Next Article in Journal
Numerical Simulation of the Effect of Wall-Equiaxed Crystal Density on the Number of Columnar Crystals and the Thickness of an Equiaxed Crystal Layer for Al-4.7%Cu Alloy Ingot Based on 3D LBM-CA Method
Next Article in Special Issue
Synthesis, Optical, Magnetic and Thermodynamic Properties of Rocksalt Li1.3Nb0.3Mn0.4O2 Cathode Material for Li-Ion Batteries
Previous Article in Journal
Fabrication of Superhydrophobic Ni-Co-BN Nanocomposite Coatings by Two-Step Jet Electrodeposition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Progress of Additive Engineering for CH3NH3PbI3 Photo-Active Layer in the Context of Perovskite Solar Cells

by
Mayuribala Mangrulkar
* and
Keith J. Stevenson
Center for Energy Science and Technology, Skolkovo Institute of Science and Technology, 121205 Moscow, Russia
*
Author to whom correspondence should be addressed.
Crystals 2021, 11(7), 814; https://doi.org/10.3390/cryst11070814
Submission received: 14 June 2021 / Revised: 7 July 2021 / Accepted: 9 July 2021 / Published: 13 July 2021
(This article belongs to the Special Issue Progress in Advanced Battery Materials)

Abstract

:
Methylammonium lead triiodide (CH3NH3PbI3/MAPbI3) is the most intensively explored perovskite light-absorbing material for hybrid organic–inorganic perovskite photovoltaics due to its unique optoelectronic properties and advantages. This includes tunable bandgap, a higher absorption coefficient than conventional materials used in photovoltaics, ease of manufacturing due to solution processability, and low fabrication costs. In addition, the MAPbI3 absorber layer provides one of the highest open-circuit voltages (Voc), low Voc loss/deficit, and low exciton binding energy, resulting in better charge transport with decent charge carrier mobilities and long diffusion lengths of charge carriers, making it a suitable candidate for photovoltaic applications. Unfortunately, MAPbI3 suffers from poor photochemical stability, which is the main problem to commercialize MAPbI3-based perovskite solar cells (PSCs). However, researchers frequently adopt additive engineering to overcome the issue of poor stability. Therefore, in this review, we have classified additives as organic and inorganic additives. Organic additives are subclassified based on functional groups associated with N/O/S donor atoms; whereas, inorganic additives are subcategorized as metals and non-metal halide salts. Further, we discussed their role and mechanism in terms of improving the performance and stability of MAPbI3-based PSCs. In addition, we scrutinized the additive influence on the morphology and optoelectronic properties to gain a deeper understanding of the crosslinking mechanism into the MAPbI3 framework. Our review aims to help the research community, by providing a glance of the advancement in additive engineering for the MAPbI3 light-absorbing layer, so that new additives can be designed and experimented with to overcome stability challenges. This, in turn, might pave the way for wide scale commercial use.

1. Introduction

Alternative renewable energy sources are considered reliable options for long-term usage due to the limited availability of traditional energy resources (i.e., coal, oil, and gas). Since solar energy is abundant and free (i.e., it comes from a natural source), it is referred to as clean and green energy. Also, solar energy can help reduce air pollution and global warming caused by greenhouse gas emissions if low-cost manufacturing is achieved [1]. This notion of utilizing solar energy further led to the development of photovoltaic technology, resulting in three generations. The first generation of photovoltaic technology was silicon wafer-based solar cells. However, they were not commercially successful due to their high cost. In contrast, second generation solar cells offered a low cost, but compromised in efficiency. Hence, the third generation was developed to be cost-effective and efficient. In this context, perovskite solar cells are considered advanced level third generation photovoltaic technology that could offer low-cost processing with ease of manufacturing and equally high efficiency when compared to their traditional counterparts [2]. This is why the research community has paid attention to perovskite solar cells in the last decade.
“Perovskite” is the name of the mineral CaTiO3 discovered in the Ural Mountains of Russia; it is named after the Russian nobleman and mineralogist, Lev Perovski. Nevertheless, this term is being used for all compounds, with the general formula ABX3, with the same crystal structure as CaTiO3 or derived from this structure. These materials consist of two cations. Cation A is 12-fold coordinated by anion X, and cation B is 6-fold, where X can be either oxygen or a halide. In terms of perovskite solar cells, the most commonly used A-site cations are MA/[CH3NH3]+, FA/[H2NCHNH2]+, Rb+, Cs+, B site cations are Pb2+, Sn2+, Ge2+. While the X site is halogen atoms, I, Cl, Br [3]. Among many different perovskite absorber layers, methyl ammonium lead iodide (MAPbI3) is the most prominently studied hybrid organic–inorganic perovskite for perovskite solar cell application. MAPbI3 offers several unique properties and advantages to be applicable in perovskite solar cells (PSC). It has a suitable bandgap~1.6 eV, a high absorption coefficient (the absorption coefficient of MAPbI3 lies in the range of 104–105 cm−1, which is more than one order of magnitude larger than that of silicon for the visible light spectrum). The high absorption coefficient of MAPbI3 allows it to absorb in the low light region. Moreover, a high absorption coefficient allows light to be absorbed by a thin film (generally in the range of 0.3–0.6 μm) of the perovskite layer, while crystalline silicon-based solar cells are usually made thicker ~300 μm. This, in turn, reduces the quantity of required material, thereby reducing the cost. Besides, MAPbI3 can be easily solution-processed to produce efficient solar cells, which further lowers manufacturing costs. Moreover, MAPbI3 can result in high open-circuit voltages (Voc), low Voc loss/deficit, and low exciton binding energy: 2–70 meV, resulting in better charge transport with decent charge carrier motilities of 2–66 cm2 V−1 s−1, long diffusion lengths of charge carriers ~1 um. Because of these properties, scientists and researchers have recommended the fabrication of perovskite solar cells using MAPbI3 as the active layer [4,5,6]. Nevertheless, long-term operational stability has been a significant factor that acts as a blockade for the commercialization of PSCs. In comparison, the traditionally made silicon-based solar cells remain operational for up to 20–25 years. On the contrary, perovskite solar cells are stable for a few hundred hours to a maximum of one year [4]. Thus, PSCs face serious stability issues compared to already available PV technology in the market. These stability issues involve both extrinsic and intrinsic challenges to overcome.
The perovskite-based active layer is susceptible to external factors, such as water, air, and moisture [7,8,9]. Reports have shown that, upon exposure to these external factors, PSC degrades and loses its operational stability [10]. It was found that, in the presence of water, MAPbI3 crystal structure forms hydrates/complex with H2O (Equation (1)), and deforms the perovskite structure, destroying optoelectronic properties, resulting in loss of photovoltaic performance [7,11,12].
4CH3NH3PbI3 + H2O ↔ 4CH3NH3PbI3·H2O ↔ (CH3NH3)4PbI6·2H2O + 3PbI2 + 2H2O
Moreover, it was demonstrated that exposure of CH3NH3PbI3 photoactive layers to light and oxygen results in the formation of superoxide (O2) species. This reactive O2 species can deprotonate the methyl ammonium cation (CH3NH3+), leading to the formation of PbI2, water, methylamine and iodine [13,14,15], as shown in Equation (2).
4CH3NH3PbI3 + O2 → 4CH3NH2 + 2I2 + 2H2O + 4PbI2
Furthermore, when MAPbI3 is exposed to moisture, methyl-ammonium cation degrades and results in PbI2, I2 and water formation (Equations (3)–(6)) [4], and the solar cell is destabilized.
CH3NH3PbI3(s) → CH3NH3I (aq) + PbI2 (s)
CH3NH3I (aq) → CH3NH2 (aq) + HI (aq)
4HI (aq) + O2 (g) → 2I2 (s) + 2H2O
2HI (aq) → H2 (g) + I2 (g)
Nevertheless, the impact of these external factors can still be minimized by encapsulating the active layer. However, it may affect performance and the cost. Ironically, MAPbI3 is intrinsically unstable as well. It means that, in the presence of light and heat, MAPbI3 tends to decompose. Sadly, these conditions are unavoidable for the operation of a solar cell.
When the temperature exceeds 85 °C in the dark, MAPbI3 forms ammonia or some other volatile species, and PbI2 (Equation (7)) [12,16].
CH3NH3PbI3 → PbI2 + CH3I + NH3
Whereas, in the presence of light and temperature, PbI2 intermediate forms and further decomposes, resulting in the formation of metallic lead and molecular iodine. At the same time, volatile species/by-product may rise from CH3NH3I. In the presence of light and heat, the identified degradation pathways are presented in Equations (8) and (9) [2,17,18,19,20,21,22,23]:
CH3NH3PbI3 → [PbI2] + volatile species (rising from CH3NH3I)
PbI2 → Pb2+ + 2I → Pb+ + I0 + I → (1/2)Pb0 + (1/2)PbI2 + (1/2)I2
Thus, these Pb and iodine sites create defect sites/trap states, resulting in instability of the CH3NH3PbI3 active layer and CH3NH3PbI3-based perovskite solar cell devices overall. These defect sites are often reported to be within the grains, at grain boundaries or at the surface of the CH3NH3PbI3 perovskite layer. Moreover, these trap states can also alter the energy level alignment and destroy charge transfer properties in CH3NH3PbI3 solar cell devices. Additionally, non-radiative recombination becomes more prominent in such a scenario, a clear indication of deterioration of device performance. Therefore, it is crucial/important to chemically passivate these defect sites to achieve longer stability of the CH3NH3PbI3 layer [14]. To achieve the aforementioned, additive engineering has been widely applied. Although, the passivation can also be achieved by physically passivating the defects with the help of encapsulating layer [24,25,26,27].
In this review, however, we will focus only on the additive approach. To accomplish this, we have analyzed/scrutinized additives used in the CH3NH3PbI3 photoactive layer. Further, the role of the additive and mechanism of the additive interaction with CH3NH3PbI3 has also been discussed. Their impact on film morphology, charge transport properties, and impact on stability has been intensively described. It should be noted that, here, we discussed only those additives that are added in the CH3NH3PbI3 layer. In addition, scrutiny of the physical passivation layers/encapsulation approach that may result in the same is beyond the scope of this review, as we mainly focused only on the aspect of additive engineering. Further, we aim to discuss those additives that do not alter/destroy the perovskite structure of CH3NH3PbI3, i.e., do not result in mixed cation or mixed halide perovskite. The additives are mainly classified as organic additives and inorganic salt additives. Furthermore, organic additives are analyzed based on the functional group associated with the N/O/S donor atom. At the same time, the inorganic salt additives are further classified as per the group name in the periodic table, i.e., alkali metals, transition metals, other metals, and non-metals. This review aims to analyze the research progress/advancement in the context of additives employed to improve the performance and stability of the CH3NH3PbI3 photoactive light-absorbing layer. This can further help design additives, ultimately resolving stability challenges to commercialize the CH3NH3PbI3-based perovskite solar cell.

2. Organic Additives

Organic additives are the most common types of additives used in perovskite solar cells. These organic additives can be further categorized based on N donor, O donor, and S donor atoms. These electron donor atoms can bind/coordinate with the Pb2+ species, resulting in adduct formation, passivating the grain boundaries and, thus, improving efficiency and stability. Subsequently, based on the electron-donating atoms, additives can be further subcategorized as amines, nitriles belonging to the N donor atom, amides, acids, acetates, alcohol, esters, and ethers, belonging to oxygen donor atoms and sulfides, thiocyanates belonging to the S donor atom [28].

2.1. N Donor Atom-Based Additives

2.1.1. Amine Additives

Following the notion of amine-based additives, the non-stoichiometric use of methylammonium iodide as an amine additive was advised in the early phases of developing perovskite solar cells. It was established that adding excess MAI reduces trap density, increases PL lifetime, and further increases PCE regardless of fabrication methods [29,30,31]. Likewise, some other organic amine derivatives, such as benzylammonium iodide (BAI) and phenethylammonium iodide (PEAI), have also been investigated as additives and were found to improve light harvesting properties and exciton lifetime with reduced charge recombination [32,33]. Moreover, it was discovered that, if hydrophobic cations containing amine additives are employed, such as hexylamine hydrochloride (1-HH), 1,6-diaminohexane dihydrochloride (1,6-DD) and phenylhydrazinium iodide (PHAI), it may even work as a protection for the MAPbI3 film against moisture, improving the stability in ambient conditions [34,35]. Later, it was found that, if instead of an iodide, a chloride-based counterpart (i.e. MACl instead of MAI) was used as an additive, perovskite films became more crystalline, with increased grain size and homogenous, smooth morphology, which further resulted in grain boundary passivation, resulting in better performance. Unfortunately, the addition of MACl was reported to form mixed halide perovskite CH3NH3PbI3−xClx [36,37,38]. Although there were opposite reports about mixed halide formation. It was shown that when other halide ions, i.e., X = Br/Cl, are used with a minimum concentration in MAPbI3 precursor, it does not affect MAPbI3 characteristics, and the final perovskite remains crystalline MAPbI3 phase-only instead of mixed halide perovskite. This is because a low concentration of halide ions can quickly evaporate as the films are annealed during the fabrication process. Some of the examples of such additives are amphiphilic hexadecyl trimethyl ammonium bromide(CTMAB), ethyl ammonium chloride (EACl), 1,3-diaminoguanidine monohydrochloride (DAGCl), benzamidine hydrochloride (BMCl), acetamidine hydrochloride (AcHc), methoxyammonium chloride (MeOCl), 2, 2, 2-trifluoroethylamine hydrochloride (TFEACl), benzenamine hydrochloride(BACl), 3-chloropropylamine Hydrochloride (3-CPACl), diethylamine hydrochloride (DEACl) [39,40,41,42,43,44,45].
Furthermore, nitrogen-containing heterocyclic amines play a great role in improving stability [46]. For instance, the introduction of 4,4′-bipyridine in MAPbI3 demonstrated complex formation with PbI2, thus improving intrinsic stability against illumination [47]. Additionally, it has been shown that the presence of N atoms prevents the loss of volatile species from the perovskite film and acts in a similar way as polymeric passivation coating. Further, such a passivation effect was visible, morphologically (in the SEM cross-section images and energy dispersive X-ray analysis), when PVC was added in MAPbI3 [48]. Later, scientists designed pyridine derivatives with units containing different multifunctional groups. One such additive is (C60-PyP), which contains C60 units that are hydrophobic in nature and pyridine units, which chelates Pb2+ by donating the lone electron pair on the N atom. It is known that uncoordinated Pb2+ ions are considered trap states at grain boundaries and can block charge extraction [49]. Thus, grain boundaries are passivated by assisting coordination interactions with the Pb2+ ion of MAPbI3 and PyP unit that further improved short circuit current density Jsc and eliminated ion migration. While the hydrophobic C60 unit does not let moisture directly affect the film [49]. Similar results were noticeable when pyridine-2-carboxylic lead salt (PbPyA2) and polyvinylpyrrolidone (PVP) additives were employed in MAPbI3 precursor [36,50]. Additionally, solvent additives with heterocyclic nitrogen-containing units also demonstrated similar results. In an investigation, Zhang and co-workers demonstrated that NMP as a solvent additive (when added in DMF) resulted in a high quality of perovskite film due to the Lewis acid–base reaction with Pb atom. The study revealed that intermediates obtained with different PbI2/NMP ratios are of the same kind, because of which, the solar cell performance and stability do not depend on the NMP ratio in the precursor [51]. Following the idea of the Lewis acid–base interaction, KIM and co-workers explored pyrrolidone-based solvent additives and compared with commonly used DMF-based precursor [52]. For pyrrolidone-based solvent additives, different N-substituents: N-methyl-2-pyrrolidone (NMP), N-ethyl-2-pyrrolidone (NEP), N-cyclohexyl-2-pyrrolidone (CHP), and N-octyl-2-pyrrolidone (NOP) was employed. During the in situ crystal growth, it was observed that solvent additive containing pyrrolidone structure with higher boiling point results in lower vapor pressure over pristine DMF-based precursor. Thus, it resulted in a further improvement in the morphology of perovskite film since the N substituent becomes bulkier with a higher boiling point (the boiling points of NMP, NEP, CHP, and NOP are 154 °C, 202 °C, 204 °C, 286 °C, 303 °C). Furthermore, it was discovered that the strength of coordination of solvent component influenced intermediate formation with PbI2, which, in turn, was influenced by the boiling point of the solvent. This was based on the evidence when the intermediate phase vanished with the CHP solvent additive, whereas pure DMF solvent without pyrrolidone-based solvent additives showed DMF: PbI2 solvate formation. Therefore, it was suggested that the bulkier additive solvent suppress DMF: PbI2 solvate formation, which shows lack of stability when used without any cosolvent [52]. Some other derivatives, such as DMI, 1-(4-ethenylbenzyl)-3-(3,3,4,4,5,5,6,6,7,7,8,8,8-tridecafluorooctylimidazolium iodide (ETI), 1-methyl-3- propylimidazolium bromide (MPIB), 1-butyl-3- tetrafluoroborate (BMIMBF4) are other examples that result in improved PCE and ambient stability due to hydrophobicity of N atoms with a ring-like structure [53,54,55]. Figure 1 below shows the chemical structures of additives based on amine derivatives.

2.1.2. Cyano/Nitrile Additives

Subsequently, it was observed that the cyano group-containing N atom could be equally effective to passivate grain boundaries [56]. Triazine-graphdiyne (Tra-GD) is a graphene-like material that consists of pyridine-like nitrogen in the highly conjugated framework. It was suggested that balanced charge distribution in the C=N bond in the triazine ring promotes stronger interaction with Pb2+, resulting in very tight contact between these two materials. This, in turn, did not allow the ion migration on the surface and successfully passivated grain boundaries [56]. Similarly, graphitic carbon nitride addition also resulted in good morphology due to highly compact bonding between Pb2+ and N of C=N that further endorsed to achieve power conversion efficiencies (PCEs) up to 21.1% (aperture 0.16 cm2) and 19.5% (aperture 1.0 cm2), featuring an open-circuit voltage of 1.16 V (corresponding to a slight voltage loss of 0.39 V), and improved operational device stability (~90% initial efficiency retained after constant one sun illumination for 500 h) [57,58]. Further, Zhou and co-workers demonstrated that, by employing 1,8-Diazabicyclo[5.4.0]undec-7-ene (DBU), the formation of I can also be suppressed due to adduct formation with PbI2 [59]. Moreover, another solvent additive, acetonitrile (ACN), also showed improvement in morphology due to adduct formation with PbI2 when added in a DMF-based precursor. This further proved to increase PCE from 15.04% to 19.7%, which remained stable without encapsulation for 160 h, with the PCE remaining ~60% of the initial value when exposed to the white light [60]. The chemical structures of the mentioned nitrile-based additives are illustrated in Figure 2.
Further, at a glance, the summary of additives that improve stability is presented in Figure 3. In Figure 3, “Extrinsic” represents results reported by conducting stability tests in air, moisture, or in ambient conditions, whereas “Intrinsic” represents results reported by conducting stability tests in inert, inside glovebox (nitrogen or argon filled), or vacuum. All subsequent figures depicting stability time versus additive will use these terms and represent the aforementioned. Here, the stability time refers to the time for which the active layer with additives was reported to be operational/functional in corresponding references. Additional details of additive influence on power conversion efficiency (PCE) and stability are described in Table 1, displaying the additives based on the N donor atom for the MAPbI3 active layer.

2.2. O Donor Atom-Based Additives

Similar to N donor, O donor and S donor can also act as Lewis bases due to lone pair of electrons available to co-ordinate with Pb defect site. Additives containing O atom as the Lewis base may consist of various functional groups, such as carbonyls, carboxyl, carboxylates, ester, ethers, and alcohols [28].

2.2.1. Carbonyl and Amide Additives

One such systematically investigated additive is urea. Urea is an eco-friendly compound consisting of a carbonyl group known to interact with PbI2 [65]. It has been illustrated that PbI2 and urea forms an intermediate PbI2.O=C(NH2)2 when added to the PbI2 precursor solution in the double-step spin coating process. The oxygen atoms act as a Lewis base. Whereas Pb2+ acts as Lewis acid, resulting in the formation of Lewis acid–base adduct. This adduct formation further results in large, flat grains [66,67]. Similar results were observed for the single-step spin coating process, resulting in grain boundary passivation, improved crystalline film, and suppressed non-radiative recombination losses, further improving the efficiency of solar cells [67,68,69]. Moreover, it was shown that incorporating amides in MAPbI3 reduces the Fermi level by interacting I defect vacancy, reducing the trap sites, and increasing the work function of MAPbI3 perovskite. At the same time, the carbonyl group transfers electrons to perovskite, reducing the Fermi level, which in turn helps to achieve better charge transport properties [70]. Recently, a variant of urea, biuret incorporation in MAPbI3 precursor attributed the intermediate formation with PbI2 to the electron delocalization in the N–C=O–N system in the presence of carbonyl group and explained this is why the peak shift for C=O vibration is noticed in FTIR spectra (from 1722 to 1713 cm−1 in case of biuret-modified MAPbI3 film) [71], resulting in higher PCE over urea incorporated devices [72]. Some other additives, such as benzoquinone (BQ), triazine perylene diimide (TPDI), 3,4-dihydroxybenzhydrazide (DOBD), containing the carbonyl group have shown similar properties [73,74,75,76]. Similarly, it was also shown that, if a hydrophobic ring is attached within the additive framework in combination with the carbonyl group (Isatin-Cl additive), it also acts as a shield against humidity in ambient conditions, helping to maintain the PCE nearly to 95% of the initial state [77]. Moreover, the usage of the carbonyl group can also be expanded to improve the flexibility of flexible, printable solar cells (FPSC). Polycaprolactone (PCL) additive in MAPbI3 has shown desirable improvement in PCE as well as mechanical strength. This improvement in mechanical stability was ascribed to a two-fold increase in grain size (from 200 to 400 nm) and homogenous grain distribution, leading to improved mechanical stability that maintained PCE more than 90% of the initial state after 300 bending cycles for the radius from 20 to 4 mm and stability under illumination for 350 h [78].

