ETL/MAPbI3/spiro-OMeTAD/Au20.9 → 22.61.11 → 1.1072.9 → 75.517.0 → 18.8CAS, 2021 [36]
MgO/SnO2Spin CoatingFTO/ETL/MAPbI3/spiro-OMeTAD/Au21.6 → 22.71.07 → 1.1071.0 → 73.016.4 → 18.2Wuhan, 2017 [37]
SnO2/TiO2Chemical Bath DepositionFTO/ETL/MAPbI3/spiro-OMeTAD/Ag22.2 → 22.50.97 → 1.0156.0 → 79.012.0 → 18.1ECUST, Griffith, 2017 [38]
SnO2/ZnOSpin CoatingITO/ETL/CsPbI2Br/spiro-OMeTAD/MoO3/Ag14.7 → 15.01.06 → 1.2375.7 → 78.811.9 → 14.6South China UT, 2018 [39]
Table 3. Summary of perovskite-solar cells based on surface modification of SnO2 ETL by ionic compounds.
Table 3. Summary of perovskite-solar cells based on surface modification of SnO2 ETL by ionic compounds.
ETLMethodDevice StackJSC
(mA/cm2)
VOC
(V)
FF
(%)
η
(%)
Institute, Year [Ref.]
SnO2/RbFSpin CoatingITO/ETL/CsMAFAPbIBr/CH3O-PEAI/Spiro-OMeTAD/Au24.2 → 24.31.20 → 1.2177.4 → 79.322.4 → 23.4CAS, 2021 [40]
p-SnO2:RbF/m-SnO2/RbFSpin CoatingITO/ETL/CsMAFAPbIBrCl/Spiro-OMeTAD/Ag23.7 → 24.51.11 → 1.1577.8 → 82.120.6 → 22.7Southwest Petrolium, 2022 [41]
SnO2/4-Imidazoleacetic acid hydrochloride (ImAcHCl)Spin CoatingFTO/ETL/MA0.05FA0.95Pb(I0.95Br0.05)3/Spiro-OMeTAD/Au22.7 → 23.11.09 → 1.1579.0 → 79.019.5 → 21.0SKKU, 2019 [42]
SnO2/KOHCBD, Spin Coating FTO/ETL/Cs0.05(FA0.85MA0.15)0.95Pb(I0.85Br0.15)3/Spiro-OMeTAD/Au22.5 → 22.61.10 → 1.1578.0 → 79.019.3 → 20.5Wuhan UT, 2018 [43]
SnO2/Cs2CO3Spin CoatingFTO/ETL/Cs0.05(MA0.17FA0.83)0.95Pb(I0.83Br0.17)3/Spiro-OMeTAD/Au20.7 → 23.31.14 → 1.1766.2 → 71.415.6 → 19.5Shanghai UES, 2021 [44]
SnO2/KClSpin CoatingITO/ETL/MAPbI3/Spiro-OMeTAD/Au22.0 → 22.71.05 → 1.0677.3 → 77.617.8 → 18.7UST Beijing, 2020 [45]
SnO2/NaClSpin CoatingITO/ETL/MAPbI3/Spiro-OMeTAD/Au22.0 → 22.11.05 → 1.0677.3 → 79.017.8 → 18.5UST Beijing, 2020 [45]
SnO2/LiClSpin CoatingITO/ETL/MAPbI3/Spiro-OMeTAD/Au22.0 → 22.51.05 → 1.0677.3 → 76.917.8 → 18.3UST Beijing, 2020 [45]
SnO2/RbClSpin CoatingITO/ETL/MAPbI3/Spiro-OMeTAD/Au22.0 → 22.21.05 → 1.0477.3 → 77.417.8 → 17.9UST Beijing, 2020 [45]
SnO2/CsClSpin CoatingITO/ETL/MAPbI3/Spiro-OMeTAD/Au22.0 → 22.11.05 → 1.0577.3 → 76.117.8 → 17.7UST Beijing, 2020 [45]
Table 4. Summary of perovskite-solar cells based on bulk incorporation of ionic compounds into SnO2 ETL.
Table 4. Summary of perovskite-solar cells based on bulk incorporation of ionic compounds into SnO2 ETL.