2.2.2. Sulfonyl Additives

Furthermore, carbonyl-based additives are often compared with the sulfonyl-based group, where the O atom is connected to different species, (i.e., S atom instead of C atom). Fang and co-workers have done one such study in which the strength of Pb coordination of the sulfonyl group containing additive was compared with the carbonyl-based additive. For this purpose, two novel fused ring non-fullerene acceptor materials IDIS-Th and IDIC-Th were introduced into MAPbI3. IDIC-Th had the Lewis base functional group carbonyl (C=O), and IDIS-Th had a sulfonyl group (O=S=O). Since the sulfonyl group (O=S=O) is a stronger electron-withdrawing group compared with carbonyl (C=O); the sulfonyl group was predicted to have a stronger interaction with Pb ions than carbonyl (C=O) because the two sulfur-oxygen double bonds can chelate Pb2+ with stronger interaction. Moreover, it was noticed that replacing carbonyl with the sulfonyl group could effectively downshift the lowest unoccupied molecular orbital (LUMO) level of IDIS-Th, which indicated the sulfonyl group is a stronger electron-withdrawing group. Thus, the interaction with Pb would reduce the defect traps density. Further, to confirm the defect passivation effect of IDIC-Th and IDIS-Th, the trap densities (ntrap) in bulk MAPbI3 films were calculated and found to be 8.85 × 1015 cm−3 and 4.20 × 1015 cm−3, respectively. Undoubtedly, the sulfonyl group reduced the defect states. This further reflected improvements in the contact angles of the film (the hydrophobicity of the films). The improvement in hydrophobicity indicated that the additive molecules at the GBs and surface could inhibit the raid of moisture to MAPbI3 films. However, both IDIC-Th and IDIS-Th are known for their hydrophobic nature. Nevertheless, these changes reflect improvement in hydrophobic nature. Moreover, the influence of the two molecules on photovoltaic performance was also examined. The PSC with IDIS-Th molecule showed significant improvement in PCE and reached up to 20.01%. Additionally, due to the strong passivation effect of IDIS-Th, PSC with IDIS-Th demonstrated superior stability over PSC with IDIC-Th in ambient conditions, under solar radiation and in the dark at high temperature [79]. Likewise, a comparison of solvent additives hexamethylphosphoric triamide (HMPA) and dimethyl sulfoxide (DMSO) was made, which was further compared to solvent additive NMethyl-2-pyrrolidone (NMP). Here, Lewis basicity, donor number (DN), and the boiling point of solvents were utilized as comparison parameters to correlate the role of solvents in regulating morphology. This investigation, conducted by Cao and Wei, and co-workers, highlighted that the larger the donor number, the stronger is the Lewis basicity of the solvent additive. The donor number of DMF, NMethyl-2-pyrrolidone (NMP), dimethyl sulfoxide (DMSO), and hexamethylphosphoric triamide (HMPA) are 26.6, 27.3, 29.8, and 38.8, respectively. Therefore, it was suggested that Pb-O bond strength would increase with an increase in the donor number of the additive [80,81]. Hence, NMP, DMSO, and HMPA were expected to form stronger intermediates over pure DMF. In addition, it is also observed that the strong Lewis base additives exhibit higher boiling points. (The boiling points of the DMF, NMP, DMSO, and HMPA are 152, 202, 189, 235 °C, respectively) [82,83]. This, in turn, reduces the evaporation rate of the solvent during the spin-coating process, resulting in a good film morphology. As a result, DMSO followed by NMP cosolvents displayed excellent PCE of 14.66% and 16.17% over pristine (DMF) (12.25%). No residue of PbI2 remained and complete conversion into perovskite resulted in less resistance [84]. Further, other solvent additives based on O donor atoms, such as tetramethylene sulfone (TMS) [85] and tetrahydrothiophene oxide (THTO), also utilized a similar mechanism of forming the Pb–O bond and improved performance and stability [86]. The chemical structures and influence of amides/carbonyl/sulfonyl group-based additives onto the stability of the MAPbI3 layer are presented in Figure 4 and Figure 5, respectively, with a glance summary shown in Table 2.

2.2.3. Carboxylic/Acid Additives

Acid additives have also been incorporated in the MAPbI3 precursor. The most commonly reported acid additive is 5-amino valeric acid (5-AVA). Dauskardt and co-workers showed that the introduction of 5-AVA into MAPbI3 precursor could be helpful to increase the mechanical robustness of the perovskite film. To test the mechanical reliability, the double cantilever beam method was used, and the cohesion energy (Gc) was calculated from the critical load at which the film created cracks. It was observed that the incorporation of 5% concentration of 5-AVA into MAPbI3 precursor increased cohesion energy (Gc) from 0.61 ± 0.27 J/m2 (pristine) to 6.04 ± 2.04 J/m2 (with AVA additive) almost 12-fold. This improvement in Gc with 5-AVA was attributed to enhanced interaction forces due to the longer alkyl chains present in the 5-AVA [87]. Additionally, a further increase in plasticity and crack deflection around the additive-containing perovskite grain boundaries was observed. Later, the elastic moduli of MAPbI3 perovskite films with 5-AVA showed a slight decrease in the elastic modulus from 20.5 GPa (pristine) to 18.9 GPa (5-AVA) (using the nanoindentation method). This decrease in stiffness of the planar perovskite structure made it more mechanically robust and less brittle. Moreover, it was confirmed that adding 5 AVA in MAPbI3 precursor does not change MAPbI3 lattice parameters [87]. Simultaneously, another study done by Durrant and the group reported enhanced photostability of MAPbI3 solar cells due to surface defect passivation in screen printed, HTM free, carbon electrode-based PSC consisting of 5 AVA in MAPbI3. They found a 40-fold increase in device lifetime measured under full sun illumination in ambient air (RH ± 15%). Further, it was proposed that AVA is located at grain boundaries and, thus, able to passivate surface defect sites, resulting in enhanced resistivity to oxygen-induced degradation [88]. The impressive improvement in the stability due to the AVA additive is attributed to the interaction of halides with –COOH and NH3+ group, through hydrogen bonding ((O–H----I) and (N–H----I)), and further verified in the large area cell with an area of 0.25 cm2, which demonstrated the efficiency of 6.6% with a decent shelf life of 75 days, maintaining more than 90% of initial efficiency [89]. Later, using polyacrylic acid (PAA), Cao and workers demonstrated that acid additives could be useful for large area (1 cm2) films using the doctor blade method. Using the carboxyl functional groups of PAA, iodide ion vacancies at the perovskite crystal surface could be cross-linked. Thus, the interaction of the PAA molecule with MAPbI3 passivates defects and improves PCE and stability [90]. In addition, it was demonstrated that bifacial passivation could also be obtained by the interaction of a functional group of –COOH, combined with –C, –S (of thioctic acid (TA)). Lewis acid–base reaction between under-coordinated Pb2+ and S atom in MAPbI3 can passivate one side, and at the same time, TiO2 can interact with COOH through hydrogen bonding, resulting in double-sided passivation [91]. Likewise, various studies reported that if multiple –COOH groups are attached with hydrophobic phenyl ring—it can chelate with Pb2+ atom and act as a shield against moisture and humidity under ambient conditions, maintaining the overall crystallinity of film. Terephthalic acid (TPA) and trimesic acid (TMA) are examples. The proposed mechanism of stability in such a case is given as follows: negatively charged groups (–COO) strongly attract Pb2+ ions in the precursor solution via electrostatic interactions. At the same time, the hydrogen atom from carboxylic acid can bond covalently with halide anion in solution due to the strong electronegativity of halide ions. Thus, it overall suppresses the loss of iodine molecules and, in turn, slows down the decomposition of MAPbI3 light absorbing layer. Simultaneously, benzene ring with rigidity and π–π bond effect does not allow interacting with water and protects against thermal stress and UV-illumination in ambient conditions [92,93]. Similar effects were observed when 3,3’,5,5’-azobenzene-tetracarboxylic acid (H4abtc) containing two benzene rings connected via azo bonding is introduced in PbI2 precursor in optimum amounts (2 mass %). The purpose of using the azo group was to reduce the stiffness of the perovskite film [94]. However, due to the hydrophobic nature of the attached cation in the acid-containing additive, most of them increase ambient stability (see Figure 6 for chemical structures of acid-containing additives and Figure 7 for associated stability). Further details of acid–based additives in MAPbI3 are provided in Table 3 below.

2.2.4. Carboxylate/Acetate Additives

In a similar fashion, the acetates in the form of additives also show improvement in crystallinity and grain size, passivating the grain boundaries. One of the examples is formamidine acetate salt (FAAc). It was reported that FAAc controlled the film morphology and improved fill-factor over 80%, which in turn improved PCE to 16.59%. The improved photovoltaic parameter was further associated with the incorporation of FAAc that eliminated the defect and trap density in the MAPbI3 film and, thus, improved the charge transport efficiency and reduced the hysteresis [98]. On the contrary, methyl ammonium acetate (MAAc) resulted in lower crystallinity and a smaller grain size ~200 nm [99]. However, it was suggested to use MAAC together with thiosemicarbazide (TSC) salt in the precursor to overcome lower crystallinity. This combination of additives (10–15% MAAc (molar ratio) and a tiny amount of TSC (3–5% molar ratio)) gave a certified PCE of 19.19% for an aperture area of 1.025 cm2, and the high-performing devices were able to sustain over 80% of the initial PCE after 500 h of thermal ageing at 85 °C [100]. Moreover, acetate additives were reported to be used as complementary to halogens. For instance, ammonium acetate (NH4Ac) was found to produce full film coverage with higher crystallinity. It was speculated that NH4Ac resulted in the formation of methylamine acetate salt by forming NH4PbI3 as an intermediate phase with a volatile by-product NH3 resulting in a higher nucleation density of grains and full coverage of the surface. This, in turn, resulted in a higher efficiency of 13.9% in carbon-based solar cells (FTO/TiO2/perovskite/Carbon), and 17.02% in the two-step spin coating in FTO/TiO2/CH3NH3PbI3/spiro-OMeTAD/Au configuration [101,102]. Further, the influence of NH4Ac addition was compared with NaAc and ZnAc2. Sadly, the incorporation of sodium acetate (NaAc) into MAPbI3 precursor reported no improvement in photovoltaic performance. However, the PCEs were comparable. While compared to Zinc-acetate (ZnAc2) additive, devices with ZnAc2 demonstrated improved fill factor and short circuit current density over reference cells, resulting in improved PCE from 11.1% to 12.30%. Furthermore, against the speculation that Zn2+ might partially replace Pb2+, due to the smaller ionic radius, the small quantity of ZnAc2 (MAI/PbI2/ZnAc2 (1:1−x:x, x = 7% as of molar ratio)) was reported to assimilate well within the MAPbI3 framework, and maintained ~89% of initial PCE after 1900 h when stored in Rh=40%, dark. However, NH4Ac incorporated devices exhibited ~96% PCE of the initial stage after 1900 h with the same conditions [101,102]. Later, lead acetate (PbAc2) was employed as an additive, demonstrating excellent PCE improvement from 17.25% to 19.07%. It was further demonstrated that incorporation of PbAc2 in precursor creates an intermediate phase by forming hydrogen bonding due to interaction with MA+ and O from acetate. Thus, this hydrogen bonding acts as a cross-linking agent for the intermediate phase, which later improves the intrinsic stability of devices that maintain almost 95% of the initial PCE for 20 days [103]. A similar mechanism was utilized in another study where the addition of barium acetate (BaAc2) was explored in MAPbI3 perovskite with a lower concentration (0–2 mg/mL), which suppressed ion migration and improved thermal stability. The devices could sustain ~90% PCE of the initial stage after exposure to thermal stress at 90 °C in the dark in inert conditions without any encapsulation up to 400 h [104]. Thus, by employing acetate-based additives (Figure 8), improved stability can be obtained both in ambient and inert conditions (Figure 9). Further, the acetate-based additives are summarized in Table 4.

2.2.5. Ester- and Ether-Based Additives

Interestingly, phenyl-C61-butyric acid methyl ester (PCBM), which is commonly employed as a charge transport layer in organic–inorganic hybrid solar cells, was also examined as an additive in the perovskite layer to achieve hole transport material (HTM) free fully printable MPSC (mesoporous perovskite solar cell). It was shown that, when the concentration of PCBM increased from 0 to 0.25 mg/mL, the PCE significantly improved from 8.58% to 12.36%, achieving a PCE enhancement of more than 44%. This enhancement of PCE was further ascribed to increased Jsc from 14.56 mA cm−2 to 20.26 mA cm−2 due to higher photogenerated charge separation, and suppressed the charge recombination process. Later, it was illustrated that the PCBM-perovskite intermediate phase was facilitated due to the formation of an intermediate adduct with PbI2 and carbonyl groups in PCBM [105]. Afterwards, Hu et al. compared PCBM incorporation with only the C60 unit and C60-taurine unit in a single-step spinning coating method in p-i-n configuration. It was reported that C60 (16.59%), PCBM (15.94%), C60-Ta (16.59%) all improved PCE over pristine (14.87%), which was consistent with the previously published report. Further, it was found that the addition of C60 and its derivatives decreased the trap densities and exhibited higher stability when exposed to ambient conditions RH = 25–50% without encapsulation [106].
Moreover, investigations have shown that ethers can help passivate grain boundaries on the perovskite surface. Poly (propylene glycol) bis (2-aminopropyl ether) (PEA) and Jeffamine are some of the examples that showed interaction with Pb ions due to the lone pair of the electron on the oxygen-containing ether part of the additives. Simultaneously, the MA+ cation from perovskite forms hydrogen bonding with the counter ion of the additive. Thus, the polymer works as a cross-linking agent that decreases trap densities and hysteresis, improving the device performance and stability. Moreover, the cross-linking properties can also improve the ductility of the perovskite film while stretching when fabricated on a flexible substrate (such as D2000, a Jeffamine variation) [107,108]. Furthermore, using the mechanism above, ethyl cellulose (EC), a low cost, environmentally friendly, thermally stable, and water-insoluble compound, also demonstrated chemical passivation of defect traps. Thus, because of the passivation effect of EC additive, the MAPbI3 crystal structure remained stable against moisture and air, maintaining 80% of PCE when exposed unencapsulated to ambient air at RH = 45% in the dark for 30 days [109]. Later on, tetra-orthosilicate (TEOS) additive in MAPbI3 was also reported to lower the cost of processing by making the fabrication process feasible in the air instead of the glove box, which was found valid for both single and double step spin-coating methods. This achievement is due to the fact that, when TEOS is introduced to MAPbI3 in air, SiO2 precipitates, acting as a passivation for the perovskite surface and shielding it from moisture and air. This improved the overall stability in ambient conditions [110,111,112]. Further, cyclic ether compound THF (tetrahydrofuran) also reported improving stability in ambient air, which the pioneer of the PSCs, the Miyasaka group, investigated. The study demonstrated that THF and PbI2 in DMF interact and form a complex, resulting in dense, homogenous, and pinhole-free film due to complex formation between the Pb and O donor from THF, because of the Lewis acid nature of Pb and Lewis basicity of THF. This, in turn, avoids a direct interaction of PbI2 with water, improving T80 (T80 is the time at which PCE is degraded to 80% of initial PCE) for 20 days in the ambient environment, with 50% RH, when devices were exposed unencapsulated [113]. Similarly, the liquid additive diethyl ether was also found to influence the crystallization process, resulting in large grains, leading to improved photocurrent properties and performance [114]. The chemical structures (Figure 10) and an overview illustrating the influence of ester and ether-based additives on the stability of MAPbI3-based solar cells is shown in Figure 11, respectively. Further, a table describing ester and ether containing additives and their role in the MAPbI3 layer is presented in Table 5.

2.2.6. Hydroxyl/Alcohol

The Hydroxyl group can also play a vital part in interacting with perovskite through hydrogen bonding, further resulting in improved crystallinity and performance. For example, enhanced PCE and stability was observed when dibutyl hydroxyl toluene (BHT) additive containing a phenol group was mixed in MAPbI3 precursor [115]. Furthermore, methanol and isopropanol showed larger grain size, reduced grain boundaries, lessened defect density, and demonstrated efficient charge carrier extraction at the interfaces, leading to improved PCE stability in ambient conditions without encapsulation. Similarly, improvement in PCE was also noticed due to other hydroxyl-based additives, such as 2-methoxy ethanol (ME), 2-ethoxyethanol (EE), and 2-propoxy ethanol (PE) and n butanol [116,117,118]. The chemical structures of the discussed additives are shown in Figure 12. The influence of some of the alcohol additives on the stability of MAPbI3 perovskite is displayed in Figure 13. Further, a summary of alcohol-based additives, their influence on the PCE, and their role with the perovskite framework is described in Table 6.

2.2.7. Oxygen-Based Multifunctional Group Containing Additives

Other than these additives, other oxygen atom-based additives with multiple functional groups, consisting of C=O, COOH, and OH, were also employed in the MAPbI3 precursor. One of the examples is reduced graphene oxide (rGO), which was used to achieve fast electron transport rates toward anode, including the growth of large, uniform, smooth, and crystalline perovskite film, resulting in a massive improvement in device performance from 13.6% to 20% [120]. Another example includes poly (amic acid) (PAA) and polyimide (PI) additive. However, the PAA and PI additives also consisted of the N atom, other than C=O, COOH and the OH functional group. The O atoms in PAA and PI form hydrogen bonds with H atoms in CH3NH3+. The lone pair of electrons of N atoms interact with Pb ions, which stabilized the PVSK framework. Furthermore, these interactions improved the optoelectronic properties of the perovskite layer. As a result, an increment in the grain size was observed when 0.0497 mg/mL PAA-derived perovskite was employed, illustrating the Lewis acid−base interaction between C=O and Pb2+ that controls the crystallization process and defect passivation of PVSK. Moreover, both PAA and PI are hydrophobic and highly heat-resistant polymers and further contribute to the stability of PSC when operated in a humid and high-temperature environment [121]. The chemical structure of these additives and their influence on stability are shown in Figure 14 and Figure 15, respectively. Table 7 presents the detailed description of these additives.

2.3. S Donor Atom-Based Additives

2.3.1. Sulfide and Organosulfur Additives

Further, sulfur-based donor atoms have also improved crystallinity and device performance [28]. For instance, Yang and co-workers used dimethyl sulfide (DS) additive in MAPbI3 precursor solution (based on DMSO precursor) and demonstrated a record high PCE of ~18.4% for a flexible perovskite solar cell. In addition, the formation of a stable intermediate complex between PbI2 and DS molecule was observed, which further enhanced stability. The enhancement of stability due to DS solvent additive was ascribed to the smaller electronegativity of the sulfur atom (2.5) in the DS molecule compared to an oxygen atom (3.5) in the DMSO molecule, which allowed sulfur molecules to provide electrons to empty 5d Pb orbitals, ultimately resulting in a hexa-coordinated complex formation. Therefore, with better crystallization and stable intermediate formation, 86% of initial PCE remained for 60 days when unencapsulated devices were exposed to moisture RH~35% in the dark at room temperature [122]. Thiourea is another example of such an organosulfur compound, the addition of which, into the photoactive layer, results in improved performance and stability due to intermediate formation because of the Lewis acid–base reaction. However, the addition of thiourea in the active layer is often compared to the similar compound urea, where the O atom from the carbonyl group (in urea) is replaced with the S atom for thiourea. (i.e., the carbon atom is connected to different species—C=O for urea and C=S for thiourea). Both thiourea and urea were found to show similar effects of increasing grain size, crystallinity, grain boundary passivation, and coordination with PbI2 in the MAPbI3 framework [123]. Moreover, many other organosulfur derivatives were reported to form an adduct with PbI2 as the Pb atom acts as the Lewis acid. One such example is a volatile Lewis base, thioacetamide (CH3CSNH2), abbreviated as TAA [124]. Another example is 2-pyridylthiourea, which has an S donor from thiourea, with a better ability to coordinate with PbI2, and an N donor from pyridine, which can co-ordinate with PbI2 as a Lewis acid base and regulate the co-ordination strength with PbI2. This further resulted in an improvement in crystallinity, which in turn resulted in better light-harvesting ability, charge transport, the reduction of defect and recombination loss, leading to PCE of 18.2%. In addition, this allowed retaining the stability of PSC more than 90% when exposed to ambient conditions, at room temperature or at higher temperature in the dark [125]. Recently, Zhang and his co-workers combined thiourea with water and iodine (ITU) ions to reduce iodine defect vacancy produced during the film fabrication process. Afterwards, the grain size was found increased with ITU (3.4 um) compared to when thiourea was added (800–2000 nm), reducing the defects at grain boundaries. Moreover, PSCs with ITU additives displayed a higher Voc, longer decay time, and longer carrier lifetime, with a lower charge recombination rate. Thus, overall, the addition of ITU resulted in improvement in PCE from 17.75% to 20.3%. In addition, the trap densities of the perovskite film significantly reduced as the ITU additive was incorporated into PSC, allowing PSC to stabilize and maintain ~80 of initial PCE after 100 h of exposure to 1.5 AMG solar irradiation and 30 days in the ambient atmosphere [126].

2.3.2. Thiocyanates

Thiocyanates (SCN containing anions) additives are another subcategory for S atom-based additives. It should be noted that thiocyanates are also considered pseudo halides. However, the role of thiocyanates in the perovskite film is very controversial. For example, Yanfa Yan et al. demonstrated that a small amount (5%) of Pb(SCN)2 addition in the precursor can significantly increase the grain size and crystalline quality of perovskite thin films using a one-step solution process method. This increased the average PCE from 15.57 to 17.80%. It was proposed that SCN anions do not incorporate in the perovskite film. SCN react with CH3NH3+ and form volatile products HSCN and CH3NH2. This allows excess Pb2+ and I to form PbI2 at grain boundaries that sources passivation effect [127]. On the contrary, Kim et al. demonstrated that Pb(SCN)2 in CH3NH3PbI3 partially substitutes I anions and forms CH3NH3Pb(SCN)xI3−x at lower temperatures. This phenomena was observed as not dependent on the concentration of Pb(SCN)2. However, similar reports of enhancement of crystal size were found. Thus, it was concluded that the influence of processing temperature could affect the purity of the final perovskite film when Pb(SCN)2 additive is used with MAPbI3 [128]. Whereas, the addition of potassium thiocyanate (KSCN) also reported to form volatile products HSCN and CH3NH2, consistent with earlier mentioned results, with no traces of SCN found in the final perovskite film [127]. In the case of the KSCN additive, however, the residue was KI instead of PbI2. Nevertheless, the authors illustrated that the final product upon KSCN incorporation was MAPbI3, not KPbI3, due to the smaller radius of the K cation. In addition, KI residue improves the crystal quality of perovskite film, resulting in enhancement of charge transport, reduced carrier recombination, and eliminating hysteresis [129].
Conversely, when NH4SCN was incorporated in MAPbI3, it was reported to form an unstable intermediate phase NH4PbI3−xSCNx due to lower formation enthalpy (△Hf) of the NH4PbI3 over MAPbI3, which later resulted in CH3NH3PbI3−xSCNx a mixed halide perovskite after annealing of the films [130]. Therefore, when methylammonium thiocyanate (MASCN) was introduced as an additive to the precursor, a rapid vacuum-based drying approach was used to extract volatile intermediate product, so that MAPbI3 was formed as the final product with no traces of SCN. The film fabricated, using this fabrication technique, was reported to gain grain size of more than one micrometer with uniform surface morphology, resulting in high crystallinity and significantly large carrier lifetimes (τ1 = 931.94 ± 89.43 ns; τ2 = 320.41 ± 43.69 ns), improving the ambient stability up to 1000 h [131]. However, when a larger cation guanidinium (Gu) was directly added in MAPbI3 as GuSCN, GuPbI3 was formed as an intermediate phase present with the MAPbI3 phase. It was contradictory to the earlier belief and reports, which showed that Gu single-handedly could not form three-dimensional perovskite materials due to a larger ionic size (278 pm) over MA cation (217 pm). As a result, improved crystallinity, grain size, and reduced trap density were noticed further, resulting in improved PCE (from 15.57% to 16.70%) that maintained stability ~90% of the initial value after being stored for 15 days without encapsulation [132]. On the other hand, when guanidinium isothiocyanate was added in the PbI2 precursor, with the dual step fabrication method; mixed cation perovskite was formed (GU)x(MA)(1−x)PbI3 [133]. Nevertheless, most S donor atom-based additives (Figure 16) were found to influence ambient stability, as displayed in Figure 17. At a glance, a summary of additives containing S atom and their role in the MAPbI3 framework are presented in Table 8.

2.4. Alkane Additives

Alkane-based additives are another class of additives that was briefly investigated within the MAPbI3 framework. Additives such as poly(vinylidene fluoride-co-hexafluoropropylene) (PVDF-HFP) have shown interaction with MA cation via hydrogen bonding with fluorine atoms in the additive. Such an additive is quite useful for indoor photovoltaic applications, as the additive is reported to control the nucleation and growth rate on a very thin (~150 nm) active layer of MAPbI3 through a one-step solution processing method. Thinner light absorber perovskite is usually full of voids, results in a low fill factor, and lower internal quantum efficiency due to poor perovskite film quality. However, increasing the thickness is not suitable to obtain semi-transparent films [134]. As a solution to this problem, a poly(vinylidene fluoride-co-hexafluoropropylene) (PVDF-HFP) additive illustrated less rough morphology when a ~150 nm thick perovskite layer was obtained. In addition, it was found that perovskite solar cells with PVDF-HFP in the active layer have a higher charge transfer rate and lower carrier recombination rate [135]. Similar observations were seen when another derivative of PVDF polyvinylidene fluoride-trifluoroethylene polymer P(VDF-TrFE) was employed in a two-step fabrication process instead [136]. Other than these additives, some other haloalkane additives, such as diiodomethane (CH2I2) and diiodooctane (C8H16I2), were also reported to reduce the trap states, resulting in improved morphology and enhanced PCE [137,138]. The chemical structures of alkane additives are shown in Figure 18. Further, Table 9 summarizes their role in the MAPbI3 framework.