ETLMethodDevice StackJSC
(mA/cm2)
VOC
(V)
FF
(%)
η
(%)
Institute, Year [Ref.]
CoCl2:SnO2Spin CoatingITO/ETL/MAFAPbIBrCl/Spiro-OMeTAD/Au24.2 → 24.61.16 → 1.2078.9 → 80.822.2 → 23.8Nankai, 2021 [46]
Heparin potassium:SnO2Spin CoatingFTO/ETL/Cs0.05MA0.10FA0.85Pb(I0.97Br0.03)3/Spiro-OMeTAD/Au24.3 → 25.01.13 → 1.1675.4 → 79.420.7 → 23.1Huazhong UST, 2020 [47]
KCl:SnO2Spin CoatingITO/ETL/MAFAPbIBr/Spiro-OMeTAD/Au24.0 → 24.21.08 → 1.1477.9 → 80.720.2 → 22.2Nanjing, 2020 [48]
Girard’s Reagent T (GRT):SnO2Spin CoatingITO/ETL/Rb0.05(FA0.95MA0.05)0.95PbI2.85Br0.15/Spiro-OMeTAD/Au22.6 → 22.91.08 → 1.1581.2 → 82.319.8 → 21.6Chongqing, 2021 [49]
NH4Cl:SnO2Spin CoatingITO/ETL/MAFAPbIBr/Spiro-OMeTAD/Au23.2 → 24.31.10 → 1.1573.5 → 76.818.7 → 21.4Soochow, 2019 [50]
Potassium sodium tartrate (PSTA):SnO2Spin CoatingITO/ETL/MA0.85FA0.15PbI3/Spiro-OMeTAD/Ag23.5 → 24.51.08 → 1.1271.9 → 76.918.3 → 21.1Jilin Normal, 2021 [51]
Phosphoric acid:SnO2Spin CoatingITO/ETL/MA0.15FA0.85Pb(I0.85Br0.15)3/Spiro-OMeTAD/Ag22.5 → 23.21.18 → 1.1773.8 → 77.419.7 → 21.0CAS, 2019 [52]
Tetramethylammonium hydroxide (TMAH):SnO2Spin CoatingFTO/ETL/MA0.25FA0.75PbI2.5Br0.5/Spiro-OMeTAD/Au22.8 → 23.31.13 → 1.1470.4 → 77.418.1 → 20.5Shenzhen, 2018 [53]
CsF:SnO2Spin CoatingFTO/ETL/Cs0.05(MA0.15FA0.85)0.95Pb(I0.85Br0.15)3/4-tert-butyl-D-phenylalanine (D4TBP)/Spiro-OMeTAD/Au22.6 → 23.21.13 → 1.1675.0 → 76.019.3 → 20.5KMU, 2020 [54]
(NH4)2S:SnO2Spin CoatingITO/ETL/MAFAPbIBr/Spiro-OMeTAD/Ag22.4 → 23.01.13 → 1.1573.4 → 76.018.7 → 20.0CAS, 2019 [55]
Table 5. Summary of perovskite-solar cells based on surface modification of SnO2 ETL by carbon materials.
Table 5. Summary of perovskite-solar cells based on surface modification of SnO2 ETL by carbon materials.
ETLMethodDevice StackJSC
(mA/cm2)
VOC
(V)
FF
(%)
η
(%)
Institute, Year [Ref.]
SnO2/fulleropyrrolidine (NMBF-Cl)Spin CoatingITO/ETL/MAFAPbIBr/Spiro-OMeTAD/Ag25.2 → 26.01.13 → 1.1275.0 → 77.021.4 → 22.3Wuhan UT, 2020 [56]
SnO2/Polystyrene (PS)Spin CoatingITO/ETL/MAFAPbIBr/PS/Spiro-OMeTAD/Au23.8 → 24.01.09 → 1.1074.0 → 76.019.3 → 20.5CAS, 2019 [57]
SnO2/Graphene quantum dotsSpin CoatingITO/ETL/MAPbI3/spiro-OMeTAD/Au22.1 → 23.11.10 → 1.1373.6 → 77.817.9 → 20.3Zhejiang, 2017 [58]
SnO2/[6,6]-phenyl-C61-butyric acid methyl ester (PCBM)Spin CoatingFTO/ETL/MAPbI3/spiro-OMeTAD/PbS/Au22.3 → 23.31.13 → 1.1475.0 → 74.018.8 → 19.6Wuhan, 2017 [59]
SnO2/PCBMSpin CoatingFTO/ETL/MAPbI3/spiro-OMeTAD/Au21.1 → 22.61.09 → 1.1271.5 → 75.816.5 → 19.1Toledo, Wuhan, 2016 [60]
SnO2/C60-SAMPEALD, Spin CoatingFTO/ETL/MAPbI3/spiro-OMeTAD/Au21.2 → 21.41.07 → 1.1375.5 → 79.117.2 → 19.0Toledo, 2016 [61]
SnO2/[6,6]-phenyl-C61-butyric acid (PCBA)Spin CoatingITO/ETL/MA0.34FA0.66PbI2.85Br0.15/Spiro-OMeTAD/MoO3/Au22.0 → 22.21.10 → 1.1064.0 → 76.015.4 → 18.6Eindhoven, 2019 [62]
Table 6. Summary of perovskite-solar cells based on bulk incorporation of carbon materials into SnO2 ETL.
Table 6. Summary of perovskite-solar cells based on bulk incorporation of carbon materials into SnO2 ETL.
ETLMethodDevice StackJSC
(mA/cm2)
VOC
(V)
FF
(%)
η
(%)
Institute, Year [Ref.]
Polymeric carbon nitrides (cPCN):SnO2Spin CoatingFTO/ETL/MAFAPbIBr/Spiro-OMeTAD/Ag23.4 → 24.91.11 → 1.1382.0 → 82.521.3 → 23.2CAS, 2021 [63]
Nb2C:SnO2Spin CoatingITO/ETL/Cs0.05MA0.07FA0.88PbI3/Spiro-OMeTAD/MoO3/Au24.7 → 25.31.11 → 1.1469.1 → 79.519.0 → 22.9China UPB, 2021 [64]
Carbon quantum dot:SnO2Spin CoatingITO/ETL/Cs0.05(MA0.17FA0.83)0.95Pb(I0.83Br0.17)3/Spiro-OMeTAD/MoO3/Au23.1 → 24.11.07 → 1.1477.8 → 82.919.2 → 22.8CAS, 2020 [65]
Polyacrylamide (PAM):SnO2Spin CoatingITO/ETL/MAFAPbIBrCl/Spiro-OMeTAD/Au23.2 → 24.81.10 → 1.1279.2 → 81.120.2 → 22.6Guilin UT, CAS, 2022 [66]
g-C3N4:SnO2Spin CoatingITO/ETL/CsMAFAPbIBr/Spiro-OMeTAD/Au23.7 → 24.01.11 → 1.1276.2 → 78.320.2 → 22.1Xian Jiaotong, 2020 [67]
Graphdiyne:SnO2Spin CoatingITO/ETL/CsMAFAPbIBr/Spiro-OMeTAD/Au22.9 → 23.31.13 → 1.1474.3 → 79.619.2 → 21.1UST Beijing, 2020 [68]
Naphthalene diimide graphene:SnO2Spin CoatingITO/ETL/MA0.17FA0.83PbI2.63Br0.37/Spiro-OMeTAD/Au23.2 → 22.71.10 → 1.0874.6 → 82.119.0 → 20.2Huazhong UST, 2018 [69]
Carbon nanodot:SnO2Spin CoatingITO/ETL/CsMAFAPbIBr/Spiro-OMeTAD/Au22.5 → 23.11.08 → 1.1076.0 → 79.018.5 → 20.0UST Beijing, 2019 [70]
Table 7. Summary of perovskite-solar cells based on surface modification of SnO2 ETL by organic molecules.
Table 7. Summary of perovskite-solar cells based on surface modification of SnO2 ETL by organic molecules.
ETLMethodDevice StackJSC
(mA/cm2)
VOC
(V)
FF
(%)
η
(%)
Institute, Year [Ref.]