2.5. Quantum Dot-Based Additives

Recently, a new area of additive engineering has emerged, based on quantum dots containing different functional groups introduced in the precursor. This allows the functional group to chelate under coordinated Pb vacancy or iodine defect sites at the surface of the perovskite layer (Figure 19). At the same time, a quantum dot improves the crystallinity of film due to the quantum confinement effect and reduces non-radiative recombination centers; thus, passivating defects at grain boundaries [139,140,141,142,143,144]. In some cases, quantum dots can even promote charge transfer from the oxide surface to the perovskite layer, further reducing the charge trap density, boosting fill factor as high as 84% and efficiency as 21.04% [143,145]. The influence of incorporation of QDs onto the stability is shown in Figure 20. A summary of such additives based on quantum dots (QDs) is presented in Table 10.

3. Inorganic Additives

Inorganic salts can be further explained as per the category in the periodic table, i.e., based on alkali metals, alkaline earth metals, transition metals, other metals, and other non-metals. These additives may exist in the form of inorganic halide salts or inorganic acids [146].

3.1. Alkali Metals Additives

Hydro halides are at the top, as hydrogen is the first element in group one of the periodic table. Soe et al. studied the incorporation of hydrohalic acids (HX, X=I, Br, Cl) in MAPbI3 precursor prepared using DMF. It was found that HX incorporation in the perovskite layer alters the bandgap and unit cell parameters. The HI addition compressed bandgap, whereas HBr widens the bandgap at high concentrations (20–25 vol%). Further, these changes were found correlated with the types of defects present in polycrystalline perovskite thin films combined with the structural strain induced in very small crystallites. Thus, it was concluded that these acids could influence crystallization rate, surface coverage, and improve morphology. However, HCl incorporation showed no influence on bandgap [147,148]. The same was verified by Yan and co-workers who investigated HI (hydroiodic acid) as an iodine quencher in MAPbI3 precursor solution (prepared with isopropyl alcohol (IPA)), that further lead to chemically passivate grain boundaries and increase in PCE from 16.13% (pristine) to 18.21% (with HI additive) [149]. Likewise, alkali metals, such as Li, Na, and K, combined with halogen counterparts, have been inspected as additives for the MAPbI3 light absorber layer. Grätzel and Friend co-workers demonstrated the influence of NaI (sodium iodide) on the optical, excitonic, and electrical properties of CH3NH3PbI3 perovskite prepared by a two-step sequential deposition technique. It was revealed that the NaI additive helps in the complete conversion of PbI2 into CH3NH3PbI3, leaving no unreacted PbI2 and enhancing the crystallinity. Furthermore, KPFM measurements showed a reduction in work function (towards Au metal) for contact-potential difference (CPD) and proved the chemical passivation of perovskite surface due to additive incorporation in the precursor. Additionally, the hypothesis of partial substitution of Pb2+ with monovalent cations Na was discarded, because replacing Pb2+ with the Na cations required high energy, which could not be obtained at room temperature, since additives and films were processed at room temperature [150]. Similar reports were published from Chu and co-workers who investigated alkali metal chloride additives (LiCl, NaCl, and KCl) in MAPbI3 and suggested that mixed halide formation does not occur as Cl atoms easily evaporate due to the annealing process, improving the crystallinity and carrier charge transport [151]. However, as the size of the nuclei increases, the perovskite shows partial replacement and forms a perovskite with a double cation (i.e., in the case of RbX and CsX, where X represents halide) [152]. The partial substitution of cation was also observed when alkaline earth metals were employed as additives in the precursor. Hence, they are not the center of attraction here and it is suggested that alkaline earth metals are best suited to replace Pb due to their divalent nature [153,154]. The chemical structures of alkali metal additives are presented in Figure 21. The example of the influence of alkali metal additives on the stability of MAPbI3 PSC is shown in Figure 22. Further, the role of alkali metal additives in the MAPbI3 framework is summarized in Table 11.

3.2. Transition Metals Additives

Additionally, some transition metals were also reported as additives in MAPbI3 precursor. One such example is n-type goethite (FeOOH) quantum dots that act as multifunctional additives. The addition of FeOOH QDs into precursor solution improves not only the performance, but also stability. The investigation by Wang and et al. showed that FeOOH QDs could produce heterogeneous nucleation, passivate the trap states, and obstruct the ion migration. The investigation demonstrated that the oxygen (in FeOOH) as a Lewis base could coordinate with Lewis acid Pb2+. The –OH group (in FeOOH) can coordinate with hydrogen atoms of MA+. This interaction delays crystal growth kinetics, resulting in good quality of perovskite films that result in high PCE. Further, the iron in FeOOH QDs as a Lewis acid interact with Lewis base I. Thus, ion migration of I and MA+ is controlled, resulting in enhanced intrinsic stability of solar cells [157]. Further, the transitional metal halides, either with monovalent cations (i.e., AgI, CuI, CuBr), divalent cation (NiCl2, CuCl2, ZnCl2, CdCl2) [158,159,160,161], or trivalent cation (RhI3) [162] frequently pointed out the improvement in crystallinity and grain boundary passivation.
Interestingly, in all these reports, chlorine-based additive salts were reported to form the MAPbI3 phase without mixed iodide and chloride as it was proposed that Cl atoms evaporate while annealing the films. Similarly, trivalent cation, such as Rh3+, did not replace lead atoms when used in small quantities (e.g., 1 mol%). However, the addition of 5% could result in partial replacement of the Pb atom. Figure 23 and Figure 24 demonstrate the chemical structures and the effect of transition metal additives on the stability of MAPbI3 PSC, respectively. Further, Table 12 summarizes the details associated with transitional metal additives.

3.3. Other Metals Additives

However, the most widely reported metal additive is PbI2. Several groups have suggested using excess PbI2 as an additive and studied the effect of the stoichiometric and non-stoichiometric composition of PbI2 in MAPbI3 perovskite [163,164]. Reports suggest that the addition of excess PbI2 in perovskite increases power conversion efficiency [163,165]. However, the influence of excess PbI2 on photochemical stability is debatable [166]. Some research groups claim that excess PbI2 passivates grain boundary [167]; thus, creates extra passivation [168]; enhancing performance and stability overall [164,169]. On the contrary, some reports showed that excess PbI2 could increase PCE, but decreased PSC stability [170,171]. To resolve the issue, Stevenson et al. demonstrated that the stability associated with excess PbI2 is affected by solvent coordination capacity with Pb atom, and suggested that excess PbI2 can work as a stabilizer to improve intrinsic stability if the correct solvent (NMP) is chosen over commonly-used solvent DMF [172]. Other than PbI2, adding PbCl2 also resulted in an equally good perovskite film quality with improved performance, but the outcome was reported as a mixed halide [173,174]. Hence, it is not focused on here. Interestingly, Ngo and the group demonstrated a different approach by using PbS quantum dots as capping ligands in the precursor solution to control nucleation and morphology. This approach allowed improving crystallinity and enhanced grain size, resulting in improved solar cell performance [175]. See Figure 25 for the chemical structures of PbI2 and PbS. The influence of excess PbI2 additive on the stability of MAPbI3 is presented in Figure 26. Further, a relevant summary is presented in Table 13.

3.4. Non-Metal Inorganic Salts

Furthermore, the controlled nucleation, improved morphology, enhanced crystalline size, and grain size with lower trap density were also noticed with additives based on other non-metals salts. During the two-step method, which is one of the commonly used fabrication approaches to fabricate planar heterojunction complex lead halide perovskite solar cells, the complete conversion of PbI2 into CH3NH3PbI3 is often observed restricted by the inadequate diffusion of CH3NH3I into PbI2 film, which affects the short circuit current density. To overcome this problem, Pathipati and co-workers used ammonium iodide (NH4I) in a small quantity (5 wt%) in the PbI2 solution and found this approach increases the porosity of the PbI2 film, allowing to improve film morphology and increase grain size up to 500 nm. This, in turn, improves the fill factor and short circuit current density, improving the PCE compared to the pristine solar cell [177]. Similarly, a drastic improvement in fill factor (from 27.53 without additive to 80.11 with NH4Cl additive) was noticed when NH4Cl was added in the MAPbI3 precursor. Moreover, as reported earlier, due to the quick evaporation of chlorine atoms, mixed halide formation did not take place [178]. Furthermore, improved performance and higher stability were observed with different ammonium halide salts (NH4X, where X can be I, Cl, Br) even when using the single-step solution process method. Although, NH4I showed the highest performance amongst all [177,178,179]. Another ammonium-based additive, NH4H2PO2 (ammonium hypophosphite-AHP), was reported to show similar behavior. Surprisingly, one of the by-products (H3PO2) from CH3NH3I synthesis was found suppressing the formation of molecular iodine (i.e., the oxidation of I to I2) [180,181]. This led to the idea of using NH4H2PO2 (ammonium hypophosphite-AHP) as an additive in the PbI2/CH3NH3I precursor. Huang and co-workers found that AHP forms a complex with PbI2, which leads to improvement in crystallinity and grain size. In turn, this improves PV performance due to crystalline grain boundaries [182]. Moreover, in a recent report, it was illustrated that hydrazinium iodide (N2H5I), an iodine quencher in the MAPbI3 active layer, can form an intermediate complex with PbI2, and enhance crystallinity and grain size ~1100 nm, causing grain boundary passivation. This, in turn, reduces hysteresis in solar cells, improves efficiency, and intrinsic stability of solar cells up to 4400 h [183]. There were similar reports by Huang et al., who reported hydrazinium chloride (N2H5Cl) incorporation in MAPbI3 [184]. Whereas hypophosphorous acid (HPA) was used to assist crystallization dynamics for large area substrates [185]. The chemical structures and influence of non-metal inorganic additives on the stability of MAPbI3 are presented in Figure 27 and Figure 28, respectively. A summary of these additives is presented in Table 14.

4. Conclusions and Outlook

To conclude, we covered the progress and updates in the area of additive engineering for the MAPbI3 photoactive layer in perovskite solar cells. As known, the poor intrinsic and extrinsic stability causes a restriction in the substantial commercialization of the PSC. Intrinsic instability is considered due to the presence of under-coordinated Pb sites or escape of a volatile product or formation of molecular iodine; either of these creates a defect in the perovskite surface. Therefore, in this review, we summarized various additives that can coordinate with the mentioned defect sites and improve stability. Higher efficiency (more than 25%) was achieved, but achieving higher stability is still far away. Further, the role of additives into MAPbI3 precursor was presented, which influences nucleation growth, film morphology, crystallinity, grain size, and grain boundary passivation via complex/intermediate formation were conferred. To understand the aforementioned, we classified additives based on organic and inorganic compounds. Furthermore, organic additives were scrutinized as per their functional group, and a relevant coordination mechanism that impacted stability was described. Similarly, inorganic additives were classified based on metal and non-metal halides. Moreover, we focused on additive influence on stability concerning different conditions (intrinsic/ambient/mechanical robustness). Our analysis suggests that amines, heterocyclic amines, and their derivatives, followed by carbonyl and acetate functional groups, can improve intrinsic stability. On the other hand, acid, esters, ether, and alcohol, and S atom-based additives can be mainly valuable to boost ambient stability. This summary of additives further aims to boost designing new additives that can finally help achieve stability challenges.
Yet, further developments are required while adopting the additive approach in the fabrication of MAPbI3-based PSCs. So far, additives employed in the MAPbI3 light-absorbing layer do not have any rule of selection. The majority of additives have opted from existing literature available for DSSCs or organic solar cells, utilizing the N/O/S donor atoms to combine with under-coordinated Pb2+ sites to improve stability because of the Lewis acid–base reaction. Hence, the first and the topmost prospect is to design an additive that can help sustain MAPbI3 under operational conditions, such as illumination, heat, moisture, oxygen, water. To do so, applying machine learning is highly recommended. While additives are primarily being experimented on (on a trial and error basis), the number of published reports are growing exponentially, with time, without any outcome to overcome stability issues. Therefore, a systematic empirical investigation should be conducted using analytical, statistical tools to understand the correlation of the additive, its properties, and its influence on morphology, optoelectronic properties, and stability with various conditions concerning numerous characterization techniques. This might help in the suitable material selections to further experiment by establishing the relationship between the chemical structure of additives and photovoltaic characteristics and lifetime. Thus, it would reduce the time and cost required for the experiment to improve stability, eventually driving industrial interest in product development.
The second prospect includes improved operational stability, which is reported to gain only a few thousand hours. While at least 10,000 h of operational stability is required to commercialize as the indoor application, with respect to thin-film photovoltaics. On the other hand, traditional photovoltaic has a lifetime of ~25 years. Thus, the second vision is to overcome short-term stability issues to commercialize the MAPbI3-based PSC. Moreover, the duration of operational stability should also be reproducible. Henceforth, fabricating reproducible stability is the third prospect to work on, as testing short- or long-term reproducibility is a time-consuming process. The next challenge is upscaling. Currently, the majority of research reports are based on small area devices. Nevertheless, large-area devices will be required for commercial purpose. However, morphology, performance, distribution of surface defects, grain size, and trap states will be changed when the large area will be used. Therefore, the fabrication of large-area devices is another prospect that requires high PCE and stability. Another future scope of development is in the area of toxicity of PSC. Since some precursors used to prepare MAPbI3 ink are not environment friendly. Hence, alternative, environment-friendly green solvent/precursors should be further explored to give equally good efficiency and stability.
In the end, we hope this review will help scientists and researchers to design and synthesize new additives that can provide the solution for the stability challenge of MAPbI3. Thus, ultimately achieving stability by applying additive engineering, commercializing MAPbI3-based PSC can be achieved.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interests.