KCl:SnO2/BTAC4Spin CoatingITO/ETL/Cs0.05(MA0.15FA0.85)0.95Pb(I0.85Br0.15)3/Spiro-OMeTAD/Ag23.1 → 24.21.23 → 1.2575.0 → 76.121.2 → 23.1CAS, 2021 [71]
KCl:SnO2/Y6Spin CoatingITO/ETL/Cs0.05(MA0.15FA0.85)0.95Pb(I0.85Br0.15)3/Spiro-OMeTAD/Ag23.1 → 23.61.23 → 1.2475.0 → 75.621.2 → 22.1CAS, 2021 [71]
SnO2/Si(OCH3)3(CH2)3I (I-SAM)Spin Coating, Immerse in solutionITO/ETL/Cs0.05(FA0.85MA0.15)0.95Pb(I0.85Br0.15)3/Spiro-OMeTAD/Au23.0 → 23.31.13 → 1.1977.4 → 77.820.2 → 21.4Brown, 2021 [72]
SnO2/Triphenylphosphine oxide (TPPO)Spin CoatingITO/ETL/Cs0.05(FA0.85MA0.15)0.95Pb(I0.85Br0.15)3/Spiro-OMeTAD/Au24.4 → 24.31.08 → 1.1172.2 → 77.019.0 → 20.7SUST, 2019 [73]
SnO2/Thiophene-3-acetic acidSpin CoatingITO/ETL/MAPbI3/spiro-OMeTAD/MoO3/Ag22.3 → 23.01.07 → 1.1273.5 → 80.117.5 → 20.6South China UT, 2021 [74]
SnO2/Aminosulfonic acidSpin Coating, Immerse in solutionITO/ETL/Cs0.05(FA0.85MA0.15)0.95Pb(I0.85Br0.15)3/Spiro-OMeTAD/Au21.8 → 22.81.12 → 1.1574.8 → 77.918.2 → 20.4Beihang, 2020 [75]
SnO2/p-amino benzenesulfonic acid (ABSA)Spin CoatingITO/ETL/MAPbI3/Spiro-OMeTAD/MoO3/Ag22.4 → 22.91.10 → 1.1373.0 → 78.818.0 → 20.3South China UT, 2021 [76]
SnO2/Choline chlorideSpin Coating, Immerse in solutionFTO/ETL/MAPbI3/Spiro-OMeTAD/Au21.3 → 22.81.07 → 1.1573.9 → 72.416.8 → 18.9Renmin U. China, 2020 [77]
SnO2/4-pyridinecarboxylic acid (PA-SAM)Spin CoatingITO/SnO2-SAM/MAPbI3/spiro-OMeTAD/Au21.7 → 22.01.06 → 1.1074.9 → 77.417.2 → 18.8UCLA, 2017 [78]
SnO2/Potassium hexylxanthateSpin CoatingITO/ETL/MAPbI3/Spiro-OMeTAD/Au21.7 → 22.61.03 → 1.0673.7 → 76.916.6 → 18.4Kyushu Tech., 2018 [79]
SnO2/4-cyanobenzoic acid (CBA-SAM)Spin CoatingITO/SnO2-SAM/MAPbI3/spiro-OMeTAD/Au21.7 → 21.71.06 → 1.0874.9 → 78.117.2 → 18.3UCLA, 2017 [78]
SnO2/benzoic acid (BA-SAM)Spin CoatingITO/SnO2-SAM/MAPbI3/spiro-OMeTAD/Au21.7 → 21.91.06 → 1.1174.9 → 74.617.2 → 18.1UCLA, 2017 [78]
Table 8. Summary of perovskite-solar cells based on bulk incorporation of organic molecules into SnO2 ETL.
Table 8. Summary of perovskite-solar cells based on bulk incorporation of organic molecules into SnO2 ETL.
ETLMethodDevice StackJSC
(mA/cm2)
VOC
(V)
FF
(%)
η
(%)
Institute, Year [Ref.]
Poly(ethylene glycol) diacrylate (PEGDA):SnO2Spin CoatingITO/ETL/FAPbI3/Spiro-OMeTAD/MoO3/Ag24.8 → 25.31.09 → 1.1480.6 → 81.021.8 → 23.3CAS, Chongqing, 2021 [80]
2,2,2-trifluoroethanol:SnO2, O2 plasmaSpin CoatingITO/ETL/MAFAPbIBr/Spiro-OMeTAD/Au23.1 → 23.9 → 24.11.10 → 1.12 → 1.1275.5 → 78.0 → 80.219.2 → 20.9 → 21.7CAS, 2019 [81]
Ethylene diamine tetraacetic acid (EDTA):SnO2Spin CoatingITO/ETL/Cs0.05FA0.95PbI3/Spiro-OMeTAD/Au22.8 → 24.61.10 → 1.1175.5 → 79.218.9 → 21.6Shaanxi Normal, 2018 [82]
Polyethylene glycol (PEG):SnO2Spin CoatingITO/ETL/Cs0.05FA0.81MA0.14Pb(I0.85Br0.15)3/Spiro-OMeTAD/Au22.6 → 22.71.09 → 1.1277.9 → 81.919.2 → 20.8Peking, 2018 [83]
Polyethylenimine (PEIE):SnO2Spin CoatingITO/ETL/CsMAFAPbICl/Spiro-OMeTAD/Ag22.9 → 23.81.08 → 1.1476.0 → 76.018.7 → 20.6Xidian, 2020 [84]
Table 9. Summary stability of ETL-modified perovskite-solar cells.
Table 9. Summary stability of ETL-modified perovskite-solar cells.
ETLDevice StackEncapsulatedConditionsContinuous 1 SUN Illumination?Durationη
Maintained
Institute, Year [Ref.]
F:SnO2FTO/bilayer F:SnO2/(FAPbI3)0.85(MAPbBr3)0.15/Spiro-OMeTAD/AuY40–50%, Air, RTN300 h>85%Huazhong UST, 2018 [23]
Nb:SnO2FTO/Nb:SnO2/(FAPbI3)0.85(MAPbBr3)0.15/Spiro-OMeTAD/AuNAir, RTN288 h90%Shaanxi Normal U., 2017 [25]
Sb:SnO2ITO/Sb:SnO2/MAPbI3/Spiro-OMeTAD/AuNDessicator, RTN504 h>95%UNL, 2016 [27]
Ga:SnO2AZO/Ga:SnO2/CsFAMAPb(Br,I)3/Spiro-OMeTAD/AuNN2, 1 SUNY1000 h~70%Adolphe Merkle Inst. & HZB, 2018 [28]
Mg:SnO2FTO/Mg:SnO2/MAPbI3/Spiro-OMeTAD/AuN<20%, AirN720 h>90%Wuhan U., 2016 [29]
In2O3/SnO2ITO/ETL/MAFAPbICl/spiro-OMeTAD/AuNN2N1920 h98%Nankai, 2020 [31]
N Y180 h91%
N75%N120 h~80%
SnO2/Eu:WOxFTO/ETL/CsFAMAPbIBr/spiro-OMeTAD/Eu:WOx/AuN16–25 °C, 20–30%Y500 h>90%Jilin, 2021 [32]
NAmbientN2000 h>90%
TiO2/SnO2FTO/TiO2-SnO2/MAPbI3/PTAA/AuN~20%, AirN1200 h>95%EPFL, 2017 [34]
SnO2/MgOITO/ETL/MAPbI3/spiro-OMeTAD/AuN30%N2568 h67%RTV, 2018 [35]
PbO:SnO2/SnO2FTO/ETL/MAPbI3/spiro-OMeTAD/Au RT, 15%N1080 h>90%CAS, 2021 [36]
SnO2/ZnOITO/ETL/CsPbI2Br/spiro-OMeTAD/MoO3/AgN85 °C, N2N300 h80%South China UT, 2018 [39]
SnO2/RbFITO/ETL/CsMAFAPbIBr/CH3O-PEAI/Spiro-OMeTAD/AuNWhite LED light illuminationN200 h~75%CAS, 2021 [40]
p-SnO2/RbF/m-SnO2/RbFITO/ETL/CsMAFAPbIBrCl/Spiro-OMeTAD/AgNMPPTY300 h90%Southwest Petrolium, 2022 [41]