References

  1. Wilson, G.M.; Al-Jassim, M.; Metzger, W.K.; Glunz, S.W.; Verlinden, P.; Xiong, G.; Mansfield, L.M.; Stanbery, B.J.; Zhu, K.; Yan, Y.; et al. The 2020 photovoltaic technologies roadmap. J. Phys. D Appl. Phys. 2020, 53, 493001. [Google Scholar] [CrossRef]
  2. Pham, H.D.; Yang, T.C.; Jain, S.M.; Wilson, G.J.; Sonar, P. Development of Dopant-Free Organic Hole Transporting Materials for Perovskite Solar Cells. Adv. Energy Mater. 2020, 10. [Google Scholar] [CrossRef]
  3. Fan, Y.; Meng, H.; Wang, L.; Pang, S. Review of Stability Enhancement for Formamidinium-Based Perovskites. Sol. RRL 2019, 3, 1900215. [Google Scholar] [CrossRef]
  4. Roy, P.; Sinha, N.K.; Tiwari, S.; Khare, A. A review on perovskite solar cells: Evolution of architecture, fabrication techniques, commercialization issues and status. Sol. Energy 2020, 198, 665–688. [Google Scholar] [CrossRef]
  5. Green, M.A.; Ho-Baillie, A.; Snaith, H.J. The emergence of perovskite solar cells. Nat. Photonics 2014, 8, 506–514. [Google Scholar] [CrossRef]
  6. Huang, J.; Yuan, Y.; Shao, Y.; Yan, Y. Understanding the physical properties of hybrid perovskites for photovoltaic applications. Nat. Rev. Mater. 2017, 2, 17042. [Google Scholar] [CrossRef]
  7. Leguy, A.M.A.; Hu, Y.; Campoy-Quiles, M.; Alonso, M.I.; Weber, O.; Azarhoosh, P.; van Schilfgaarde, M.; Weller, M.T.; Bein, T.; Nelson, J.; et al. Reversible Hydration of CH3NH3PbI3in Films, Single Crystals, and Solar Cells. Chem. Mater. 2015, 27, 3397–3407. [Google Scholar] [CrossRef]
  8. Ke, J.C.-R.; Walton, A.S.; Lewis, D.J.; Tedstone, A.; O’Brien, P.; Thomas, A.G.; Flavell, W.R. In situ investigation of degradation at organometal halide perovskite surfaces by X-ray photoelectron spectroscopy at realistic water vapour pressure. Chem. Commun. 2017, 53, 5231–5234. [Google Scholar] [CrossRef] [Green Version]
  9. Abdelmageed, G.; Jewell, L.; Hellier, K.; Seymour, L.; Luo, B.; Bridges, F.; Zhang, J.Z.; Carter, S. Mechanisms for light induced degradation in MAPbI3 perovskite thin films and solar cells. Appl. Phys. Lett. 2016, 109, 233905. [Google Scholar] [CrossRef]
  10. Bryant, D.; Aristidou, N.; Pont, S.; Sanchez-Molina, I.; Chotchunangatchaval, T.; Wheeler, S.; Durrant, J.R.; Haque, S.A. Light and oxygen induced degradation limits the operational stability of methylammonium lead triiodide perovskite solar cells. Energy Environ. Sci. 2016, 9, 1655–1660. [Google Scholar] [CrossRef] [Green Version]
  11. Krishnan, U.; Kaur, M.; Kumar, M.; Kumar, A. Factors affecting the stability of perovskite solar cells: A comprehensive review. J. Photonics Energy 2019, 9, 021001. [Google Scholar] [CrossRef]
  12. Manser, J.S.; Saidaminov, M.; Christians, J.; Bakr, O.M.; Kamat, P.V. Making and Breaking of Lead Halide Perovskites. Acc. Chem. Res. 2016, 49, 330–338. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Aristidou, N.; Eames, C.; Sanchez-Molina, I.; Bu, X.; Kosco, J.; Islam, M.S.; Haque, S.A. Fast oxygen diffusion and iodide defects mediate oxygen-induced degradation of perovskite solar cells. Nat. Commun. 2017, 8, 15218. [Google Scholar] [CrossRef] [PubMed]
  14. Li, B.; Li, Y.; Zheng, C.; Gao, D.; Huang, W. Advancements in the stability of perovskite solar cells: Degradation mechanisms and improvement approaches. RSC Adv. 2016, 6, 38079–38091. [Google Scholar] [CrossRef]
  15. Aristidou, N.; Sanchez-Molina, I.; Chotchuangchutchaval, T.; Brown, M.; Martinez, L.; Rath, T.; Haque, S.A. The Role of Oxygen in the Degradation of Methylammonium Lead Trihalide Perovskite Photoactive Layers. Angew. Chem. Int. Ed. 2015, 54, 8208–8212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Kim, N.-K.; Min, Y.H.; Noh, S.; Cho, E.; Jeong, G.; Joo, M.; Ahn, S.-W.; Lee, J.S.; Kim, S.; Ihm, K.; et al. Investigation of Thermally Induced Degradation in CH3NH3PbI3 Perovskite Solar Cells using In-situ Synchrotron Radiation Analysis. Sci. Rep. 2017, 7, 1–9. [Google Scholar] [CrossRef] [PubMed]
  17. Akbulatov, A.F.; Frolova, L.A.; Dremova, N.N.; Zhidkov, I.S.; Martynenko, V.M.; Tsarev, S.A.; Luchkin, S.; Kurmaev, E.Z.; Aldoshin, S.M.; Stevenson, K.J.; et al. Light or Heat: What Is Killing Lead Halide Perovskites under Solar Cell Operation Conditions? J. Phys. Chem. Lett. 2020, 11, 333–339. [Google Scholar] [CrossRef]
  18. Juarez-Perez, E.J.; Ono, L.K.; Maeda, M.; Jiang, Y.; Hawash, Z.; Qi, Y. Photodecomposition and thermal decomposition in methylammonium halide lead perovskites and inferred design principles to increase photovoltaic device stability. J. Mater. Chem. A 2018, 6, 9604–9612. [Google Scholar] [CrossRef] [Green Version]
  19. Kim, M.; Ham, S.; Cheng, D.; Wynn, T.A.; Jung, H.S.; Meng, Y.S. Advanced Characterization Techniques for Overcoming Challenges of Perovskite Solar Cell Materials. Adv. Energy Mater. 2020, 11, 1–26. [Google Scholar] [CrossRef]
  20. Latini, A.; Gigli, G.; Ciccioli, A. A study on the nature of the thermal decomposition of methylammonium lead iodide perovskite, CH3NH3PbI3: An attempt to rationalise contradictory experimental results. Sustain. Energy Fuels 2017, 1, 1351–1357. [Google Scholar] [CrossRef]
  21. Brunetti, B.; Cavallo, C.; Ciccioli, A.; Gigli, G.; Latini, A. Erratum: Corrigendum: On the Thermal and Thermodynamic (In)Stability of Methylammonium Lead Halide Perovskites. Sci. Rep. 2017, 7, 46867. [Google Scholar] [CrossRef] [PubMed]
  22. Akbulatov, A.; Luchkin, S.Y.; Frolova, L.A.; Dremova, N.N.; Gerasimov, K.L.; Zhidkov, I.S.; Anokhin, D.V.; Kurmaev, E.Z.; Stevenson, K.J.; Troshin, P.A. Probing the Intrinsic Thermal and Photochemical Stability of Hybrid and Inorganic Lead Halide Perovskites. J. Phys. Chem. Lett. 2017, 8, 1211–1218. [Google Scholar] [CrossRef] [PubMed]
  23. Juarez-Perez, E.J.; Hawash, Z.; Raga, S.R.; Ono, L.K.; Qi, Y. Thermal degradation of CH3NH3PbI3 perovskite into NH3 and CH3I gases observed by coupled thermogravimetry–mass spectrometry analysis. Energy Environ. Sci. 2016, 9, 3406–3410. [Google Scholar] [CrossRef] [Green Version]
  24. Deng, X.; Cao, Z.; Yuan, Y.; Chee, M.O.L.; Xie, L.; Wang, A.; Xiang, Y.; Li, T.; Dong, P.; Ding, L.; et al. Coordination modulated crystallization and defect passivation in high quality perovskite film for efficient solar cells. Coord. Chem. Rev. 2020, 420, 213408. [Google Scholar] [CrossRef]
  25. Zhang, F.; Zhu, K. Additive Engineering for Efficient and Stable Perovskite Solar Cells. Adv. Energy Mater. 2020, 10. [Google Scholar] [CrossRef]
  26. Azam, M.; Liu, K.; Sun, Y.; Wang, Z.; Liang, G.; Qu, S.; Fan, P.; Wang, Z. Recent advances in defect passivation of perovskite active layer via additive engineering: A review. J. Phys. D Appl. Phys. 2020, 53, 183002. [Google Scholar] [CrossRef]
  27. Mahapatra, A.; Prochowicz, D.; Tavakoli, M.M.; Trivedi, S.; Kumar, P.; Yadav, P.K. A review of aspects of additive engineering in perovskite solar cells. J. Mater. Chem. A 2020, 8, 27–54. [Google Scholar] [CrossRef]
  28. Zhang, L.; Cole, J.M. Anchoring Groups for Dye-Sensitized Solar Cells. ACS Appl. Mater. Interfaces 2015, 7, 3427–3455. [Google Scholar] [CrossRef]
  29. Liao, K.; Yang, J.-A.; Li, C.; Li, T.S.; Hao, F. Off-Stoichiometric Methylammonium Iodide Passivated Large-Grain Perovskite Film in Ambient Air for Efficient Inverted Solar Cells. ACS Appl. Mater. Interfaces 2019, 11, 39882–39889. [Google Scholar] [CrossRef]
  30. Dänekamp, B.; Droseros, N.; Palazon, F.; Sessolo, M.; Banerji, N.; Bolink, H.J. Efficient Photo- and Electroluminescence by Trap States Passivation in Vacuum-Deposited Hybrid Perovskite Thin Films. ACS Appl. Mater. Interfaces 2018, 10, 36187–36193. [Google Scholar] [CrossRef] [Green Version]
  31. Xie, Y.; Shao, F.; Wang, Y.; Xu, T.; Wang, D.; Huang, F. Enhanced Performance of Perovskite CH3NH3PbI3 Solar Cell by Using CH3NH3I as Additive in Sequential Deposition. ACS Appl. Mater. Interfaces 2015, 7, 12937–12942. [Google Scholar] [CrossRef]
  32. Yang, Y.; Song, J.; Zhao, Y.; Zhu, L.; Gu, X.; Gu, Y.; Che, M.; Qiang, Y. Ammonium-iodide-salt additives induced photovoltaic performance enhancement in one-step solution process for perovskite solar cells. J. Alloys. Compd. 2016, 684, 84–90. [Google Scholar] [CrossRef]
  33. Xu, C.; Zhang, Z.; Hu, Y.; Sheng, Y.; Jiang, P.; Han, H.; Zhang, J. Printed hole-conductor-free mesoscopic perovskite solar cells with excellent long-term stability using PEAI as an additive. J. Energy Chem. 2018, 27, 764–768. [Google Scholar] [CrossRef] [Green Version]
  34. Wang, Y.; Liu, S.; Zeng, Q.; Wang, R.; Qin, W.; Cao, H.; Yang, L.; Li, L.; Yin, S.; Zhang, F. Enhanced performance and stability of inverted planar perovskite solar cells by incorporating 1,6-diaminohexane dihydrochloride additive. Sol. Energy Mater. Sol. Cells 2018, 188, 140–148. [Google Scholar] [CrossRef]
  35. Laskar, A.R.; Luo, W.; Ghimire, N.; Chowdhury, A.H.; Bahrami, B.; Gurung, A.; Reza, K.M.; Pathak, R.; Bobba, R.S.; Lamsal, B.S.; et al. Phenylhydrazinium Iodide for Surface Passivation and Defects Suppression in Perovskite Solar Cells. Adv. Funct. Mater. 2020, 30, 1–11. [Google Scholar] [CrossRef]
  36. Du, C.; Wang, S.; Miao, X.; Sun, W.; Zhu, Y.; Wang, C.; Ma, R. Polyvinylpyrrolidone as additive for perovskite solar cells with water and isopropanol as solvents. Beilstein J. Nanotechnol. 2019, 10, 2374–2382. [Google Scholar] [CrossRef] [Green Version]
  37. Wang, N.; Liu, Z.; Zhou, Z.; Zhu, H.; Zhou, Y.; Huang, C.; Wang, Z.; Xu, H.; Jin, Y.; Fan, B.; et al. Reproducible One-Step Fabrication of Compact MAPbI3−xClx Thin Films Derived from Mixed-Lead-Halide Precursors. Chem. Mater. 2014, 26, 7145–7150. [Google Scholar] [CrossRef]
  38. Yu, H.; Wang, F.; Xie, F.; Li, W.; Chen, J.; Zhao, N. The Role of Chlorine in the Formation Process of “CH3NH3PbI3−xClx” Perovskite. Adv. Funct. Mater. 2014, 24, 7102–7108. [Google Scholar] [CrossRef]
  39. Yefang, J.; Ru, D.; Xuediao, C.; Jiangshan, F.; Zhike, L.; Shengzhong, L. Effect of Amphiphilic Quaternary Ammonium Salt Additive on Performance and Stability of Perovskite Solar Cells. J. Chin. Univ. 2019, 40, 1697–1705. [Google Scholar] [CrossRef]
  40. Mateen, M.; Arain, Z.; Liu, X.; Iqbal, A.; Ren, Y.; Zhang, X.; Liu, C.; Chen, Q.; Ma, S.; Ding, Y.; et al. Boosting optoelectronic performance of MAPbI3 perovskite solar cells via ethylammonium chloride additive engineering. Sci. China Mater. 2020, 63, 2477–2486. [Google Scholar] [CrossRef]
  41. Yao, J.; Wang, H.; Wang, P.; Gurney, R.S.; Intaniwet, A.; Ruankham, P.; Choopun, S.; Liu, D.; Wang, T. Trap passivation and efficiency improvement of perovskite solar cells by a guanidinium additive. Mater. Chem. Front. 2019, 3, 1357–1364. [Google Scholar] [CrossRef]
  42. Wang, P.; Wang, H.; Ye, F.; Zhang, H.; Chen, M.; Cai, J.; Li, D.; Liu, D.; Wang, T. Contrasting Effects of Organic Chloride Additives on Performance of Direct and Inverted Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2019, 11, 37833–37841. [Google Scholar] [CrossRef] [PubMed]
  43. Zheng, G.; Li, L.; Wang, L.; Gao, X.; Zhou, H. The investigation of an amidine-based additive in the perovskite films and solar cells. J. Semicond. 2017, 38, 14001. [Google Scholar] [CrossRef]
  44. Bae, S.; Jo, J.W.; Lee, P.; Ko, M.J. Controlling the Morphology of Organic–Inorganic Hybrid Perovskites through Dual Additive-Mediated Crystallization for Solar Cell Applications. ACS Appl. Mater. Interfaces 2019, 11, 17452–17458. [Google Scholar] [CrossRef] [PubMed]
  45. Wang, Y.; Song, N.; Feng, L.; Deng, X. Effects of Organic Cation Additives on the Fast Growth of Perovskite Thin Films for Efficient Planar Heterojunction Solar Cells. ACS Appl. Mater. Interfaces 2016, 8, 24703–24711. [Google Scholar] [CrossRef] [PubMed]
  46. Chen, C.-C.; Hong, Z.; Li, G.; Chen, Q.; Zhou, H.; Yang, Y. One-step, low-temperature deposited perovskite solar cell utilizing small molecule additive. J. Photonics Energy 2015, 5, 057405. [Google Scholar] [CrossRef]
  47. Mangrulkar, M.; Luchkin, S.Y.; Boldyreva, A.G.; Troshin, P.A.; Stevenson, K.J. Influence of pyridine-based ligands on photostability of MAPbI3 thin films. Mendeleev Commun. 2021, 31, 319–322. [Google Scholar] [CrossRef]
  48. Frolova, L.A.; Davlethanov, A.I.; Dremova, N.N.; Zhidkov, I.S.; Akbulatov, A.F.; Kurmaev, E.Z.; Aldoshin, S.M.; Stevenson, K.J.; Troshin, P. Efficient and Stable MAPbI3-Based Perovskite Solar Cells Using Polyvinylcarbazole Passivation. J. Phys. Chem. Lett. 2020, 11, 6772–6778. [Google Scholar] [CrossRef]
  49. Zhen, J.; Zhou, W.; Chen, M.; Li, B.; Jia, L.; Wang, M.; Yang, S. Pyridine-functionalized fullerene additive enabling coordination interactions with CH3NH3PbI3 perovskite towards highly efficient bulk heterojunction solar cells. J. Mater. Chem. A 2019, 7, 2754–2763. [Google Scholar] [CrossRef]
  50. Fu, S.; Li, X.; Wan, L.; Wu, Y.; Zhang, W.; Wang, Y.; Bao, Q.; Fang, J. Efficient Passivation with Lead Pyridine-2-Carboxylic for High-Performance and Stable Perovskite Solar Cells. Adv. Energy Mater. 2019, 9, 1–10. [Google Scholar] [CrossRef]
  51. Cheng, F.; Jing, X.; Chen, R.; Cao, J.; Yan, J.; Wu, Y.; Huang, X.; Wu, B.; Zheng, N. N-Methyl-2-pyrrolidone as an excellent coordinative additive with a wide operating range for fabricating high-quality perovskite films. Inorg. Chem. Front. 2019, 6, 2458–2463. [Google Scholar] [CrossRef]
  52. Lee, S.; Tang, M.-C.; Munir, R.; Barrit, D.; Kim, Y.-J.; Kang, R.; Yun, J.-M.; Smilgies, D.-M.; Amassian, A.; Kim, D.-Y. In situ study of the film formation mechanism of organic–inorganic hybrid perovskite solar cells: Controlling the solvate phase using an additive system. J. Mater. Chem. A 2020, 8, 7695–7703. [Google Scholar] [CrossRef]
  53. Xia, R.; Fei, Z.; Drigo, N.; Bobbink, F.D.; Huang, Z.; Jasiūnas, R.; Franckevičius, M.; Gulbinas, V.; Mensi, M.; Fang, X.; et al. Retarding Thermal Degradation in Hybrid Perovskites by Ionic Liquid Additives. Adv. Funct. Mater. 2019, 29. [Google Scholar] [CrossRef]
  54. Luo, C.; Li, G.; Chen, L.; Dong, J.; Yu, M.; Xu, C.; Yao, Y.; Wang, M.; Song, Q.; Zhang, S. Passivation of defects in inverted perovskite solar cells using an imidazolium-based ionic liquid. Sustain. Energy Fuels 2020, 4, 3971–3978. [Google Scholar] [CrossRef]
  55. Zheng, X.; Jiang, T.; Bai, L.; Chen, X.; Chen, Z.; Xu, X.; Song, D.; Xu, X.; Li, B.; Yang, Y. Enhanced thermal stability of inverted perovskite solar cells by interface modification and additive strategy. RSC Adv. 2020, 10, 18400–18406. [Google Scholar] [CrossRef]
  56. Chen, S.; Pan, Q.; Li, J.; Zhao, C.; Guo, X.; Zhao, Y.; Jiu, T. Grain boundary passivation with triazine-graphdiyne to improve perovskite solar cell performance. Sci. China Mater. 2020, 63, 2465–2476. [Google Scholar] [CrossRef]
  57. Liao, J.-F.; Wu, W.-Q.; Zhong, J.-X.; Jiang, Y.; Wang, L.; Kuang, D.-B. Enhanced efficacy of defect passivation and charge extraction for efficient perovskite photovoltaics with a small open circuit voltage loss. J. Mater. Chem. A 2019, 7, 9025–9033. [Google Scholar] [CrossRef]
  58. Jiang, L.-L.; Wang, Z.-K.; Li, M.; Zhang, C.-C.; Ye, Q.-Q.; Hu, K.-H.; Lu, D.-Z.; Fang, P.; Liao, L.-S. Passivated Perovskite Crystallization via g-C3N4 for High-Performance Solar Cells. Adv. Funct. Mater. 2018, 28, 1–8. [Google Scholar] [CrossRef]
  59. Hu, L.; Liu, T.; Sun, L.; Xiong, S.; Qin, F.; Jiang, X.; Jiang, Y.; Zhou, Y. Suppressing generation of iodine impurity via an amidine additive in perovskite solar cells. Chem. Commun. 2018, 54, 4704–4707. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Li, L.; Chen, Y.; Liu, Z.; Chen, Q.; Wang, X.; Zhou, H. The Additive Coordination Effect on Hybrids Perovskite Crystallization and High-Performance Solar Cell. Adv. Mater. 2016, 28, 9862–9868. [Google Scholar] [CrossRef] [PubMed]
  61. Hao, X.; Bo-Xin, Z.; Wei, J.; Qing-Hong, Z.; Hua-Qing, X. Polymer PVP Additive for Improving Stability of Perovskite Solar Cells. J. Inorg. Mater. 2019, 34, 96–102. [Google Scholar] [CrossRef]
  62. Zhang, T.; Cao, Z.; Shang, Y.; Cui, C.; Fu, P.; Jiang, X.; Wang, F.; Xu, K.; Yin, D.; Qu, D.; et al. Multi-functional organic molecules for surface passivation of perovskite. J. Photochem. Photobiol. A Chem. 2018, 355, 42–47. [Google Scholar] [CrossRef]
  63. Zhi, L.; Li, Y.; Cao, X.; Li, Y.; Cui, X.; Ci, L.; Wei, J. Perovskite Solar Cells Fabricated by Using an Environmental Friendly Aprotic Polar Additive of 1,3-Dimethyl-2-imidazolidinone. Nanoscale Res. Lett. 2017, 12, 632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Gao, L.; Wang, L.; Ding, X.; Zhao, E.; Yang, S.; Zhao, Y.; Li, Y.; Wang, S.; Ma, T. Incredible PCE enhancement induced by damaged perovskite layers: Deeply understanding the working principle of additives in bulk heterojunction perovskite solar cells. J. Mater. Chem. A 2018, 6, 4365–4373. [Google Scholar] [CrossRef]
  65. Li, Y.; Li, L.; Yerramilli, A.; Chen, Y.; Fang, D.; Shen, Y.; Alford, T. Enhanced power conversion efficiency and preferential orientation of the MAPbI3 perovskite solar cells by introduction of urea as additive. Org. Electron. 2019, 73, 130–136. [Google Scholar] [CrossRef]
  66. Han, L.; Cong, S.; Yang, H.; Lou, Y.; Wang, H.; Huang, J.; Zhu, J.; Wu, Y.; Chen, Q.; Zhang, B.; et al. Environmental-Friendly Urea Additive Induced Large Perovskite Grains for High Performance Inverted Solar Cells. Sol. RRL 2018, 2, 1–9. [Google Scholar] [CrossRef]
  67. Wen, X.; Cai, Q.; Shen, G.; Xu, X.; Dong, P.; Du, Y.; Dong, H.; Mu, C. Enhanced crystallization of solution-processed perovskite using urea as an additive for large-grain MAPbI3 perovskite solar cells. Nanotechnology 2021, 32, 30LT02. [Google Scholar] [CrossRef]
  68. Lee, J.-W.; Bae, S.-H.; Hsieh, Y.-T.; De Marco, N.; Wang, M.; Sun, P.; Yang, Y. A Bifunctional Lewis Base Additive for Microscopic Homogeneity in Perovskite Solar Cells. Chem 2017, 3, 290–302. [Google Scholar] [CrossRef]
  69. Liu, B.; Wang, S.; Ma, Z.; Ma, J.; Ma, R.; Wang, C. High-performance perovskite solar cells with large grain-size obtained by the synergy of urea and dimethyl sulfoxide. Appl. Surf. Sci. 2019, 467–468, 708–714. [Google Scholar] [CrossRef]
  70. Liu, S.; Li, S.; Wu, J.; Wang, Q.; Ming, Y.; Zhang, D.; Sheng, Y.; Hu, Y.; Rong, Y.; Mei, A.; et al. Amide Additives Induced a Fermi Level Shift To Improve the Performance of Hole-Conductor-Free, Printable Mesoscopic Perovskite Solar Cells. J. Phys. Chem. Lett. 2019, 10, 6865–6872. [Google Scholar] [CrossRef]
  71. Shi, X.; Wu, Y.; Chen, J.; Cai, M.; Yang, Y.; Liu, X.; Tao, Y.; Guli, M.; Ding, Y.; Dai, S. Thermally stable perovskite solar cells with efficiency over 21% via a bifunctional additive. J. Mater. Chem. A 2020, 8, 7205–7213. [Google Scholar] [CrossRef]
  72. Lv, Y.; Zhang, H.; Wang, J.; Chen, L.; Bian, L.; An, Z.; Qian, Z.; Ren, G.; Wu, J.; Nüesch, F.; et al. All-in-One Deposition to Synergistically Manipulate Perovskite Growth for High-Performance Solar Cell. Research 2020, 2020, 1–10. [Google Scholar] [CrossRef]
  73. Qin, C.; Matsushima, T.; Fujihara, T.; Adachi, C. Multifunctional Benzoquinone Additive for Efficient and Stable Planar Perovskite Solar Cells. Adv. Mater. 2017, 29, 1–8. [Google Scholar] [CrossRef]
  74. Yu, W.; Yu, S.; Zhang, J.; Liang, W.; Wang, X.; Guo, X.; Li, C. Two-in-one additive-engineering strategy for improved air stability of planar perovskite solar cells. Nano Energy 2018, 45, 229–235. [Google Scholar] [CrossRef]
  75. Kuo, D.-W.; Liu, G.-Z.; Lee, R.-H. Star-shaped molecule with planar triazine core and perylene diimide branches as an n-type additive for bulk-heterojunction perovskite solar cells. Dye. Pigment. 2019, 170, 107562. [Google Scholar] [CrossRef]
  76. Li, H.; Zhu, K.; Zhang, K.; Huang, P.; Li, D.; Yuan, L.; Cao, T.; Sun, Z.; Li, Z.; Chen, Q.; et al. 3,4-Dihydroxybenzhydrazide as an additive to improve the morphology of perovskite films for efficient and stable perovskite solar cells. Org. Electron. 2019, 66, 47–52. [Google Scholar] [CrossRef]
  77. Xiong, S.; Song, J.; Yang, J.; Xu, J.; Zhang, M.; Ma, R.; Li, D.; Liu, X.; Liu, F.; Duan, C.; et al. Defect-Passivation Using Organic Dyes for Enhanced Efficiency and Stability of Perovskite Solar Cells. Sol. RRL 2020, 4, 1–9. [Google Scholar] [CrossRef]
  78. Lan, Y.; Wang, Y.; Song, Y. Efficient flexible perovskite solar cells based on a polymer additive. Flex. Print. Electron. 2020, 5, 014001. [Google Scholar] [CrossRef]