SnO2/4-Imidazoleacetic acid hydrochloride (ImAcHCl)FTO/ETL/MA0.05FA0.95Pb(I0.95Br0.05)3/Spiro-OMeTAD/AuNRT, 46–60%N840 h94%SKKU, 2019 [42]
FTO/ETL/MA0.05FA0.95Pb(I0.95Br0.05)3/PTAA/AuN85 °C, N2N40 h90%SKKU, 2019 [42]
SnO2/Cs2CO3FTO/ETL/Cs0.05(MA0.17FA0.83)0.95Pb(I0.83Br0.17)3/Spiro-OMeTAD/AuN35–45%N340 h91%Shanghai UES, 2021 [44]
SnO2/NaClITO/ETL/MAPbI3/Spiro-OMeTAD/AuN N960 h>90%UST Beijing, 2020 [45]
CoCl2:SnO2ITO/ETL/MAFAPbIBrCl/Spiro-OMeTAD/AuNN2Y200 h84%Nankai, 2021 [46]
N60 °C, >50%N100 h80%
Heparin potassium:SnO2FTO/ETL/Cs0.05MA0.10FA0.85Pb(I0.97Br0.03)3/Spiro-OMeTAD/AuY60–65 °CY1000 h~97%Huazhong UST, 2020 [47]
KCl:SnO2ITO/ETL/MAFAPbIBr/Spiro-OMeTAD/AuN Y120 h88%Nanjing, 2020 [48]
Girard’s Reagent T (GRT):SnO2ITO/ETL/Rb0.05(FA0.95MA0.05)0.95PbI2.85Br0.15/Spiro-OMeTAD/AuN5–10%, RTN1440 h96%Chongqing, 2021 [49]
N60 °C, N2N720 h100%
NN2Y672 h59%
NH4Cl:SnO2ITO/ETL/MAFAPbIBr/Spiro-OMeTAD/AuNN2N1000 h>95%Soochow, 2019 [50]
Potassium sodium tartrate (PSTA):SnO2ITO/ETL/MA0.85FA0.15PbI3/Spiro-OMeTAD/AgN25 °C, 45%, AirN1440 h>95%Jilin Normal, 2021 [51]
Tetramethylammonium hydroxide (TMAH):SnO2FTO/ETL/MA0.25FA0.75PbI2.5Br0.5/Spiro-OMeTAD/AuY15%N360 h97%Shenzhen, 2018 [53]
CsF:SnO2FTO/ETL/Cs0.05(MA0.15FA0.85)0.95Pb(I0.85Br0.15)3/4-tert-butyl-D-phenylalanine (D4TBP)/Spiro-OMeTAD/AuNMPPTY800 h>90%KMU, 2020 [54]
SnO2/NMBF-ClITO/ETL/MAFAPbIBr/Spiro-OMeTAD/AgN25–35 °C, 45–60%, AirN1000 h>95%Wuhan UT, 2020 [56]
SnO2/Polystyrene (PS)ITO/ETL/MAFAPbIBr/PS/Spiro-OMeTAD/AuNAirN2800 h>90%CAS, 2019 [57]
N25 °C, 25%,MPPTY72 h>90%
SnO2/GQDsITO/ETL/MAPbI3/spiro-OMeTAD/AuNN2 → 20–30%, AirN2160 h>95%Zhejiang, 2017 [58]
SnO2/PCBMFTO/ETL/MAPbI3/spiro-OMeTAD/PbS/AuNAirN1000 h~100%Wuhan, 2017 [59]
SnO2/C60-SAMFTO/ETL/MAPbI3/spiro-OMeTAD/AuN<10%, room light, N2N480 h> 98%Toledo, 2016 [61]
Polymeric carbon nitrides (cPCN):SnO2FTO/ETL/MAFAPbIBr/Spiro-OMeTAD/AgNN2N2880 h95%CAS, 2021 [63]
N25–35%, AirN2000 h88%
Nb2C:SnO2ITO/ETL/Cs0.05MA0.07FA0.88PbI3/Spiro-OMeTAD/MoO3/AuN25 °C, 40–60%N960 h98%China UPB, 2021 [64]
Carbon quantum dot:SnO2ITO/ETL/Cs0.05(MA0.17FA0.83)0.95Pb(I0.83Br0.17)3/Spiro-OMeTAD/MoO3/AuN25 °C, 40–60%, AirN1000 h96%CAS, 2020 [65]
Polyacrylamide (PAM):SnO2ITO/ETL/MAFAPbIBrCl/Spiro-OMeTAD/AuN45–55%N1080 h90%Guilin UT, CAS, 2022 [66]
g-C3N4:SnO2ITO/ETL/CsMAFAPbIBr/Spiro-OMeTAD/AuN25 °C, 60%, AirN1500 h90%Xian Jiaotong, 2020 [67]
N85 °C, 60%, AirN75 h80%
Carbon nanodot:SnO2ITO/ETL/CsMAFAPbIBr/Spiro-OMeTAD/AuN20 °C, 20–30%, Air, UVN200 h90%UST Beijing, 2019 [70]
NDry AirN1200 h>90%
KCl:SnO2/BTAC4ITO/ETL/Cs0.05(MA0.15FA0.85)0.95Pb(I0.85Br0.15)3/Spiro-OMeTAD/AgN35%N768 h~90%CAS, 2021 [71]
KCl:SnO2/Y6 N35%N768 h~90%
SnO2/Si(OCH3)3(CH2)3I (I-SAM)ITO/ETL/Cs0.05(FA0.85MA0.15)0.95Pb(I0.85Br0.15)3/Spiro-OMeTAD/AuNMPPT, N2, RTY1200 h>90%Brown, 2021 [72]
SnO2/Thiophene-3-acetic acidITO/ETL/MAPbI3/spiro-OMeTAD/MoO3/AgNN2N1440 h>90%South China UT, 2021 [74]
N85 °C, 70%, AirN130 h>80%
SnO2/Aminosulfonic acidITO/ETL/Cs0.05(FA0.85MA0.15)0.95Pb(I0.85Br0.15)3/Spiro-OMeTAD/AuN25–35%, AirN1000 h>80%Beihang, 2020 [75]
N60 °C, N2N500 h>75%
SnO2/p-amino benzenesulfonic acid (ABSA)ITO/ETL/MAPbI3/Spiro-OMeTAD/MoO3/AgNN2N720 h57%South China UT, 2021 [76]
SnO2/Potassium hexylxanthateITO/ETL/MAPbI3/Spiro-OMeTAD/AuNRTN1680 h~90%Kyushu Tech., 2018 [79]
Poly(ethylene glycol) diacrylate (PEGDA):SnO2ITO/ETL/FAPbI3/Spiro-OMeTAD/MoO3/AgNN2Y850 h>90%CAS, Chongqing, 2021 [80]
N30–35%, AirN1000 h98%
2,2,2-trifluoroethanol:SnO2ITO/ETL/MAFAPbIBr/Spiro-OMeTAD/AuN30–40%N720 h>90%CAS, 2019 [81]
Ethylene diamine tetraacetic acid (EDTA):SnO2ITO/ETL/Cs0.05FA0.95PbI3/Spiro-OMeTAD/AuN35%N2880 h92%Shaanxi Normal, 2018 [82]
N1 SUNY120 h86%
Polyethylene glycol (PEG):SnO2ITO/ETL/Cs0.05FA0.81MA0.14Pb(I0.85Br0.15)3/Spiro-OMeTAD/AuN28–35 °C, 30–80%, AirN2160 h>97%Peking, 2018 [83]
Polyethylenimine (PEIE):SnO2ITO/ETL/MAFAPbI/Spiro-OMeTAD/AgN40%N1680 h82%Xidian, 2020 [84]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Park, H.H. Modification of SnO2 Electron Transport Layer in Perovskite Solar Cells. Nanomaterials 2022, 12, 4326. https://doi.org/10.3390/nano12234326

AMA Style

Park HH. Modification of SnO2 Electron Transport Layer in Perovskite Solar Cells. Nanomaterials. 2022; 12(23):4326. https://doi.org/10.3390/nano12234326

Chicago/Turabian Style

Park, Helen Hejin. 2022. "Modification of SnO2 Electron Transport Layer in Perovskite Solar Cells" Nanomaterials 12, no. 23: 4326. https://doi.org/10.3390/nano12234326

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

Article Metrics

Back to TopTop