  79. Song, C.; Li, X.; Wang, Y.; Fu, S.; Wan, L.; Liu, S.; Zhang, W.; Song, W.; Fang, J. Sulfonyl-based non-fullerene electron acceptor-assisted grain boundary passivation for efficient and stable perovskite solar cells. J. Mater. Chem. A 2019, 7, 19881–19888. [Google Scholar] [CrossRef]
  80. Cao, X.B.; Li, C.; Zhi, L.L.; Li, Y.H.; Cui, X.; Yao, Y.W.; Ci, L.J.; Wei, J.Q. Fabrication of high quality perovskite films by modulating the Pb–O bonds in Lewis acid–base adducts. J. Mater. Chem. A 2017, 5, 8416–8422. [Google Scholar] [CrossRef]
  81. Cao, X.; Zhi, L.; Li, Y.; Cui, X.; Ci, L.; Ding, K.; Wei, J. Enhanced performance of perovskite solar cells by strengthening a self-embedded solvent annealing effect in perovskite precursor films. RSC Adv. 2017, 7, 49144–49150. [Google Scholar] [CrossRef] [Green Version]
  82. Gregorio, R.; Borges, D.S. Effect of crystallization rate on the formation of the polymorphs of solution cast poly(vinylidene fluoride). Polymer 2008, 49, 4009–4016. [Google Scholar] [CrossRef]
  83. Hao, F.; Stoumpos, C.; Guo, P.; Zhou, N.; Marks, T.J.; Chang, R.P.H.; Kanatzidis, M.G. Solvent-Mediated Crystallization of CH3NH3SnI3 Films for Heterojunction Depleted Perovskite Solar Cells. J. Am. Chem. Soc. 2015, 137, 11445–11452. [Google Scholar] [CrossRef]
  84. Cao, X.; Zhi, L.; Li, Y.; Fang, F.; Cui, X.; Yao, Y.; Ci, L.; Ding, K.; Wei, J. Control of the morphology of PbI 2 films for efficient perovskite solar cells by strong Lewis base additives. J. Mater. Chem. C 2017, 5, 7458–7464. [Google Scholar] [CrossRef]
  85. Ren, Y.; Ding, X.; Zhu, J.; Hayat, T.; Alsaedi, A.; Li, Z.; Xu, X.; Ding, Y.; Yang, S.; Kong, M.; et al. A Bi-functional additive for linking PI 2 and decreasing defects in organo-halide perovskites. J. Alloys Compd. 2018, 758, 171–176. [Google Scholar] [CrossRef]
  86. Foley, B.J.; Girard, J.; Sorenson, B.A.; Chen, A.Z.; Niezgoda, J.S.; Alpert, M.; Harper, A.; Smilgies, D.-M.; Clancy, P.; Saidi, W.A.; et al. Controlling nucleation, growth, and orientation of metal halide perovskite thin films with rationally selected additives. J. Mater. Chem. A 2017, 5, 113–123. [Google Scholar] [CrossRef]
  87. Gutwald, M.; Rolston, N.; Printz, A.D.; Zhao, O.; Elmaraghi, H.; Ding, Y.; Zhang, J.; Dauskardt, R.H. Perspectives on intrinsic toughening strategies and passivation of perovskite films with organic additives. Sol. Energy Mater. Sol. Cells 2020, 209, 110433. [Google Scholar] [CrossRef]
  88. Lin, C.-T.; De Rossi, F.; Kim, J.; Baker, J.; Ngiam, J.; Xu, B.; Pont, S.; Aristidou, N.; Haque, S.A.; Watson, T.; et al. Evidence for surface defect passivation as the origin of the remarkable photostability of unencapsulated perovskite solar cells employing aminovaleric acid as a processing additive. J. Mater. Chem. A 2019, 7, 3006–3011. [Google Scholar] [CrossRef] [Green Version]
  89. Santhosh, N.; Sitaaraman, S.; Pounraj, P.; Govindaraj, R.; Pandian, M.S.; Ramasamy, P. Fabrication of hole-transport-free perovskite solar cells using 5-ammonium valeric acid iodide as additive and carbon as counter electrode. Mater. Lett. 2019, 236, 706–709. [Google Scholar] [CrossRef]
  90. Li, N.; Xu, F.; Qiu, Z.; Liu, J.; Wan, X.; Zhu, X.; Yu, H.; Li, C.; Liu, Y.; Cao, B. Sealing the domain boundaries and defects passivation by Poly(acrylic acid) for scalable blading of efficient perovskite solar cells. J. Power Sources 2019, 426, 188–196. [Google Scholar] [CrossRef]
  91. Chen, H.; Liu, T.; Zhou, P.; Li, S.; Ren, J.; He, H.; Wang, J.; Wang, N.; Guo, S. Efficient Bifacial Passivation with Crosslinked Thioctic Acid for High-Performance Methylammonium Lead Iodide Perovskite Solar Cells. Adv. Mater. 2020, 32, 1905661. [Google Scholar] [CrossRef]
  92. Zhang, C.; Luo, Q.; Deng, X.; Zheng, J.; Ou-Yang, W.; Chen, X.; Huang, S. Enhanced efficiency and stability of carbon based perovskite solar cells using terephthalic acid additive. Electrochim. Acta 2017, 258, 1262–1272. [Google Scholar] [CrossRef]
  93. Su, L.; Xiao, Y.; Han, G.; Lu, L.; Li, H.; Zhu, M. Performance enhancement of perovskite solar cells using trimesic acid additive in the two-step solution method. J. Power Sources 2019, 426, 11–15. [Google Scholar] [CrossRef]
  94. Su, L.; Xiao, Y.; Lu, L.; Han, G.; Zhu, M. Enhanced stability and solar cell performance via π-conjugated Lewis base passivation of organic inorganic lead halide perovskites. Org. Electron. 2020, 77, 105519. [Google Scholar] [CrossRef]
  95. Yao, X.; Zheng, L.; Zhang, X.; Xu, W.; Hu, W.; Gong, X. Efficient Perovskite Solar Cells through Suppressed Nonradiative Charge Carrier Recombination by a Processing Additive. ACS Appl. Mater. Interfaces 2019, 11, 40163–40171. [Google Scholar] [CrossRef]
  96. Pockett, A.; Raptis, D.; Meroni, S.M.P.; Baker, J.; Watson, T.M.; Carnie, M. Origin of Exceptionally Slow Light Soaking Effect in Mesoporous Carbon Perovskite Solar Cells with AVA Additive. J. Phys. Chem. C 2019, 123, 11414–11421. [Google Scholar] [CrossRef] [Green Version]
  97. Han, F.; Luo, J.; Malik, H.A.; Zhao, B.; Wan, Z.; Jia, C. A functional sulfonic additive for high efficiency and low hysteresis perovskite solar cells. J. Power Sources 2017, 359, 577–584. [Google Scholar] [CrossRef]
  98. Gao, C.; Dong, H.; Bao, X.; Zhang, Y.; Saparbaev, A.; Yu, L.; Wen, S.; Yang, R.; Dong, L. Additive engineering to improve the efficiency and stability of inverted planar perovskite solar cells. J. Mater. Chem. C 2018, 6, 8234–8241. [Google Scholar] [CrossRef]
  99. Wang, M.; Li, B.; Siffalovic, P.; Chen, L.-C.; Cao, G.; Tian, J. Monolayer-like hybrid halide perovskite films prepared by additive engineering without antisolvents for solar cells. J. Mater. Chem. A 2018, 6, 15386–15394. [Google Scholar] [CrossRef]
  100. Wu, Y.; Xie, F.; Chen, H.; Yang, X.; Su, H.; Cai, M.; Zhou, Z.; Noda, T.; Han, L. Thermally Stable MAPbI3Perovskite Solar Cells with Efficiency of 19.19% and Area over 1 cm2achieved by Additive Engineering. Adv. Mater. 2017, 29, 1–8. [Google Scholar] [CrossRef] [PubMed]
  101. Wu, Q.; Zhou, P.; Zhou, W.; Wei, X.; Chen, T.; Yang, S. Acetate Salts as Nonhalogen Additives To Improve Perovskite Film Morphology for High-Efficiency Solar Cells. ACS Appl. Mater. Interfaces 2016, 8, 15333–15340. [Google Scholar] [CrossRef] [PubMed]
  102. Zhang, Z.; Fan, W.; Wei, X.; Zhang, L.; Yang, Z.; Wei, Z.; Shen, T.; Si, H.; Qi, J. Promoted performance of carbon based perovskite solar cells by environmentally friendly additives of CH3COONH4 and Zn(CH3COO)2. J. Alloys Compd. 2019, 802, 694–703. [Google Scholar] [CrossRef]
  103. Tang, G.; You, P.; Tai, Q.; Wu, R.; Yan, F. Performance Enhancement of Perovskite Solar Cells Induced by Lead Acetate as an Additive. Sol. RRL 2018, 2, 1–9. [Google Scholar] [CrossRef]
  104. Wang, Y.; Wu, Y.; Fu, S.; Song, C.; Wan, L.; Zhang, W.; Li, X.; Yang, W.; Song, W.; Fang, J. Barium acetate as an additive for high performance perovskite solar cells. J. Mater. Chem. C 2019, 7, 11411–11418. [Google Scholar] [CrossRef]
  105. Guan, Y.; Mei, A.; Rong, Y.; Duan, M.; Hou, X.; Hu, Y.; Han, H. Fullerene derivative as an additive for highly efficient printable mesoscopic perovskite solar cells. Org. Electron. 2018, 62, 653–659. [Google Scholar] [CrossRef]
  106. Hu, L.; Li, S.; Zhang, L.; Liu, Y.; Zhang, C.; Wu, Y.; Sun, Q.; Cui, Y.; Zhu, F.; Hao, Y.; et al. Unravelling the role of C60 derivatives as additives into active layers for achieving high-efficiency planar perovskite solar cells. Carbon 2020, 167, 160–168. [Google Scholar] [CrossRef]
  107. Chen, N.; Yi, X.; Zhuang, J.; Wei, Y.; Zhang, Y.; Wang, F.; Cao, S.; Li, C.; Wang, J. An Efficient Trap Passivator for Perovskite Solar Cells: Poly(propylene glycol) bis(2-aminopropyl ether). Nano-Micro Lett. 2020, 12, 1–13. [Google Scholar] [CrossRef]
  108. Hsu, H.-L.; Jiang, B.-H.; Chung, C.-L.; Yu, Y.-Y.; Jeng, R.-J.; Chen, C.-P. Commercially available jeffamine additives for p–i–n perovskite solar cells. Nanotechnol. 2020, 31, 274002. [Google Scholar] [CrossRef]
  109. Yang, J.; Xiong, S.; Qu, T.; Zhang, Y.; He, X.; Guo, X.; Zhao, Q.; Braun, S.; Chen, J.; Xu, J.; et al. Extremely Low-Cost and Green Cellulose Passivating Perovskites for Stable and High-Performance Solar Cells. ACS Appl. Mater. Interfaces 2019, 11, 13491–13498. [Google Scholar] [CrossRef]
  110. Guan, M.; Zhang, Q.; Wang, F.; Liu, H.; Zhao, J.; Jia, C.; Chen, Y. Employing tetraethyl orthosilicate additive to enhance trap passivation of planar perovskite solar cells. Electrochim. Acta 2019, 293, 174–183. [Google Scholar] [CrossRef]
  111. De Carvalho, B.A.; Kavadiya, S.; Huang, S.; Niedzwiedzki, D.; Biswas, P. Highly Stable Perovskite Solar Cells Fabricated Under Humid Ambient Conditions. IEEE J. Photovoltaics 2017, 7, 532–538. [Google Scholar] [CrossRef]
  112. Isa, M.J.A.; Sulistianto, J.; Kevin, L.; Poespawati, N.R. Optimization of Perovskite Solar Cell Performance using Optimized Level of Tetraethyl Orthosilicate Concentration. In Proceedings of the 2019 11th International Conference on Information Technology and Electrical Engineering (ICITEE), Pattaya, Thailand, 10–11 October 2019; Volume 7. [Google Scholar] [CrossRef]
  113. Chaudhary, B.; Kulkarni, A.; Jena, A.K.; Ikegami, M.; Miyasaka, T. Tetrahydrofuran as an Oxygen Donor Additive to Enhance Stability and Reproducibility of Perovskite Solar Cells Fabricated in High Relative Humidity (50%) Atmosphere. Energy Technol. 2020, 8, 1–8. [Google Scholar] [CrossRef]
  114. Wang, H.; Zeng, W.; Xia, R. Antisolvent diethyl ether as additive to enhance the performance of perovskite solar cells. Thin Solid Film. 2018, 663, 9–13. [Google Scholar] [CrossRef]
  115. Kumar, S.; Choi, Y.; Kang, S.-H.; Oh, N.K.; Lee, J.; Seo, J.; Jeong, M.; Kwon, H.W.; Seok, S.I.; Yang, C.; et al. Multifaceted Role of a Dibutylhydroxytoluene Processing Additive in Enhancing the Efficiency and Stability of Planar Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2019, 11, 38828–38837. [Google Scholar] [CrossRef]
  116. Feng, M.; You, S.; Cheng, N.; Du, J. High quality perovskite film solar cell using methanol as additive with 19.5% power conversion efficiency. Electrochim. Acta 2019, 293, 356–363. [Google Scholar] [CrossRef]
  117. You, S.; Bi, S.; Qiushuang, J.; Jia, Q.; Yuan, Y.; Xia, Y.; Xiao, Z.; Sun, Z.; Liu, J.; Sun, S.; et al. Additive-Enhanced Crystallization of Solution Process for Planar Perovskite Solar Cells with Efficiency Exceeding 19 %. Chem. A Eur. J. 2017, 23, 18140–18145. [Google Scholar] [CrossRef] [PubMed]
  118. Ugur, E.; Sheikh, A.D.; Munir, R.; Khan, J.I.; Barrit, D.; Amassian, A.; Laquai, F. Improved Morphology and Efficiency of n–i–p Planar Perovskite Solar Cells by Processing with Glycol Ether Additives. ACS Energy Lett. 2017, 2, 1960–1968. [Google Scholar] [CrossRef] [Green Version]
  119. Wu, W.; Li, H.; Liu, S.; Zheng, B.; Xue, Y.; Liu, X.; Gao, C. Tuning PbI2layers by n-butanol additive for improving CH3NH3PbI3light harvesters of perovskite solar cells. RSC Adv. 2016, 6, 89609–89613. [Google Scholar] [CrossRef]
  120. Balis, N.; Zaky, A.A.; Athanasekou, C.; Silva, A.; Sakellis, E.; Vasilopoulou, M.; Stergiopoulos, T.; Kontos, A.G.; Falaras, P. Investigating the role of reduced graphene oxide as a universal additive in planar perovskite solar cells. J. Photochem. Photobiol. A Chem. 2020, 386, 112141. [Google Scholar] [CrossRef]
  121. Yu, Y.-Y.; Tseng, C.; Chien, W.-C.; Hsu, H.-L.; Chen, C.-P. Photovoltaic Performance Enhancement of Perovskite Solar Cells Using Polyimide and Polyamic Acid as Additives. J. Phys. Chem. C 2019, 123, 23826–23833. [Google Scholar] [CrossRef]
  122. Feng, J.; Zhu, X.; Yang, Z.; Zhang, X.; Niu, J.; Wang, Z.; Zuo, S.; Priya, S.; Liu, F.; Yang, D. Record Efficiency Stable Flexible Perovskite Solar Cell Using Effective Additive Assistant Strategy. Adv. Mater. 2018, 30, 1801418. [Google Scholar] [CrossRef]
  123. Hsieh, C.-M.; Liao, Y.-S.; Lin, Y.-R.; Chen, C.-P.; Tsai, C.-M.; Diau, E.W.-G.; Chuang, S.-C. Low-temperature, simple and efficient preparation of perovskite solar cells using Lewis bases urea and thiourea as additives: Stimulating large grain growth and providing a PCE up to 18.8%. RSC Adv. 2018, 8, 19610–19615. [Google Scholar] [CrossRef] [Green Version]
  124. Cui, C.; Xie, D.; Lin, P.; Hu, H.; Che, S.; Xiao, K.; Wang, P.; Xu, L.; Yang, D.; Yu, X. Thioacetamide additive assisted crystallization of solution-processed perovskite films for high performance planar heterojunction solar cells. Sol. Energy Mater. Sol. Cells 2020, 208, 110435. [Google Scholar] [CrossRef]
  125. Sun, M.; Zhang, F.; Liu, H.; Li, X.; Xiao, Y.; Wang, S. Tuning the crystal growth of perovskite thin-films by adding the 2-pyridylthiourea additive for highly efficient and stable solar cells prepared in ambient air. J. Mater. Chem. A 2017, 5, 13448–13456. [Google Scholar] [CrossRef]
  126. Gao, Y.; Wu, Y.; Liu, Y.; Chen, C.; Bai, X.; Yang, L.; Shi, Z.; Yu, W.W.; Dai, Q.; Zhang, Y. Dual Functions of Crystallization Control and Defect Passivation Enabled by an Ionic Compensation Strategy for Stable and High-Efficient Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2019, 12, 3631–3641. [Google Scholar] [CrossRef]
  127. Ke, W.; Xiao, C.; Wang, C.; Saparov, B.; Duan, H.-S.; Zhao, D.; Xiao, Z.; Schulz, P.; Harvey, S.P.; Liao, W.; et al. Employing Lead Thiocyanate Additive to Reduce the Hysteresis and Boost the Fill Factor of Planar Perovskite Solar Cells. Adv. Mater. 2016, 28, 5214–5221. [Google Scholar] [CrossRef] [PubMed]
  128. Kim, M.K.; Jeon, T.; Park, H.I.; Lee, J.M.; Nam, S.A.; Kim, S.O. Effective control of crystal grain size in CH3NH3PbI3 perovskite solar cells with a pseudohalide Pb(SCN)2additive. CrystEngComm 2016, 18, 6090–6095. [Google Scholar] [CrossRef]
  129. Zhang, R.; Li, M.; Huan, Y.; Xi, J.; Zhang, S.; Cheng, X.; Wu, H.; Peng, W.; Bai, Z.; Yan, X. A potassium thiocyanate additive for hysteresis elimination in highly efficient perovskite solar cells. Inorg. Chem. Front. 2019, 6, 434–442. [Google Scholar] [CrossRef]
  130. Zhang, H.; Hou, M.; Xia, Y.; Wei, Q.; Wang, Z.; Cheng, Y.; Chen, Y.; Huang, W. Synergistic effect of anions and cations in additives for highly efficient and stable perovskite solar cells. J. Mater. Chem. A 2018, 6, 9264–9270. [Google Scholar] [CrossRef]
  131. Han, Q.; Bai, Y.; Liu, J.; Du, K.; Li, T.; Ji, D.; Zhou, Y.; Cao, C.; Shin, D.; Ding, J.; et al. Additive engineering for high-performance room-temperature-processed perovskite absorbers with micron-size grains and microsecond-range carrier lifetimes. Energy Environ. Sci. 2017, 10, 2365–2371. [Google Scholar] [CrossRef]
  132. Cheng, N.; Li, W.; Zhang, M.; Wu, H.; Sun, S.; Zhao, Z.; Xiao, Z.; Sun, Z.; Zi, W.; Fang, L. Enhance the performance and stability of methylammonium lead iodide perovskite solar cells with guanidinium thiocyanate additive. Curr. Appl. Phys. 2019, 19, 25–30. [Google Scholar] [CrossRef]
  133. Zou, J.; Liu, W.; Deng, W.; Lei, G.; Zeng, S.; Xiong, J.; Gu, H.; Hu, Z.; Wang, X.; Li, J. An efficient guanidinium isothiocyanate additive for improving the photovoltaic performances and thermal stability of perovskite solar cells. Electrochim. Acta 2018, 291, 297–303. [Google Scholar] [CrossRef]
  134. Correa-Baena, J.-P.; Anaya, M.; Lozano, G.; Tress, W.; Domanski, K.; Saliba, M.; Matsui, T.; Jacobsson, J.; Calvo, M.; Abate, A.; et al. Unbroken Perovskite: Interplay of Morphology, Electro-optical Properties, and Ionic Movement. Adv. Mater. 2016, 28, 5031–5037. [Google Scholar] [CrossRef]
  135. Zhang, S.; Lu, Y.; Lin, B.; Zhu, Y.; Zhang, K.; Yuan, N.-Y.; Ding, J.-N.; Fang, B. PVDF-HFP additive for visible-light-semitransparent perovskite films yielding enhanced photovoltaic performance. Sol. Energy Mater. Sol. Cells 2017, 170, 178–186. [Google Scholar] [CrossRef]
  136. Sun, C.; Guo, Y.; Fang, B.; Yang, J.; Qin, B.; Duan, H.; Chen, Y.; Li, H.; Liu, H. Enhanced Photovoltaic Performance of Perovskite Solar Cells Using Polymer P(VDF-TrFE) as a Processed Additive. J. Phys. Chem. C 2016, 120, 12980–12988. [Google Scholar] [CrossRef]
  137. Eze, V.O.; Seike, Y.; Mori, T. Synergistic Effect of Additive and Solvent Vapor Annealing on the Enhancement of MAPbI3 Perovskite Solar Cells Fabricated in Ambient Air. ACS Appl. Mater. Interfaces 2020, 12, 46837–46845. [Google Scholar] [CrossRef] [PubMed]
  138. AnkiReddy, K.; Ghahremani, A.H.; Martin, B.; Gupta, G.; Druffel, T. Rapid thermal annealing of CH3NH3PbI3 perovskite thin films by intense pulsed light with aid of diiodomethane additive. J. Mater. Chem. A 2018, 6, 9378–9383. [Google Scholar] [CrossRef]
  139. Ma, Y.; Zhang, H.; Zhang, Y.; Hu, R.; Jiang, M.; Zhang, R.; Lv, H.; Tian, J.; Chu, L.; Zhang, J.; et al. Enhancing the Performance of Inverted Perovskite Solar Cells via Grain Boundary Passivation with Carbon Quantum Dots. ACS Appl. Mater. Interfaces 2018, 11, 3044–3052. [Google Scholar] [CrossRef]
  140. Kırbıyık, Ç.; Toprak, A.; Başlak, C.; Kuş, M.; Ersöz, M. Nitrogen-doped CQDs to enhance the power conversion efficiency of perovskite solar cells via surface passivation. J. Alloys Compd. 2020, 832, 154897. [Google Scholar] [CrossRef]
  141. Guo, Q.; Yuan, F.; Zhang, B.; Zhou, S.; Zhang, J.; Bai, Y.; Fan, L.; Hayat, T.; Alsaedi, A.; Tan, Z. Passivation of the grain boundaries of CH3NH3PbI3 using carbon quantum dots for highly efficient perovskite solar cells with excellent environmental stability. Nanoscale 2019, 11, 115–124. [Google Scholar] [CrossRef] [PubMed]
  142. Hsu, H.-L.; Hsiao, H.-T.; Juang, T.-Y.; Jiang, B.-H.; Chen, S.-C.; Jeng, R.-J.; Chen, C.-P. Carbon Nanodot Additives Realize High-Performance Air-Stable p-i-n Perovskite Solar Cells Providing Efficiencies of up to 20.2%. Adv. Energy Mater. 2018, 8, 1–9. [Google Scholar] [CrossRef]
  143. Li, Z.; Wang, F.; Liu, C.; Gao, F.; Shen, L.; Guo, W. Efficient perovskite solar cells enabled by ion-modulated grain boundary passivation with a fill factor exceeding 84%. J. Mater. Chem. A 2019, 7, 22359–22365. [Google Scholar] [CrossRef]
  144. Wang, Y.; Zhang, J.; Chen, S.; Zhang, H.; Li, L.; Fu, Z. Surface passivation with nitrogen-doped carbon dots for improved perovskite solar cell performance. J. Mater. Sci. 2018, 53, 9180–9190. [Google Scholar] [CrossRef]
  145. Fang, X.; Ding, J.; Yuan, N.; Sun, P.; Lv, M.; Ding, G.; Zhu, C. Graphene quantum dot incorporated perovskite films: Passivating grain boundaries and facilitating electron extraction. Phys. Chem. Chem. Phys. 2017, 19, 6057–6063. [Google Scholar] [CrossRef] [PubMed]
  146. Brakkee, R.; Williams, R.M. Minimizing Defect States in Lead Halide Perovskite Solar Cell Materials. Appl. Sci. 2020, 10, 3061. [Google Scholar] [CrossRef]
  147. Soe, C.M.M.; Stoumpos, C.C.; Harutyunyan, B.; Manley, E.F.; Chen, L.X.; Bedzyk, M.J.; Marks, T.J.; Kanatzidis, M.G. Room Temperature Phase Transition in Methylammonium Lead Iodide Perovskite Thin Films Induced by Hydrohalic Acid Additives. ChemSusChem 2016, 9, 2656–2665. [Google Scholar] [CrossRef] [PubMed]
  148. Zhang, T.; Guo, N.; Li, G.; Qian, X.; Zhao, Y. A controllable fabrication of grain boundary PbI2 nanoplates passivated lead halide perovskites for high performance solar cells. Nano Energy 2016, 26, 50–56. [Google Scholar] [CrossRef]
  149. Wen, Y.; Tang, Y.-G.; Yan, G.-Q. Large grain size CH3NH3PbI3 film for perovskite solar cells with hydroic acid additive. AIP Adv. 2018, 8, 095226. [Google Scholar] [CrossRef] [Green Version]
  150. Abdi-Jalebi, M.; Dar, M.I.; Sadhanala, A.; Senanayak, S.P.; Franckevicius, M.; Arora, N.; Hu, Y.; Nazeeruddin, M.K.; Zakeeruddin, S.M.; Grätzel, M.; et al. Impact of Monovalent Cation Halide Additives on the Structural and Optoelectronic Properties of CH3NH3PbI3 Perovskite. Adv. Energy Mater. 2016, 6, 1502472. [Google Scholar] [CrossRef] [Green Version]
  151. Boopathi, K.M.; Mohan, R.; Huang, T.-Y.; Budiawan, W.; Lin, M.-Y.; Lee, C.-H.; Ho, K.-C.; Chu, C.-W. Synergistic improvements in stability and performance of lead iodide perovskite solar cells incorporating salt additives. J. Mater. Chem. A 2016, 4, 1591–1597. [Google Scholar] [CrossRef]
  152. Park, I.J.; Seo, S.; Park, M.A.; Lee, S.; Kim, D.H.; Zhu, K.; Shin, H.; Kim, J.Y. Effect of Rubidium Incorporation on the Structural, Electrical, and Photovoltaic Properties of Methylammonium Lead Iodide-Based Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2017, 9, 41898–41905. [Google Scholar] [CrossRef]
  153. Chan, S.-H.; Wu, M.-C.; Lee, K.-M.; Chen, W.-C.; Lin, T.-H.; Su, W.-F. Enhancing perovskite solar cell performance and stability by doping barium in methylammonium lead halide. J. Mater. Chem. A 2017, 5, 18044–18052. [Google Scholar] [CrossRef]
  154. Chen, C.; Xu, Y.; Wu, S.; Zhang, S.; Yang, Z.; Zhang, W.; Zhu, H.; Xiong, Z.; Chen, W.; Chen, W. CaI2: A more effective passivator of perovskite films than PbI2 for high efficiency and long-term stability of perovskite solar cells. J. Mater. Chem. A 2018, 6, 7903–7912. [Google Scholar] [CrossRef]
  155. Mabrouk, S.; Bahrami, B.; Gurung, A.; Reza, K.M.; Adhikari, N.; Dubey, A.; Pathak, R.; Yang, S.; Qiao, Q. Higher efficiency perovskite solar cells using additives of LiI, LiTFSI and BMImI in the PbI2 precursor. Sustain. Energy Fuels 2017, 1, 2162–2171. [Google Scholar] [CrossRef]
  156. Duan, B.; Ren, Y.; Xu, Y.; Chen, W.; Ye, Q.; Huang, Y.; Zhu, J.; Dai, S. Identification and characterization of a new intermediate to obtain high quality perovskite films with hydrogen halides as additives. Inorg. Chem. Front. 2017, 4, 473–480. [Google Scholar] [CrossRef]
  157. Chen, H.; Luo, Q.; Liu, T.; Ren, J.; Li, S.; Tai, M.; Lin, H.; He, H.; Wang, J.; Wang, N. Goethite Quantum Dots as Multifunctional Additives for Highly Efficient and Stable Perovskite Solar Cells. Small 2019, 15, e1904372. [Google Scholar] [CrossRef]
  158. Gong, X.; Guan, L.; Pan, H.; Sun, Q.; Zhao, X.; Li, H.; Pan, H.; Shen, Y.; Shao, Y.; Sun, L.; et al. Highly Efficient Perovskite Solar Cells via Nickel Passivation. Adv. Funct. Mater. 2018, 28, 1804286. [Google Scholar] [CrossRef]
  159. Kayesh, E.; Matsuishi, K.; Chowdhury, T.H.; Kaneko, R.; Noda, T.; Islam, A. Enhanced Photovoltaic Performance of Perovskite Solar Cells by Copper Chloride (CuCl2) as an Additive in Single Solvent Perovskite Precursor. Electron. Mater. Lett. 2018, 14, 712–717. [Google Scholar] [CrossRef]
  160. Almutawah, Z.S.; Watthage, S.C.; Song, Z.; Ahangharnejhad, R.H.; Subedi, K.K.; Shrestha, N.; Phillips, A.B.; Yan, Y.; Ellingson, R.J.; Heben, M.J. Enhanced Grain Size and Crystallinity in CH3NH3PbI3 Perovskite Films by Metal Additives to the Single-Step Solution Fabrication Process. MRS Adv. 2018, 3, 3237–3242. [Google Scholar] [CrossRef] [Green Version]
  161. Watthage, S.C.; Song, Z.; Shrestha, N.; Phillips, A.B.; Liyanage, G.K.; Roland, P.J.; Ellingson, R.J.; Heben, M.J. Impact of Divalent Metal Additives on the Structural and Optoelectronic Properties of CH3NH3PbI3 Perovskite Prepared by the Two-Step Solution Process. MRS Adv. 2017, 2, 1183–1188. [Google Scholar] [CrossRef]
  162. Liu, W.; Liu, N.; Ji, S.; Hua, H.; Ma, Y.; Hu, R.; Zhang, J.; Chu, L.; Li, X.; Huang, W. Perfection of Perovskite Grain Boundary Passivation by Rhodium Incorporation for Efficient and Stable Solar Cells. Nano-Micro Lett. 2020, 12, 119. [Google Scholar] [CrossRef]
  163. Carmona, C.R.; Gratia, P.; Zimmermann, I.; Grancini, G.; Gao, P.; Graetzel, M.; Nazeeruddin, M.K. High efficiency methylammonium lead triiodide perovskite solar cells: The relevance of non-stoichiometric precursors. Energy Environ. Sci. 2015, 8, 3550–3556. [Google Scholar] [CrossRef]
  164. Rafizadeh, S.; Wienands, K.; Schulze, P.S.C.; Bett, A.J.; Andreani, L.C.; Hermle, M.; Glunz, S.W.; Goldschmidt, J.C. Efficiency Enhancement and Hysteresis Mitigation by Manipulation of Grain Growth Conditions in Hybrid Evaporated–Spin-coated Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2019, 11, 722–729. [Google Scholar] [CrossRef] [PubMed]
  165. Chen, Y.; Yerramilli, A.; Shen, Y.; Zhao, Z.; Alford, T. Effect of excessive Pb content in the precursor solutions on the properties of the lead acetate derived CH3NH3PbI3 perovskite solar cells. Sol. Energy Mater. Sol. Cells 2018, 174, 478–484. [Google Scholar] [CrossRef]
  166. Yerramilli, A.; Chen, Y.; Sanni, D.; Asare, J.; Theodore, N.D.; Alford, T. Impact of excess lead on the stability and photo-induced degradation of lead halide perovskite solar cells. Org. Electron. 2018, 59, 107–112. [Google Scholar] [CrossRef]
  167. Barbé, J.; Newman, M.; Lilliu, S.; Kumar, V.; Lee, H.K.H.; Charbonneau, C.; Rodenburg, C.; Lidzey, D.G.; Tsoi, W.C. Localized effect of PbI2 excess in perovskite solar cells probed by high-resolution chemical–optoelectronic mapping. J. Mater. Chem. A 2018, 6, 23010–23018. [Google Scholar] [CrossRef]
  168. Li, Z.; Zhang, C.; Shao, Z.; Fan, Y.; Liu, R.; Wang, L.; Pang, S. Controlled surface decomposition derived passivation and energy-level alignment behaviors for high performance perovskite solar cells. J. Mater. Chem. A 2018, 6, 9397–9401. [Google Scholar] [CrossRef]
  169. Meier, T.; Gujar, T.P.; Schönleber, A.; Olthof, S.; Meerholz, K.; van Smaalen, S.; Panzer, F.; Thelakkat, M.; Köhler, A. Impact of excess PbI2 on the structure and the temperature dependent optical properties of methylammonium lead iodide perovskites. J. Mater. Chem. C 2018, 6, 7512–7519. [Google Scholar] [CrossRef]
  170. Gujar, T.P.; Unger, T.; Schönleber, A.; Fried, M.; Panzer, F.; van Smaalen, S.; Köhler, A.; Thelakkat, M. The role of PbI2 in CH3NH3PbI3 perovskite stability, solar cell parameters and device degradation. Phys. Chem. Chem. Phys. 2018, 20, 605–614. [Google Scholar] [CrossRef]
  171. Liu, F.; Dong, Q.; Wong, M.K.; Djurišić, A.B.; Ng, A.; Ren, Z.; Shen, Q.; Surya, C.; Chan, W.K.; Wang, J.; et al. Is Excess PbI2 Beneficial for Perovskite Solar Cell Performance? Adv. Energy Mater. 2016, 6, 1502206. [Google Scholar] [CrossRef]
  172. Mangrulkar, M.; Luchkin, S.Y.; Akbulatov, A.F.; Zhidkov, I.; Kurmaev, E.Z.; Troshin, P.A.; Stevenson, K.J. Rationalizing the effect of overstoichiometric PbI2 on the stability of perovskite solar cells in the context of precursor solution formulation. Synth. Met. 2021, 278, 116823. [Google Scholar] [CrossRef]
  173. Zhang, Z.; Yue, X.; Wei, D.; Li, M.; Fu, P.; Xie, B.; Song, D.; Li, Y. DMSO-based PbI2 precursor with PbCl2 additive for highly efficient perovskite solar cells fabricated at low temperature. RSC Adv. 2015, 5, 104606–104611. [Google Scholar] [CrossRef]
  174. Jiang, F.; Rong, Y.; Liu, H.; Liu, T.; Mao, L.; Meng, W.; Qin, F.; Jiang, Y.; Luo, B.; Xiong, S.; et al. Synergistic Effect of PbI2 Passivation and Chlorine Inclusion Yielding High Open-Circuit Voltage Exceeding 1.15 V in Both Mesoscopic and Inverted Planar CH3NH3PbI3(Cl)-Based Perovskite Solar Cells. Adv. Funct. Mater. 2016, 26, 8119–8127. [Google Scholar] [CrossRef]
  175. Ngo, T.T.; Masi, S.; Méndez, P.F.; Kazes, M.; Oron, D.; Seró, I.M. PbS quantum dots as additives in methylammonium halide perovskite solar cells: The effect of quantum dot capping. Nanoscale Adv. 2019, 1, 4109–4118. [Google Scholar] [CrossRef] [Green Version]
  176. Wang, S.; Dong, W.; Fang, X.; Zhang, Q.; Zhou, S.; Deng, Z.; Tao, R.; Shao, J.; Xia, R.; Song, C.; et al. Credible evidence for the passivation effect of remnant PbI2 in CH3NH3PbI3 films in improving the performance of perovskite solar cells. Nanoscale 2016, 8, 6600–6608. [Google Scholar] [CrossRef] [PubMed]
  177. Pathipati, S.R.; Shah, M.N. Performance enhancement of perovskite solar cells using NH4I additive in a solution processing method. Sol. Energy 2018, 162, 8–13. [Google Scholar] [CrossRef]
  178. Zuo, C.; Ding, L. An 80.11% FF record achieved for perovskite solar cells by using the NH4Cl additive. Nanoscale 2014, 6, 9935–9938. [Google Scholar] [CrossRef]
  179. Jahandar, M.; Khan, N.; Jahankhan, M.; Song, C.E.; Lee, H.K.; Lee, S.K.; Shin, W.S.; Lee, J.-C.; Im, S.H.; Moon, S.-J. High-performance CH3NH3PbI3 inverted planar perovskite solar cells via ammonium halide additives. J. Ind. Eng. Chem. 2019, 80, 265–272. [Google Scholar] [CrossRef]
  180. Zhang, W.; Pathak, S.; Sakai, N.; Stergiopoulos, T.; Nayak, P.K.; Noel, N.K.; Haghighirad, A.A.; Burlakov, V.M.; Dequilettes, D.W.; Sadhanala, A.; et al. Enhanced optoelectronic quality of perovskite thin films with hypophosphorous acid for planar heterojunction solar cells. Nat. Commun. 2015, 6, 10030. [Google Scholar] [CrossRef] [Green Version]
  181. Xiao, Z.; Wang, D.; Dong, Q.; Wang, Q.; Wei, W.; Dai, J.; Zeng, X.; Huang, J. Unraveling the hidden function of a stabilizer in a precursor in improving hybrid perovskite film morphology for high efficiency solar cells. Energy Environ. Sci. 2016, 9, 867–872. [Google Scholar] [CrossRef] [Green Version]
  182. Xu, W.; Lei, G.; Tao, C.; Zhang, J.; Liu, X.; Xu, X.; Lai, W.-Y.; Gao, F.; Huang, W. Precisely Controlling the Grain Sizes with an Ammonium Hypophosphite Additive for High-Performance Perovskite Solar Cells. Adv. Funct. Mater. 2018, 28, 1802320. [Google Scholar] [CrossRef]
  183. Mangrulkar, M.; Boldyreva, A.G.; Lipovskikh, S.A.; Troshin, P.A.; Stevenson, K.J. Influence of hydrazinium iodide on the intrinsic photostability of MAPbI3 thin films and solar cells. J. Mater. Res. 2021, 36, 1846–1854. [Google Scholar] [CrossRef]
  184. Zhang, X.; Yuan, S.; Lu, H.; Zhang, H.; Wang, P.; Cui, X.; Zhang, Y.; Liu, Q.; Wang, J.; Zhan, Y.; et al. Hydrazinium Salt as Additive To Improve Film Morphology and Carrier Lifetime for High-Efficiency Planar-Heterojunction Perovskite Solar Cells via One-Step Method. ACS Appl. Mater. Interfaces 2017, 9, 36810–36816. [Google Scholar] [CrossRef] [PubMed]
  185. Abzieher, T.; Mathies, F.; Hetterich, M.; Welle, A.; Gerthsen, D.; Lemmer, U.; Paetzold, U.W.; Powalla, M. Additive-Assisted Crystallization Dynamics in Two-Step Fabrication of Perovskite Solar Cells. Phys. Status Solidi Appl. Mater. Sci. 2017, 214, 1–9. [Google Scholar] [CrossRef]
Figure 1. Chemical structures of amine-based additives for MAPbI3 PSC.
Figure 1. Chemical structures of amine-based additives for MAPbI3 PSC.
Crystals 11 00814 g001
Figure 2. Chemical structures of nitrile additives in MAPbI3 PSC.
Figure 2. Chemical structures of nitrile additives in MAPbI3 PSC.
Crystals 11 00814 g002
Figure 3. Stability of MAPbI3 photoactive layer due to various additives containing N donor atom. The data are taken from references [33,34,35,39,40,41,47,48,49,50,53,54,55,56,57,59,60,61,62] respectively.
Figure 3. Stability of MAPbI3 photoactive layer due to various additives containing N donor atom. The data are taken from references [33,34,35,39,40,41,47,48,49,50,53,54,55,56,57,59,60,61,62] respectively.
Crystals 11 00814 g003
Figure 4. Chemical structures of amide/carbonyl/sulfonyl based additives for MAPbI3 PSC.
Figure 4. Chemical structures of amide/carbonyl/sulfonyl based additives for MAPbI3 PSC.
Crystals 11 00814 g004
Figure 5. Influence of carbonyl/sulfonyl group onto stability of MAPbI3-based PSCs. The data are collected from references [65,67,68,71,73,74,75,76,77,79,84,85] respectively.
Figure 5. Influence of carbonyl/sulfonyl group onto stability of MAPbI3-based PSCs. The data are collected from references [65,67,68,71,73,74,75,76,77,79,84,85] respectively.
Crystals 11 00814 g005
Figure 6. Chemical structures of acid containing additives for MAPbI3 PSC.
Figure 6. Chemical structures of acid containing additives for MAPbI3 PSC.
Crystals 11 00814 g006
Figure 7. Impact of acidic additives on the stability of MAPbI3-based solar cells. The graph is plotted using the data from references [88,89,90,91,92,93,94,95] respectively.
Figure 7. Impact of acidic additives on the stability of MAPbI3-based solar cells. The graph is plotted using the data from references [88,89,90,91,92,93,94,95] respectively.
Crystals 11 00814 g007
Figure 8. Chemical structures of acetate additives for MAPbI3 PSC.
Figure 8. Chemical structures of acetate additives for MAPbI3 PSC.
Crystals 11 00814 g008
Figure 9. Influence of acetate-based organic additives onto stability of MAPbI3 perovskite. The graph is plotted with the data taken from references [98,100,102,103,104] respectively.
Figure 9. Influence of acetate-based organic additives onto stability of MAPbI3 perovskite. The graph is plotted with the data taken from references [98,100,102,103,104] respectively.
Crystals 11 00814 g009
Figure 10. Chemical structures of ester and ether containing additives for MAPbI3 PSC.
Figure 10. Chemical structures of ester and ether containing additives for MAPbI3 PSC.
Crystals 11 00814 g010
Figure 11. Influence of ester and ether-based additives on the stability of MAPbI3 perovskite. The graph is plotted with data from [106,107,109,110,114], respectively.
Figure 11. Influence of ester and ether-based additives on the stability of MAPbI3 perovskite. The graph is plotted with data from [106,107,109,110,114], respectively.
Crystals 11 00814 g011
Figure 12. Chemical structures of alcohol additives for MAPbI3 PSC.
Figure 12. Chemical structures of alcohol additives for MAPbI3 PSC.
Crystals 11 00814 g012
Figure 13. Influence of alcohols onto stability of MAPbI3 PSK. The data to plot the graph are taken from references [115,116,117].
Figure 13. Influence of alcohols onto stability of MAPbI3 PSK. The data to plot the graph are taken from references [115,116,117].
Crystals 11 00814 g013
Figure 14. Chemical structures of oxygen-based multifunctional group containing additives.
Figure 14. Chemical structures of oxygen-based multifunctional group containing additives.
Crystals 11 00814 g014
Figure 15. Other oxygen-based additives with multiple functional groups [120,121].
Figure 15. Other oxygen-based additives with multiple functional groups [120,121].
Crystals 11 00814 g015
Figure 16. Chemical structures of organosulfur additives for MAPbI3 PSC.
Figure 16. Chemical structures of organosulfur additives for MAPbI3 PSC.
Crystals 11 00814 g016
Figure 17. Influence of S donor atom onto stability of MAPbI3-based PSC [122,124,125,126,131,132].
Figure 17. Influence of S donor atom onto stability of MAPbI3-based PSC [122,124,125,126,131,132].
Crystals 11 00814 g017
Figure 18. Chemical structures of alkane-based additives for MAPbI3 PSC.
Figure 18. Chemical structures of alkane-based additives for MAPbI3 PSC.
Crystals 11 00814 g018
Figure 19. Schematic representation of quantum dots with chelating functional group/groups.
Figure 19. Schematic representation of quantum dots with chelating functional group/groups.
Crystals 11 00814 g019
Figure 20. Quantum dots-based additives and their influence on stability of MAPbI3. Data taken from [139,141,142,145] respectively.
Figure 20. Quantum dots-based additives and their influence on stability of MAPbI3. Data taken from [139,141,142,145] respectively.
Crystals 11 00814 g020
Figure 21. Chemical structures of alkali metal additives for MAPbI3 PSC.
Figure 21. Chemical structures of alkali metal additives for MAPbI3 PSC.
Crystals 11 00814 g021
Figure 22. Influence of alkali metal additives onto stability of MAPbI3. Data taken from references [151,155].
Figure 22. Influence of alkali metal additives onto stability of MAPbI3. Data taken from references [151,155].
Crystals 11 00814 g022
Figure 23. Chemical structures of transition metal additives.
Figure 23. Chemical structures of transition metal additives.
Crystals 11 00814 g023
Figure 24. Influence of transition metal additive onto MAPbI3. Data taken from references [157,158,162] respectively.
Figure 24. Influence of transition metal additive onto MAPbI3. Data taken from references [157,158,162] respectively.
Crystals 11 00814 g024
Figure 25. Chemical structures of other metal additives in MAPbI3.
Figure 25. Chemical structures of other metal additives in MAPbI3.
Crystals 11 00814 g025
Figure 26. Ambient and inert stability due to excess PbI2 [171,172].
Figure 26. Ambient and inert stability due to excess PbI2 [171,172].
Crystals 11 00814 g026
Figure 27. Chemical structure of non-metal halide additives for MAPbI3.
Figure 27. Chemical structure of non-metal halide additives for MAPbI3.
Crystals 11 00814 g027
Figure 28. Influence of non-metal additives onto stability of the MAPbI3 layer. Data were taken from references [177,179,183], respectively.
Figure 28. Influence of non-metal additives onto stability of the MAPbI3 layer. Data were taken from references [177,179,183], respectively.
Crystals 11 00814 g028
Table 1. Investigated additives based on nitrogen donor atom containing amines or nitrile.
Table 1. Investigated additives based on nitrogen donor atom containing amines or nitrile.
Additive in Active LayerArchitecturePCE, %
(Pristine)
Stability
(Pristine)
PCE, %
(with Additive)
Amount of AdditiveStability with AdditiveStability ConditionsRole of AdditiveRef
excess MAIFTO/SnO2/CH3NH3PbI3/PCBM/Ag15.14N/A17.24molar ratio PbI2: MAI = 1:1.05N/AN/AHot casting, improved crystallinity, decrease in defect density, increased PL lifetime[29]
excess MAIITO/PCBM/CH3NH3PbI3/HTL/AuN/AN/AN/Amolar ratio
MAI: PbI2 = 3:1
N/AN/AVacuum deposition, increased PL lifetime, reduced trap states[30]
excess MAIFTO/TiO2/CH3NH3PbI3/Spiro-OmATAD/Au11.13N/A13.370.2 mMN/AN/ASequential deposition, good quality perovskite film [31]
benzylammonium
iodide
(BAI)
FTO/TiO2/CH3NH3PbI3/HTL/metal6.83N/A9.05molar ratio BAI:MAI = 0.2N/AN/ABetter light harvesting property and low charge recombination[32]
phenethylammonium iodide (PEAI)FTO/
c-TiO2/m-TiO2/mp- ZrO2/CH3NH3PbI3/Carbon
6.3Retained ~77% PCE after 80 days8.60molar ratio PEAI: MAI = 1:20Retained 90% PCE after 80 daysair, light 100 mW cm−2 Better contact with TiO2, longer exciton lifetime, good quality of perovskite film[33]
hexylamine hydrochloride
(1-HH)
Glass/ITO/PEDOT:PSS/CH3NH3PbI3/PCBM/BCP/Ag14.37Retained 43% PCE from initial after 16 days15.700.05 wt%Retained ~85% PCE from initial after 16 daysUnencapsulated, ambient, RT, air, RH = 10–20%Increases grain size and passivate defects, NH3+ group could form N-H…I. hydrogen
bond with
the I- of the [PbI6]4- passivating
“A” vacancy; hydrophobic hexane alkyl chain protects against moisture
[34]
1,6-diaminohexane dihydrochloride (1,6-DD)17.00Retained 90% PCE from initial after 16 days
phenylhydrazinium iodide (PHAI)FTO/PEDOT:PSS/CH3NH3PbI3/PCBM/Rhoda mine/Ag14.63(a) Retained ~53% PCE after 60 days
(b) died in 20 days
17.210 mg mL−1(a) Retained ~90% PCE after 60 days
(b) Retained ~85% PCE after 20 days
(a) unencapsulated, N2, dark, RH = 20%, T = 26 °C
(b) unencapsulated, ambient room environment, dark, 30 ± 5% RH, 24 ± 2 °C
PbI2 and PHAI complex/intermediate formation results in passivation against vacancy defects, hydrophobic phenyl rings acts as a barrier against moisture[35]
amphiphilic hexadecyl trimethyl ammonium bromide(CTMAB)Glass/FTO/TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au17.05Retained 70% PCE after 40 days18.0310 mg CTMAB in 1 mL DMSO. Used 20 uL of this to add in 1 mL precursorRetained 95% PCE after 40 daysNon-encapsulated, RH~40%, T = 25 °C, darkImproved crystallinity and morphology[39]
ethylammonium chloride (EACl)Glass/FTO/c-TiO2/meso-TiO2/CH3NH3PbI3/HTL/Au 17.35Retained ~30% of the original PCE after 1000 h20.32.5%, molar %Retained ~89% of the original PCE after 1000 hencapsulatedImproves morphology, grain boundary passivation[40]
1,3-diaminoguanidine monohydrochloride (DAGCl)Glass/
ITO/PolyTPD/CH3NH3PbI3/PC61BM/ZrAcac/Ag
19.1Retained 70% PCE after 20 days20.30.6%, wt% of MAIRetained 80% PCE after 20 daysNon-encapsulated, ambient, RH ~ 50% Increased grain size, reduced trap density, DAG cation can bond with I via hydrogen[41]
benzamidine hydrochloride (BMCl)ITO/TiO2/PC61BM/CH3NH3PbI3/PTAA/MoO3/Ag17.8N/A18.4N/AN/AN/AEnhance work function of perovskite film improves efficiency[42]
acetamidine salt (AcHc)ITO/TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au15.45N/A16.54molar ratio MAI:AaHc = 1:0.08N/AN/ASmooth film, full and uniform coverage, large grain size, improve carrier lifetime,[43]
methoxyammonium chloride
(MeOCl)
FTO/compact-TiO2
layer/mesoporous-TiO2 layer/CH3NH3PbI3/Spiro-MeOTAD/Ag
17.15N/A19.71molar ratio PbI2:MeOCl =1:0.10 N/AN/AImprovement in grain size and crystallinity[44]
2, 2, 2-
trifluoroethylamine hydrochloride (TFEACl)
Glass/ITO/PEDOT:PSS/CH3NH3PbI3/PCBM/AlN/AN/A4.98molar ratio of additive:MAI:PbI2 = 0.4:1:1N/AN/ACompact smooth high quality film[45]
benzenamine hydrochloride
(BACl)
11.07
3-chloropropylamine
hydrochloride
(3-CPACl)
8.21
diethylamine hydrochloride (DEACl)9.89
4-ethylamine Phenylphosphate disodium salt (EAPP)FTO/C-TiO2/mp-TiO2/CH3NH3PbI3/Spiro-OMeTAD/Ag18.83(a) Dead in 6 h
(b) Remained 60% from initial PCE after 100 h
17.611 mol%(a) Remained 90% from initial PCE after 12 h
(a) Remained ~99% from initial PCE after 100 h
(a) ATM condition, RH = 80%, unencapsulated, (b) inert, N2, unencapsulated, Phenylethylamine group protects against moisture and improves ambient stability, phosphate
sodium prevents the formation of
(CH3NH3)4PbI6.H2O
[62]
4,4’-bipyridineGlass/CH3NH3PbI3N/ADied in ~900 hN/A5 wt%Remained active for 1400 hinert, unencapsulated, light 70-80 mW cm−2, 50-60 °CForms complex with PbI2, thus slows down the formation of metallic lead[47]
poly 4-vinylpyridine (PVP)FTO/TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au6.09Remained 1.55% PCE as of final PCE (absolute value) after 3 weeks13.070.4 wt%Remained 6.6% PCE as of final PCE (absolute value) after 3 weeks.Air, 50% RH, non-encapsulatedInhibits carrier recombination, reduced defects[61]
Polyvinylcarbazole (PVC)ITO/PTAA/CH3NH3PbI3/PCBM/Al17.4Died in 1500 h18.71 mass%Retained ~70% of the initial efficiency after light soaking for 1500 h,light 50 ± 3 mW cm−2, 65 ± 2 °C, inert nitrogen, Defect passivation due to interaction with lone pair of electrons from N atom with Pb2+[48]
pyridine-functionalized
fullerene derivative (C60-PyP)
ITO/TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au17.61Retained 70% PCE after 30 days19.820.13 wt%Retained 90% PCE after 30 days25 ℃, RH = 30%, dark, non-encapsulation, Enlarged grain size, improved crystallization, interaction b/w N atom of the pyridine moiety within C60-PyP and Pb2+ ion within MAPbl3 leads to the
passivation of trap states of perovskite layer, hydrophobic nature of C60-PyP molecule increases ambient stability
[49]
pyridine-2-carboxylic lead salt (PbPyA2)ITO/P3CT-N/CH3NH3PbI3/
(PCBM)/C60/(BCP)/Ag
18.86(a) Retained 20% of
PCE after 480 h
(b) retained 20% of PCE
after 540 h
(c) retained 30% of PCE from initial after 480h
19.964 mg mL−1(a) Retained 80%
PCE after 480 h (b) retained 93% of initial PCE
after 540 h
(c) retained 90% PCE from initial after 480h
(a) 90 °C, RH = 40–60%, dark, not encapsulated, (b)MPP- tracking, non-encapsulated, white light-100 mW cm−2, inert, 25 °C, (c) Air, non-encapsulated, dark,
RH = 40–60%
Controlled crystallization, passivation of grain boundaries, the interaction of pyridine and carboxylate to cations increases hydrophobicity[50]
polyvinylpyrrolidone (PVP)Ito/SnO2/CH3NH3PbI3/Spiro-OMeTAD/Au15.33Retained 76% from initial PCE after 400 h15.191 mg mL−1Retained 80% of the initial PCE after 60 days encapsulated, ambient, RH = 10%, RTLewis base, the pyridine part (side chain) of the PVP polymer, can passivate the surface defects caused by misaligned lead ions and can fill the iodine vacancy traps on the surface of the perovskite film, C=O also stabilizes, allows not to degrade[36]
NMPFTO/ZnO-MgO-EA+
/mesoporous-TiO2/CH3NH3PbI3/spiro-OMeTAD/Au
18.0N/A19.2molar ratio PbI2:NMP =1:2N/AN/AThe same kind of intermediate, regardless of NMP ratio, results in excellent morphology[51]
NMPGlass/ITO/PEDOT:PSS/CH3NH3PbI3/PC61BM/Al1.50N/A7.0360 μL in 1 mL precursorN/AN/ALewis acid base interaction[52]
NEP 10.04 -
CHP 12.87 Suppression of solvate formation
NOP 2.79 Lewis acid–base interaction
DMIFTO/compact-TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au10.72N/A14.5410 vol% N/AN/APb-O bond formation due to Lewis adduct between DMI and Pb[63]
1-(4-ethenylbenzyl)-
3-(3,3,4,4,5,5,6,6,7,7,8,8,8-tridecafluorooctylimidazolium iodide (ETI)
FTO/C-TiO2/mp-TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au19.2(a) Retained 49% PCE from initial 700 h
(b) Retained 60% PCE from initial 700 h
19.51 mol%(a)Retained 85% PCE from initial 700 h
(b) Retained 80% PCE from initial 700 h
(a) MPPT inert, 60 °C, light 100 mW cm−2, unencapsulated,
(b) RH = 40%,RT, dark, air
Enables the full transformation into precursor,
suppresses thermal decomposition pathway and, provides outstanding hydrophobicity within the
active material
[53]
1-methyl-3- propylimidazolium bromide (MPIB)ITO/PEDOT:PSS/CH3NH3PbI3/PC61BM/BCP/Ag.15.9retain ~50% of its original PCE after 150 h18.20.5 mg
mL−1
retain 78% of its original PCE after 150 hatmospheric environment, RT,Passivation of the uncoordinated Pb2+ to reduce the defects in the perovskite film due
to the lone-pair electron in its cation group, and beneficial to promote crystal growth to improve film quality
[54]
1-butyl-3- tetrafluoroborate
(BMIMBF4)
ITO/NiOx/CH3NH3PbI3/C60/Ag18.13retaining 30% of their initial PCE after thermal ageing of 400 h at 85 °C18.070.4 mol%retaining 80% of their initial PCE after thermal ageing of 400 h at 85 °C85 °C,
unencapsulated
Thermal stability by effective suppression of
perovskite decomposition
[55]
Spiro-OMeTADFTO/Den TiO2/mp-TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au15.52N/A17.770.01 wt%N/AN/AFacilitates charge transport[64]
triazine-graphdiyne (Tra-GD),ITO/P3CT-K/CH3NH3PbI3/PC61BM/ZnO/Al17.90Died in 1100 h20.332 mg mL−1Retain above 90% after 1100 h UnencapsulatedInteracts with Pb2+ at grain boundaries and passivates grain boundaries and inhibits ion migration[56]
graphitic carbon nitride(g-C3N4)FTO/compact TiO2/CH3NH3PbI3/Spiro-MeOTAD/Au18Retained 46% PCE after 300 h21.60.1 wt%Retained 90% PCE from initial after 500 hEncapsulated, light 100 mW cm−2 Improves grain size and crystallinity, passivates grain boundaries, C=N to interact with Pb2+ that forms compact tight bonding resulting in good morphology[57]
graphitic carbon nitride (g-C3N4)FTO/
compact TiO2/CH3NH3PbI3/Spiro-OMeTAD/MoO3/Ag
16.22 ± 0.83N/A19.34 ± 0.630.4 mg mL−1N/AN/APassivation, enhanced crystallinity, C=N to interact with Pb2+, increases the conductivity
and carrier mobility
[58]
1,8- Diazabicyclo[5.4.0]undec-7-ene (DBU)Glass/ITO/NiOx/CH3NH3PbI3/PCBM/PEI/Ag15.98Retained 50% PCE from initial after 10 days18.133% weight ratioRetained 80% PCE from initial after 10 daysunencapsulated, inert (N2,) mpp trackingIodine quencher, adduct with PbI2 (C=N interaction with Pb2+), reduced defects, high-quality perovskite film[59]
ACNITO/TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au15.04 ± 0.48Remained ~20% of initial PCE after 150 h19.7molar ratio ACN: PbI2 = 0.5Remained ~60% PCE after 150 hlight 100 mW cm−2 RH = 60%–90%, in air without encapsulationMorphology enhancement[60]
Table 2. A summary of additives containing amides/carbonyls/sulfonyl functional group in MAPbI3 active layer.
Table 2. A summary of additives containing amides/carbonyls/sulfonyl functional group in MAPbI3 active layer.
Additive in Active LayerArchitecturePCE, %
(Pristine)
Stability
(Pristine)
PCE, %
(with Additive)
Amount of AdditiveStability with AdditiveStability ConditionsRole of AdditiveRef
ureaGlass/ITO/PEDOT:PSS/CH3NH3PbI3/PCMB/BCP/Al15.1Retained ~35% from original PCE after 2 h17molar ratio urea: lead acetate = 0.5Retained ~60% from original PCE after 2 hlight, 100 mW cm−2, 50–60 °C, inert, unencapsulatedInteraction of carbonyl group of urea with PbI2 results in crystalline film, resulting passivation and improvement in stability[65]
ureaITO/PEDOT:PSS/PbI2/MAI/PCBM/BPhen/Ag12.75N/A18.015 wt%, wt% of PbI2N/AN/ADouble step spin coating, urea and PbI2 together form PbI2.O=C(NH2)2 complex/intermediate, resulting big flat grains[66]
ureaFTO/SnO2/CH3NH3PbI3/Spiro-OMeTAD/Au15.48Remained 50% from original PCE after 60 days18.5220 mol%Remained 80% from original PCE after 60 daysair, 50% humidity, without any encapsulationPbI2-urea adduct results increment in grain size and passivate grain boundaries[67]
ureaITO/SnO2/CH3NH3PbI3/Spiro-OMeTAD/Ag16.80a) Retained 10% PCE as of final PCE after ~27 days
b) retained nearly 90% PCE from initial after 60 min
18.54 mol%(a) Remained 12% PCE as of final PCE after ~27 days
(b) retained nearly 100% PCE from initial after 60 min
mppt tracking
(a) dark, ambient,
(b) 1 sun, in-situ
Adduct formation,
high crystallinity, good film, suppress non-radiative recombination and passivates grain
boundaries
[68]
formamideFTO/c-TiO2/
mp-TiO2//CH3NH3PbI3/ZrO2 layer mp-ZrO2/
carbon electrode
14.26N/A15.21 N/AN/Athe amide additives shifted the Fermi
level of the MAPbI3 perovskite from −4.36 eV to −4.63,
−4.65, and −4.61 eV, respectively for formamide, acetamide and urea and suppressed non-radiative
recombination. Reduced iodide defects vacancy
[70]
acetamide15.57 N/AN/A
urea15.07 N/AN/A
biuretFTO/TiO2/CH3NH3PbI3/PCBM/Ag18.26Maintains ~50%PCE after 12 days21.162 mol%Preserves 94% of its initial efficiency after 12 day85 °C, N2, inert, non-encapsulated, Increased grain size, reduced trap state, additive acts as Lewis base and interacted with uncoordinated Pb2+,
improves the thermal stability
[71]
ureaGlass/ITO/NiOx/CH3NH3PbI3//PC61BM/Ag15N/A18.60.05 mmol N/AN/APassivate the defects and eliminate non-radiative charge recombination due to adduct with PbI2[72]
biuret 20.1N/AN/A
triuret 19.2N/AN/A
benzoquinone (BQ)ITO/PEDOT:PSS/CH3NH3PbI3/
C60/BCP/Au
10.7Remained ~50% PCE from initial after 900 h15.60.5 mol% Remained ~80% PCE from initial after 1000 hunencapsulated, light 1 sun, open circuit, without UV-filterReduces trap density[73]
benzoquinone (BQ)FTO/TiO2/CH3NH3PbI3/SpiroOMeTAD/Au17.37(a) Dead in 1200 h
(b)Dead in ~800 h
18.010.25%(a) Retained 75% PCE from initial after 2400 h
(b) Retained ~30% PCE from initial after 1200 h
unencapsulated, without UV-filter
(a) RH ~ 20%, RT,
(b) ambient air, RH = 40 - 70%, RT
Improves crystal quality, passivate grain boundary, reduce trap states, suppress PbI2 formation at GB[74]
triazine perylene diimide (TPDI)ITO/PEDOT:PSS/CH3NH3PbI3
/PC61BM/Ag
8.32Retained 30% from initial PCE after 400 h10.841.2 mg mL−1Retained 60% from initial PCE after 400hNon-encapsulated, ambientMinimize grain boundary defects and enhance the coverage and crystal grain sizes[75]
3,4-Dihydroxybenzhydrazide (DOBD)ITO/PEDOT:PSS/CH3NH3PbI3/C6 0/BCP/Al14.47Retained 70%PCE from initial after 35 days17.583 mg mL−1Retained 85%PCE from initial after 35 daysNon-encapsulated, inert, darkIncreases grain size
and decrease of grain boundaries, both of which facilitate charge
transportation and suppress charge recombination, due to Lewis acid–base interaction between Pb2+ and C=O
[76]
Isatin-ClITO/PTAA:F4TCNQ/CH3NH3PbI3/PCBM/Al18.13(a) Retained 60% PCE from initial after 350 h.
(b)Retained 30% PCE from initial after 24 h
20.180.0001 wt% (a) Retained 95% PCE from initial after 350 h.
(b) Retained 75% PCE from initial after 24 h
unencapsulated,
(a) in ambient air, RH = 45%, room temperature,
(b) in nitrogen atmosphere at 85 °C
The carbonyl groups and hydrogen-bond structures
on Isatin-Cl passivate the defect states in the perovskite
grain boundaries and improved charge transport, suppress charge recombination, hydrophobic ring attached to Isatin-Cl improves stability against humidity
[77]
polycaprolactone (PCL)glass or PET/ITO/PEDOT:PSS/CH3NH3PbI3/PC61BM/BCP/Ag10.52Retained 32% of initial PCE (10.12%) after 300 bending cycles14.490.025 wt%Retained 90% of initial PCE (10.12%) after 300 bending cyclemechanical bending stability Carbonyl (C=O) and Pb2+ bond helps the uniform coverage of perovskite, which avoids the defects and achieve the grain boundary regulation on flexible PSC[78]
IDIS-ThITO/P3CT-N/CH3NH3PbI3/PC61BM/Bphen/Ag17.78(a)Retained 50% from initial PCE after 300h
(b) Retained 78% from initial PCE after 200 h
20.10.05 mg mL−1(a) Retained 80% from initial PCE after 300 h
(b) Retained 85% from initial PCE after 200 h
(a) light 100 mW cm−2, ambient, RH = 30%, non-encapsulated, (b) 85 °C, inert, dark, Passivate grain boundary, reduce charge recombination, adduct between additive and under coordinated
Pb ions can effectively inhibit ion migration and
moisture diffusion to enhance the stability of PSC devices
[79]
IDIC-Th18.78(a) Retained 80% from initial PCE after 300 h
(b) Retained 78% from initial PCE after 200 h
NMPFTO/TiO2/PbI2/MAI/Spiro-OMeTAD/Au12.25Died after 14 days14.6630% NMP in 1 mL precursorN/A [84]
DMSO16.1720% DMSO in 1 mL precursorRetained 66% PCE from initial after 14 daysin air Morphology control based on Lewis basicity-based, donor number and boiling point
HMPA12.065% HMPAN/A
tetramethylene sulfone (TMS)FTO/bl-TiO2/mp-TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au8.7Retained ~70% from initial PCE after30 days16.2molar ratio =PbI2:MAI:TMS= 2:2:1Retained 80% from initial PCE after30 daysRT, RH = 10 - 40%, un encapsulated, darkIntermediate phase via cross-linking. PbI2 act as Lewis acid TMS with S=O group acts as Lewis base[85]
tetrahydrothiophene oxide (THTO),ITO/PEDOT:PSS/CH3NH3PbI3/PCBM/AlN/AN/A12.1molar ratio THTO:Pb = 3:1N/AN/AAdditive alters the nucleation and growth processes, lowered the free energy of the precursor by incorporating a sulfoxide (S=O), which
strongly interacts with MAPbI3 precursors, allowing an unprecedented degree of control over the
nucleation density and growth rate.
[86]
Table 3. A summary of acid additives for the MAPbI3 active layer.
Table 3. A summary of acid additives for the MAPbI3 active layer.
Additive in Active LayerArchitecturePCE, %
(Pristine)
Stability
(Pristine)
PCE, %
(with Additive)
Amount of AdditiveStability with AdditiveStability ConditionsRole of AdditiveRef
5-aminovaleric acid (5-AVA)ITO/C60/CH3NH3PbI3/PTAA/Au15.4N/A13.75% wtN/AN/ALarger grain size, improved crystallinity, improved mechanical robustness[87]
amino valeric acid
AVA
FTO/compact TiO2/mesoporous TiO2/mesoporous ZrO2/mesoporous carbon11.1Died in 110 h9.1molar ratio
AVA: MAI=3%
Retained 50% PCE from initial after 110 hRH = 15%, 1 Sun, non-encapsulated, AVA located at grain boundaries is
able to passivate surface defect sites, resulting in enhanced resistivity
to oxygen-induced degradation.
[88]
5-ammonium valeric acid iodide (5-AVAI)FTO/TiO2/CH3NH3PbI3/ZrO2/CarbonN/AN/A6.680.072M of AVAI5.62% PCE as of final PCE after 75 daysambient, Big area 0.25 cm2, lower charge transport resistance[89]
aminovaleric acid iodide (AVAI)TiO2/ZrO2/CH3NH3PbI3/Spiro-OMeTAD/Au~7N/A~9 AVA:PbI2= 3 mol%N/AN/AControls ion migration-[96]
aminopropanoic acid (APPA)ITO/PTAA/CH3NH3PbI3/
PC61BM/BCP/Ag,
17.51Remained 6% as of absolute value of PCE after 90 days19.23weight ratio MAI:APPA= 0.56Remained 13% as of absolute value of PCE after90 daysunencapsulated, RH = 10%Smaller grain size, smooth surface, suppress non-radiative charge recombination, resulting in enhanced JSC and VOC [95]
Polyacrylic acid (PAA)FTO/
NiOx:Zn/CH3NH3PbI3/
PC61BM/bathocuproine (BCP)/Ag
10.3Retained 60% from initial PCE after 24 days14.95 mg mL−1Retained 80% from initial PCE after 24 daysRT, air, ambient, RH ~ 30 ± 5%, non-encapsulated Using the doctor blade method, a smooth, uniform, and pin-hole free film of high electronic quality, passivate defects in large area devices (1 cm2)[90]
thioctic acid (TA)FTO/ c-TiO2/TiO2-Poly(TA)/CH3NH3PbI3/Spiro-OMeTAD/Au17.4(a) Died after 180 min
(b) Died after 400 h
c) N/A
20.420 mg mL−1(a) Retained 98% of its original PCE after 450 min
(b) 97% PCE after 2100 h
(c)retained 92% PCE from initial after 600 h
(a) under UV irradiation (35.8 mW cm−2).
(b) air, RH = 50 ± 10%,
(c) mppt, inert,1.5 AMG
Carboxylic acid moieties
binds at TiO2 surface.
Then five-membered ring-containing
the dynamic covalent disulfide is bonded to the other end of
the molecule through the thermal-initiated ring-opening, and interacts with Pb, forming high-quality perovskite film due to the Lewis acid–base reaction between S and Pb2+. Extra passivation on TiO2 helps for efficient charge extraction and stability under UV illumination and ambient conditions
[91]
terephthalic acid (TPA)FTO/compact-TiO2/CH3NH3PbI3/Carbon layer11.5 ±0.67(a)Retained 81% PCE from initial after 21 days
(b) ~6% PCE as of final PCE after 700 h
(c) retained 79% PCE from initial after 40 min
14. 29 ± 08 mg mL−1(a) Retained 94% PCE from initial after 21 days
(b) 10.7% PCE as of final PCE after 700 h
(c) retained 90% PCE from initial after 40 min
(a) unencapsulated, air, RH = 35%, 25 °C, ambient, dark,
(b) 60 °C
(c) 365 nm UV illumination 250 mW cm−2
(–COO) from TPA can strongly attract Pb2+ ions
crosslinking additive within perovskite, resulting in compact, dense improved morphology,
the rigidity of the phenyl skeleton allows moisture and thermal resistance
[92]
trimesic acid (TMA)FTO/TiO2/CH3NH3PbI3/Spiro-OMeTAD/Ag14.27(a)Retained 49% from initial PCE after 10h
(b) retained 46% from initial PCE after 20 days
17.211.0 mol L−1 PbI2
precursor solution
(a) Retained 49% from initial PCE after 10 h
(b) retained 71% from initial PCE after 20 days
(a) 100 °C, air, ambient, dark, non-encapsulated, (b)RT, air, RH ~ 30%, dark, non-encapsulated, TMA with a
benzene ring and three carboxyl groups maintained stability. The strong hydrogen bond
between the hydroxyl group and iodide to suppress the loss of iodide ion, preventing the perovskite from decomposing. Moreover, the benzene ring with rigidity and the π-π bond effect, and the hydrophobic alkyl chains further protects the perovskite from reacting with water
[93]
3,3’,5,5’-azobenzene-tetracarboxylic acid
(H4abtc)
FTO/TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au14.25(a) Remained 25% after 20 h
(b) retained 66% of its initial PCE after 30 d
17.672.0% (mass ratio with respect to PbI2(a) Remained 61% after 20 h
(b) retained 84% of its initial PCE after 30 d
(a) 100 °C without encapsulation, air, light 100 mW cm−2 (AM 1.5),
(b) RT, RH = 30%without encapsulation.
High-quality perovskite film, H4abtc
can passivate grain boundaries by reacting with the lead cation, therefore leading to good thermal stability and anti-moisture of perovskite films due to rigidity of the benzene ring and azo bond
[94]
4-methylbenzenesulfonic acid (4-
MSA)
FTO/bl-TiO2/mp-TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au.14.05N/A17.586 mg mL−1N/AN/AImproved performance, reduced hysterias, improve J-V characteristics, the sulfonic group is chemically bonded to
mp-TiO2 and phenyl backbone has p-conjugated structure. Jsc and Voc of 4-MSA doped PSC are enhanced.
[97]
Table 4. Acetate-based additives in MAPbI3 active layer.
Table 4. Acetate-based additives in MAPbI3 active layer.
Additive in Active LayerarchitecturePCE, %
(Pristine)
Stability
(Pristine)
PCE, %
(with Additive)
Amount of AdditiveStability with AdditiveStability ConditionsRole of AdditiveRef
formamidine acetate salt (FAAc)ITO/PEDOT:PSS/CH3NH3PbI3/PCBM/Al12.13(a) N/A
(b) N/A
16.595 mol%(a) Retained 90% PCE from initial after 30 days
(b) 16.10 as of final PCE after 320 s @ (0.87 V)
inert, unencapsulated, control
film morphology and crystallinity, improves optical and
electrical properties reduces
hysteresis
[98]
methylammonium acetate (MAAc) and thiosemicarbazide (TSC), combinedFTO/NiO/CH3NH3PbI3/PCBM/AgN/AN/A19.1910–15% MAAc (molar ratio)
3–5% TSC (molar ratio)
(a) Retained 90% PCE from initial after 1000 h
(b) Retained 80% PCE from initial after 500 h
(a) light AM 1.5, 25 °C and
RH < 25%, mppt
(b) dark, 85 °C, RH < 25%
large-area (aperture area of 1.025 cm2) using one step solution process, high crystalline quality of film[100]
ammonium acetate (NH4Ac)FTO/TiO2/CH3NH3PbI3/spiro-OMeTAD/Au13.82N/A17.0210 wt%N/AN/AImproved film morphology and increased surface coverage[101]
ammonium
acetate CH3COONH4 (NH4Ac)
FTO/TiO2/CH3NH3PbI3/carbon
devices
11.11Remained 60% from initial PCE after 1900 h13.9molar ratio MAI/PbI2/NH4Ac 1:1:x = 0.08%Remained 96% from initial PCE after 1900 hDark, RH = 40%Carbon electrode-based device, improved crystallinity[102]
zinc acetate
Zn(CH3COO)2 (ZnAc2)
12.30molar ratio MAI/
PbI2/ZnAc2 1:1−x:x = 7%
Remained 89% from initial PCE after 1900 h
lead acetate (PbAc2)Glass/
ITO/PTAA/CH3NH3PbI3/PCBM/BCP/Ag
17.25Retained 80% PCE from initial after 20 days19.07molar ratio PbAc2/PbI2=3%Retained 95% PCE from initial after 20 daysunencapsulated, inert PbAc2 additive aids cross-linking to form a strong hydrogen bond with MAI, leading to a more
stable perovskite intermediate phase, retards crystallization process, resulting, perovskite thin films with better morphology and larger grains
[103]
barium acetate (BaAc2)ITO/P3CTN/CH3NH3PbI3/PC61BM/C60/BCP/Ag18.99Retained only 20% from initial PCE after 400 h19.822 mg mL−1Retained 90% from initial PCE after 400 hNon encapsulated, inert, 90 °C, dark, suppression of ions migration,
high quality perovskite film,
grain boundary passivation
[104]
Table 5. Ester and ether-based additives for MAPbI3 and their role.
Table 5. Ester and ether-based additives for MAPbI3 and their role.
Additive in Active LayerarchitecturePCE, %
(Pristine)
Stability
(Pristine)
PCE, %
(with Additive)
Amount of AdditiveStability with AdditiveStability ConditionsRole of AdditiveRef
[6,6]-phenyl-C61-butyric acid methyl esterFTO/c-TiO2/m-TiO2/mZrO2/CH3NH3PbI3/m-carbon8.58N/A12.360.25 mg mL−1N/AN/AImproved morphology, intermediate formation between C=O of PCBM and PbI2[105]
PCBMGlass/ITO/CH3NH3PbI3/PCBM/BPhen/Ag13.94Retained ~9% (absolute value) as of final PCE after 10 days15.943.5 mg mL−1Retained ~12% (absolute value) as of final PCE after 10 dayswithout encapsulation, ambient RH = 25–50%,Improves crystallinity, passivates defects, suppress non radiative recombination[106]
C60 -Ta16.461.5 mg mL−1Retained ~12% (absolute value) as of final PCE after 10 days
C6016.590.1 mg mL−1Retained ~11% (absolute value) as of final PCE after 10 days
poly(propylene glycol) bis(2-aminopropyl ether) (PEA)FTO/c-TiO2/mp-TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au17.18Maintains ~55% of its original PCE after 30 days18.871 wt%Maintains 95% of its original PCE after 30 daysdark, ambient, air, RH = 30 ± 5% Grain boundary passivation, the oxygen atom from ether in PEA, acts as a crosslinking agent, reduces trap density and hysteresis[107]
jeffamineITO/NiOx/CH3NH3PbI3/PC61BM/BCP/Ag14.5Showed major cracks16.80.05 wt%Retained original morphology10 cycles of stretching at 30% strainDefect passivation through Lewis acid–base reaction of N atom and O atom with Pb 2+/or interaction of MA+ and hydrogen bond of Jeffamine decrease trap density,
enhances the ductility of the perovskite film to prevent cracking during stretching
[108]
ethyl cellulose (EC)FTO/c-TiO2/CH3NH3PbI3/Spiro-OMeTAD/
MoO3/Ag
17.11Completely died in 30 days19.270.1 mg mL−1Retained 80% PCE from initial after 30 daysnon –encapsulated, dark, ambient, air, RH = 45%Hydrogen bonding between EC and MAI passivates defects at the grain boundary, reduces hysteresis and improves stability[109]
tetraethyl orthosilicate
(TEOS)
TiO2/
CH3NH3PbI3/Spiro-OMeTAD/Ag
15.96Retained 73% of its initial PCE after 28 days18.380.3 mol%Retained 77% of its initial PCE after 28 daysambient, 25 °C, RH = 30%, non-encapsulated, Reduce trap density, improves carrier lifetime[110]
THFFTO/TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au11.3Retained 35% PCE from initial after 20 days15.1THF: DMF (1:10) v/vRetained 80% from initial PCE after 20 daysRH = 50%, unencapsulated, Complex formation due to Lewis acid–base reaction[113]
diethyl etherGlass/ITO/NiOx/CH3NH3PbI3/PCBM/BCP/Ag13.28N/A15.09the molar ratio of 4% with matrix solventN/AN/AGrain size improvement[114]
Table 6. Alcohol-based additives and their role in MAPbI3 perovskite framework.
Table 6. Alcohol-based additives and their role in MAPbI3 perovskite framework.
Additive in Active LayerArchitecturePCE, %
(Pristine)
Stability
(Pristine)
PCE, %
(with Additive)
Amount of AdditiveStability with AdditiveStability ConditionsRole of AdditiveRef
dibutylhydroxytoluene (BHT)ITO/PEDOT: PSS/CH3NH3PbI3/
PC61BM/ZnO/Al
17.1Retained 85% from initial PCE after 180 min18.10.02 MRetained 93% of initial PCE after 180 minlight 100 mW cm−2, RT, non-encapsulated,
RH < 5%
intermolecular hydrogen
bonds between the MA+ and −OH groups of the BHT
additive improves film crystallinity, reducing the sub-Eg states and carrier
traps,
[115]
methanolFTO/TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au16.53Retained 40% from initial PCE after 30 days19.515 vol%Retained 90% from initial PCE after 30 daysDark, ambient, nonencapsulated,morphology control/enhancement[116]
2-methoxyethanol MEITO/ZnO/CH3NH3PbI3/Spiro-OMeTAD/MoOx/Ag14.2N/A16.730 µL to 1 mL MAI solutionN/AN/Atwo-step interdiffusion protocol to prepare pin-hole free perovskite films, glycol ethers changes the lead iodide to perovskite
conversion dynamics and enhances PCE resulting in more compact
polycrystalline films, and it creates micrometer-sized perovskite crystals vertically-aligned
across the photoactive layer
[118]
2-ethoxyethanol
(EE),
13.2
2-propoxyethanol (PE)15.1
isopropanol (IPA)FTO/compact-TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au16.02Retained 40% PCE from initial after 40 days19.70precursor: IPA volume ratio = 4:1Retained 85% PCE from initial after 40 daysunencapsulated, air, ambient, RTmorphology enhancement/control[117]
n-butanolFTO/CT-IO2/mp-TiO2/CH3NH3PbI3/Spiro-OMeTAD/Ag13.8N/A15.55.0 v% for 1 mL
precursor
N/AN/Amorphology enhancement[119]
Table 7. Other oxygen atom-based additives with multifunctional group.
Table 7. Other oxygen atom-based additives with multifunctional group.
Additive in Active LayerArchitecturePCE, %
(Pristine)
Stability
(Pristine)
PCE, %
(with Additive)
Amount of AdditiveStability with AdditiveStability ConditionsRole of AdditiveRef
Reduced graphene oxide (rGO) FTO/TiO2/CH3NH3PbI3/Spiro-OMeTAD/Ag13.8Retained 20% of their initial PCE after 50 days16.59 μg mL−1 Retained 40% of their initial PCE after 50 daysdark, RH = 10%, unencapsulated,Defect-free perovskite films of enhanced crystallinity, larger and evenly
distributed grains, enhanced light
harvesting potential, increasing the photocurrent density and the resulting
improved PCE
[120]
Poly(amic acid) (PAA)ITO/NiOx/CH3NH3PbI3/PC61BM/BCP/Ag14.16(a) Remained 13% as of final PCE after 500 h
(b) Retained 50% from initial PCE after 70h
17.850.0497 mg mL−1(a) Remained 16.57% as of final PCE after 500 h
(b) Retained 88% from initial PCE after 70 h
Inert, dark, non-encapsulated, (b) 85 °C, RH = 45% without encapsulation, dark, in the gloveboxThe O atoms in PAA and
PI form hydrogen bonds with H atoms in CH3NH3
+ and lone pair of electrons of N atoms interact with Pb ions,
which stabilized the PVSK framework.
[121]
polyimide (PI)16.490.0990 mg mL−1N/AN/A
Table 8. S atom-based additive in MAPbI3.
Table 8. S atom-based additive in MAPbI3.
Additive in Active LayerArchitecturePCE, %
(Pristine)
Stability
(Pristine)
PCE, %
(with Additive)
Amount of AdditiveStability with AdditiveStability ConditionsRole of AdditiveRef
DSMgF2/PET/ITO/Nb2O5/CH3NH3PbI3/Spiro-OMeTAD/Au17.03(a) Retained 50% PCE from initial after 60 days
(b) N/A
18.4010 vol% in 1 mL precursor(a) Retained 86% PCE from initial after 60 days
(b) 15.24 (83% of initial) as of final PCE after 5000 cycles
(a) unencapsulated, RH = 35%, dark, RT
(b) bending stability- bending radius of 4 mm
Intermediate between Pb2+ and S atom[122]
thioureaITO/PEDOT:PSS/CH3NH3PbI3
/PCBM/Ag
13.4N/A16.20.5 vol%N/AN/AThiourea promotes grain growth, reduce trap states, passivate grain boundaries[123]
thioacetamide (TAA)ITO/SnO2/CH3NH3PbI3/Spiro-OMeTAD/Au17Retains ~75.1% of its initial performance after aging 816 h18.91.0% (molar ratio to PbI2)Retains 88.9% of its initial performance after aging 816 hRH = 25–35%, unsealed, air, Interaction of TAA with Pb2+
improved grain size
[124]
2-pyridylthioureaFTO/C-TiO2/mp-TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au15.5(a) Retained only ~10% PCE from initial after 30 days
(b) Retained~55% PCE from initial after 30 days
18.20.5 mg mL−1(a) Retained~95% PCE from initial after 30 days
(b) Retained~92% PCE from initial after 30 days
ambient,
air, dark
(a) RH = 55 ± 5%, dark, RT
(b) 65 °C, RH = 30%,dark
N-donor and S-donor coordinate with PbI2 and slow down the formation of PbI2. improved morphology, larger crystalline size, smooth compact, homogeneous film[125]
ITU for I− and thioureaITO/SnO2/CH3NH3PbI3/Spiro-OMeTAD/Au17.75Retained 30% of the initial PCE after 100 h20.30.003 mMRetained 80% of the initial PCE after 100 hAM 1.5 light, ambient atmosphere Reduction of iodine ions and reduces defect state concentration, and defect passivation of grain boundaries crystalline quality, improve charge transport[126]
lead thiocyanate Pb(SCN)2FTO/SnO2/CH3NH3PbI3/Spiro-OMeTAD/Au15.57N/A17.805% Pb(SCN)2 (molar ratio with respect to PbI2N/AN/AForms volatile products HSCN and CH3NH2,
resulting in formation of excess PbI2 that passivates GB
[127]
potassium thiocyanate
(KSCN)
FTO/c-TiO2/
m-TiO2/CH3NH3PbI3/Spiro-OMeTAD/Ag
19.38N/A19.61 mol%N/AN/AGrain size and crystallinity improvement reduced recombination density[129]
methylammonium thiocyanate (MASCN)
additive
Glass/FTO/TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au15.81Died in 150 h.18.7140 mol%Retained 16.34% as of final PCE after 1000 h ~89.7% from initialair, RH = 20–25%, without
encapsulation,
Rapid vacuum-based drying approach, high crystallinity, large carrier lifetimes[131]
guanidinium thiocyanate (GuSCN)FTO/TiO2/CH3NH3PbI3/PCBM/Ag15.57Retained 60% PCE from initial PCE after 15 days16.7010% mmolRetained 90% PCE from initial PCE after 15 daysunencapsulated, dark, RH = 30–40%, Enhance the crystallinity, enlarge the crystal size, and reduce the trap density of the
perovskite film
[132]
Table 9. Alkane-based additives and their role within MAPbI3 framework.
Table 9. Alkane-based additives and their role within MAPbI3 framework.
Additive in Active LayerArchitecturePCE, %
(Pristine)
Stability
(Pristine)
PCE, %
(with Additive)
Amount of AdditiveStability with AdditiveStability ConditionsRole of AdditiveRef
poly(vinylidenefluoride-co-hexafluoropropylene) (PVDF-HFP)FTO/cTiO2/CH3NH3PbI3/Spiro-MeOTAD/Au7.2N/A10.6N/AN/AN/AInteract with uncoordinated MAI molecule through hydrogen bonds between fluorine atoms. This can improve the carrier lifetimes and reduce the charge transfer resistance, which
contributes to enhancing PCE
[135]
polyvinylidene fluoride-trifluoroethylene polymer P(VDF-TrFE)FTO/c-TiO2/mesoporous-TiO2/CH3NH3PbI3/Spiro-MeOTAD/Ag9.57 ± 0.25%N/A12.54 ± 0.40N/AN/AN/ATwo-step deposition, improved crystallinity and morphology, reduced carrier recombination of charge carrier, increased carrier lifetime[136]
diiodomethane
CH2I2
Glass/FTO/TiO2/MAPbI3/Spiro-OMeTAD/Au10.0N/A16.50.25 mL in 1 mL precursorN/AN/AIodine liberation, morphology enhancement[138]
Diiodooctane
DIO
Glass/ITO/PEDOT:PSS/MAPbI3/PCBM/Ag10.61N/A17.741 vol% in 1 mL precursorN/AN/AIntermediate formation and morphology enhancement[137]
Table 10. Quantum dots-based additives for MAPbI3.
Table 10. Quantum dots-based additives for MAPbI3.
Additive in Active LayerArchitecturePCE, %
(Pristine)
Stability
(Pristine)
PCE, %
(with Additive)
Amount of AdditiveStability with AdditiveStability ConditionsRole of AdditiveRef
carbon quantum dot (CQD)ITO/NiOx/CH3NH3PbI3/PCBM/Ag15.25Remained ~10% PCE from initial after 48 h18.240.15 mg mL−1Remained~73.4% PCE from initial after 48 hunencapsulated, dark, RT, RH = 80%, Reduce non-radiative recombination loss, improve the crystallinity of film; thus, passivate grain boundaries and improve stability [139]
carbon nanodots (CNDs)ITO/NiOx/CH3NH3PbI3/PC61BM/BCP/Ag14.48% ± 0.39%N/A16.47% ± 0.26%10 mg mL−1Remained ~18% as of an absolute value of PCE after 500 hdark, 25 °C, RH = 40%
air, unencapsulated,
Increase in the crystal
size, lower content of grain boundary defects, longer carrier lifetimes.
[142]
carbon quantum dots (CQDs)FTO/TiO2/CH3NH3PbI3/Spiro-OMeTAD/Ag17.59(a) Retained 74% from initial PCE after 1500 h
(b) Retained 17% from initial PCE after 216 h
18.810.05 mg mL−1(a) Retained 90% from initial PCE after 1500 h
(b) Retained 70% from initial PCE after 216 h
(a) unencapsulated, inert.
(b) unencapsulated, RH = 50–60%,
Passivate the grain boundaries and decrease the trap-state density. The bonding between CQDs and MAPbI3 leads to a stable absorption of CQDs on the MAPbI3 surface, forming a protective layer to prevent the perovskite from coming in contact with water, thereby enhancing the
stability of PSCs
[141]
nitrogen doped CQDs (N-CQDs),ITO/PEDOT:PSS/CH3NH3PbI3/PCBM/Al8.34N/A13.9363 vol%N/AN/AN-CQDs act like an intermediate and help to form dense and smooth perovskite; passivate the trap states and decrease the
non-radiative charge recombination
[140]
Nitrogen-doped carbon dots (NCDs)FTO/blTiO2/mL-TiO2/CH3NH3PbI3/Spiro-OMeTAD/Ag12.12 ± 0.28%N/A15.93 ± 0.15%0.05 mg mL−1N/AN/ACoordinate
with the iodide ions and lead cations on the surface of
perovskite, which effectively passivates
the surface traps and help reduce non-radiative
charge carrier recombination; the
nitrogen dopants with lone electron pairs in NCDs
optimize the interfacial energy level to enhance
the charge carrier extraction efficiency at photoactive
layer/TiO2 interface
[144]
potassium cation (K+) functionalized carbon nanodots (CNDs@K)ITO/PTAA:F4TCNQ/CH3NH3PbI3/PCBM/BPhen/Ag18.25N/A21.04 N/AN/ADefects passivation and crystallization control of the perovskite film;
K+ in the grain boundary, and prevents excessive cations from occupying
interstitial sites, thereby reducing microstrain of polycrystalline film, the synergistic effect of tailored crystal size, and suppressed grain boundary defects could reduce charge trap density, facilitate charge generation, and lengthen carrier lifetime
[143]
graphene quantum dots (GQDs)FTO/c-TiO2/mp-TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au16.83Remained ~15.6% as of final PCE after 30 days18.341.5 mg mL−1Remained ~17.4% as of final PCE after 30 daysencapsulation, RH = 36%, atmPromote charge transfer from the
perovskite layer to TiO2 film. Faster electron extraction
and a slower recombination rate
passivate
hanging bonds at the perovskite GBs
[145]
Table 11. Alkali metal additives and their role in MAPbI3 perovskite.
Table 11. Alkali metal additives and their role in MAPbI3 perovskite.
Additive in Active LayerArchitecturePCE, %
(Pristine)
Stability
(Pristine)
PCE, %
(with Additive)
Amount of AdditiveStability with AdditiveStability ConditionsRole of AdditiveRef
hydroiodic acid
(HI)
ITO/PTAA/CH3NH3PbI3/PCBM/Ti/Au16.1N/A18.210.004 vol% N/AN/AGrain size improvement[149]
hydrochloric acid
(HCl)
Glass/TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au9.96N/A14.49volume ratio 8%, N/AN/AGrain size improvement
and morphology enhancement
[156]
hydrobromic acid
(HBr)
13.53volume ratio 6% N/AN/A
hydroiodic acid
(HI)
15.21volume ratio 8% N/AN/A
lithium iodide
(LiI)
FTO/TiO2/CH3NH3PbI3/electrode11.3Remained ~1% PCE as of final value after 25 days17.0112mg/mLRetained ~6% PCE as of final value after 70 daysunencapsulated, ambient, air, RH = 40%, Larger grain size and higher crystallinity and reduced PbI2 residue[155]
sodium iodide (NaI)FTO/compact TiO2/mesoporous TiO2/CH3 NH3PbI3/SpiroMeOTAD/Au14.01N/A15.140.02 mol L−1 N/AN/AImproved crystallinity and grain boundary passivation[150]
lithium chloride (LiCl)Glass/ITO/PEDOT:PSS/CH3NH3PbI3/PC61BM/C60/BCP/Al11.40Died in 50 days9.980.25%, wt%N/AN/AImproved crystallinity and homogenous nucleation, and crystallization[151]
sodium chloride (NaCl)12.771.0%, wt%N/AN/A
potassium chloride (KCl)15.080.75%, wt%Remained ~85% after 50 daysunencapsulated, inert,
dark
Table 12. Transition metal additives and their role in MAPbI3 perovskite.
Table 12. Transition metal additives and their role in MAPbI3 perovskite.
Additive in Active LayerArchitecturePCE, %
(Pristine)
Stability
(Pristine)
PCE, %
(with Additive)
Amount of AdditiveStability with AdditiveStability ConditionsRole of AdditiveRef
n-type goethite (FeOOH)FTO/cTiO2/mp-TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au15.4(a) Remained only 5.7% as of final PCE after 360 h.
(b) Retained 61% from initial PCE after 60 days.
(c) Retained only 42% from initial PCE after 480 h
19.70.1 mg mL−1(a) Retained 94% after 360 h, i.e., 17.9%.
(b) Retained 97% from initial PCE after 60 days.
(c) Retained 92% from initial PCE after 480 h
(a) inert, 85 °C, dark, non-encapsulated,
(b) 10 ± 5% RH,
(c) light 100 mW cm−2, inert, RT, mpp tracking,
Improved quality of perovskite and inhibition of iodine and methylamine ion migration [157]
silver iodide
(AgI)
FTO/compact TiO2/mesoporous TiO2/
CH3NH3PbI3/SpiroMeOTAD/Au
14.01N/A14.180.02 mol L−1 N/AN/AImproved crystallinity and grain boundary passivation[150]
copper iodide
(CuI)
15.25N/AN/A
copper bromide
(CuBr)
15.61N/AN/A
nickel chloride (NiCl2)ITO/SnO2/CH3NH3PbI3/SpiroMeOTAD/Au17.25Retained ~70% PCE after 100 h20.610.03 mM Retained ~70% PCE after 100 hen-capsulated, air, light 65 mW cm−2, ∼25 °C, RH = 45%, without
ultraviolet filter
High crystallinity, GB passivation[158]
copper chloride (CuCl2)FTO/NiOX/
CH3NH3PbI3/PCBM/BCP/Ag
9.73N/A15.222.5 mol%N/AN/AImproves morphology[159]
cadmium chloride (CdCl2)FTO/SnO2
/CH3NH3PbI3
/Spiro-OMeTAD/Au
11.7N/A13.21% molar ratioN/AN/AEnhance grain size and crystallinity[160]
zinc chloride (ZnCl2)13.760.1% molar ratioN/AN/A
rhodium iodide (RhI3)ITO/SnO2/CH3NH3PbI3/Spiro OMeTAD/Ag19.09Retained 75% of efficiency after 500 h20.711 mol%Retained 92% of efficiency after 500 hunencapsulated, dry airImproved crystallinity and grain boundary passivation[162]
Table 13. Other metal additive for MAPbI3.
Table 13. Other metal additive for MAPbI3.
Additive in Active LayerArchitecturePCE, %
(Pristine)
Stability
(Pristine)
PCE, %
(with Additive)
Amount of AdditiveStability with AdditiveStability ConditionsRole of AdditiveRef
lead iodide
(PbI2)
FTO/c-TiO2/m-TiO2/CH3NH3PbI3/Spiro-OMeTAD/Au15.95N/A18.4210% excess PbI2N/AN/AImproves the crystallinity[163]
lead iodide
(PbI2)
Glass/FTO/CH3NH3PbI3/Spiro-OMeTAD/Electrode/Encapsulation(UV-epoxy)16.2Retained 80% PCE till 12 h16.510 mg mL−1Retained 80% PCE till 10
days
ambient air, 85 °C, 60% RH, light 1 sunN/A[171]
lead iodide
(PbI2)
Glass/ITO/SnO2/CH3NH3PbI3/PTAA/MoO3/Al~8.9%Remained ~ 40% PCE after 1500 h~15.6%15% excess Remained ~ 95% PCE after 1500 hlight 30 mW cm−2, 40 °C, inertExcess PbI2 forms adduct with the cosolvent (NMP), improving PCE and stability.[172]
lead iodide
(PbI2)
FTO/TiO2/CH3NH3PbI3/Spiro-OMeTAD/Ag7.23 ± 0.10N/A14.32 ± 0.28%concentration –N/A, by annealing @130 °CN/AN/AGrain boundary passivation[176]
lead sulfide
(PbS QDs)
FTO/SnO2/CH3NH3PbI3/Spiro-OMeTAD/Au15.2N/A15.712.5 mg mL−1N/AN/ACapping ligand for surface functionalization to improve crystallinity[175]
Table 14. Non-metal additives and their role in MAPbI3 active layer.
Table 14. Non-metal additives and their role in MAPbI3 active layer.
Additive in Active LayerArchitecturePCE, %
(Pristine)
Stability
(Pristine)
PCE, %
(with Additive)
Amount of AdditiveStability with AdditiveStability ConditionsRole of AdditiveRef
ammonium iodide (NH4I)Glass/ITO/PTAA/CH3NH3PbI3/PCBM/Ag15.2Retained ~70% PCE after 20 days17.45 wt%Retained ~80% PCE after 20 daysRH = 50 ± 5%, unencapsulated, ambient, airImproved morphology, increased crystalline size and grain size, lower trap density[177]
ammonium chloride (NH4Cl)ITO/PEDOT:PSS/CH3NH3PbI3/PC61BM/Al7.97N/A9.9317.5 mg mL−1N/AN/AImproved morphology, increased crystalline size and grain size, lower trap density[178]
ammonium fluoride (NH4F)ITO/PEDOT:PSS/CH3NH3PbI3
/PC61BM/LiF/Al
13.74Retained 70% PCE after 5 weeks15.1510 mol%Retained 82% PCE after 5 weeksencapsulated, airImproved morphology, increased crystalline size and grain size, lower trap density[179]
ammonium chloride (NH4Cl)16.88Retained 87% PCE after 5 weeks
ammonium bromide (NH4Br)16.85Retained 89% PCE after 5 weeks
ammonium iodide (NH4I)17.44Retained 91% PCE after 5 weeks
ammonium hypophosphite NH4H2PO2ITO/PTAA:F4-TCNQ/CH3NH3PbI3/PC61BM/TrNBr/Ag9.4 ± 1.0%N/A16.5 ± 0.7%2.5 mg mL−1N/AN/AIntermediate with Pb and improved crystallinity grain size[182]
hydrazinium iodide (N2H5I)Glass/ITO/SnO2/CH3NH3PbI3/PTAA/MoOx/Al14.98Retained ~60% PCE after 4400 h17.030.5%, wt%Retained ~80% PCE after 4400 hinert, light 60-70 mW cm−2, 50 °C.Intermediate between PbI2 and N2H5I, improved crystallinity, grain size[183]
hydrazinium chloride (N2H5Cl)ITO/PEDOT:PSS/CH3NH3PbI3/PCBM/C60/BCP/Ag7.14N/A12.66PbI2/MAI/N2H5Cl molar ratio = 1:1:0.2N/AN/AIntermediate between PbI2 and N2H5Cl, improved crystallinity, grain size[184]
hypophosphorous acid (HPA)Glass/FTO/TiO2/CH3NH3PbI3/ETL/electrode5.1N/A80.75 uL mg−1 solution in 1 mL of MAIN/AN/ALarger grains with a smoother surface,
additive acts as a reducing agent for iodine
[185]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mangrulkar, M.; Stevenson, K.J. The Progress of Additive Engineering for CH3NH3PbI3 Photo-Active Layer in the Context of Perovskite Solar Cells. Crystals 2021, 11, 814. https://doi.org/10.3390/cryst11070814

AMA Style

Mangrulkar M, Stevenson KJ. The Progress of Additive Engineering for CH3NH3PbI3 Photo-Active Layer in the Context of Perovskite Solar Cells. Crystals. 2021; 11(7):814. https://doi.org/10.3390/cryst11070814

Chicago/Turabian Style

Mangrulkar, Mayuribala, and Keith J. Stevenson. 2021. "The Progress of Additive Engineering for CH3NH3PbI3 Photo-Active Layer in the Context of Perovskite Solar Cells" Crystals 11, no. 7: 814. https://doi.org/10.3390/cryst11070814

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