Download as pdf or txt
Download as pdf or txt
You are on page 1of 52

pubs.acs.

org/CR Review

High-Efficiency Perovskite Solar Cells


Jin Young Kim, Jin-Wook Lee, Hyun Suk Jung,* Hyunjung Shin,* and Nam-Gyu Park*

Cite This: https://dx.doi.org/10.1021/acs.chemrev.0c00107 Read Online

ACCESS Metrics & More Article Recommendations

ABSTRACT: With rapid progress in a power conversion efficiency (PCE) to reach 25%,
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

metal halide perovskite-based solar cells became a game-changer in a photovoltaic


performance race. Triggered by the development of the solid-state perovskite solar cell in
2012, intense follow-up research works on structure design, materials chemistry, process
Downloaded via EAST CAROLINA UNIV on July 29, 2020 at 00:51:36 (UTC).

engineering, and device physics have contributed to the revolutionary evolution of the solid-
state perovskite solar cell to be a strong candidate for a next-generation solar energy
harvester. The high efficiency in combination with the low cost of materials and processes
are the selling points of this cell over commercial silicon or other organic and inorganic solar
cells. The characteristic features of perovskite materials may enable further advancement of
the PCE beyond those afforded by the silicon solar cells, toward the Shockley−Queisser
limit. This review summarizes the fundamentals behind the optoelectronic properties of
perovskite materials, as well as the important approaches to fabricating high-efficiency
perovskite solar cells. Furthermore, possible next-generation strategies for enhancing the
PCE over the Shockley−Queisser limit are discussed.

CONTENTS 3.5.3. Photon Recycling AM


3.5.4. Concentrator Solar Cells AN
1. Introduction A 4. Summary AN
2. Fundamentals of the Organic Lead Halide Author Information AO
Perovskite Required to Achieve High PCE C Corresponding Authors AO
2.1. Electronic and Optical Properties of Perov- Authors AO
skites C Author Contributions AO
2.2. Parameters Affecting Charge Transport, Notes AO
Collection, And Recombination E Biographies AO
2.3. Recombination in Perovskite Materials and Acknowledgments AP
Solar Cells G References AP
3. Key Challenges in Achieving High Efficiency G
3.1. Structure Design G
3.1.1. Crystal Structure G 1. INTRODUCTION
3.1.2. Defect Structure H
3.1.3. Microstructure J The solid-state perovskite solar cell (PSC) was reported in
3.2. Materials Chemistry M 2012,1,2 with a power conversion efficiency (PCE) of
3.2.1. A Site(Cation)-Modified Perovskites N approximately 10% under a simulated 1 sun illumination.
3.2.2. B Site(Cation)-Modified Perovskites Q Furthermore, the solid-state PSC exhibited long-term stability
3.2.3. X Site (Anion)-Modified Perovskites R for 500 h, as confirmed by stability tests conducted using
3.2.4. Low-Dimensional Perovskites T unencapsulated devices under ambient conditions.1 Although
3.3. Processing Engineering T the application of organic−inorganic lead halide perovskite as a
3.3.1. One-Step vs Two-Step Processes V sensitizer in dye-sensitized solar cell (DSSC) structures was
3.3.2. Adduct Intermediated Process Z first attempted in 20093 and improved efficiency, from 3.8% to
3.4. Device Physics AA
3.4.1. Selective Contact AA Received: February 12, 2020
3.4.2. Interface Engineering AD
3.4.3. Defect Passivation AG
3.5. Over Shockley−Queisser Limit AI
3.5.1. Tandem Solar Cells AI
3.5.2. Hot Carrier Device AL

© XXXX American Chemical Society https://dx.doi.org/10.1021/acs.chemrev.0c00107


A Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 1. (a) Chronological evolution of the power conversion efficiency (PCE) of perovskite solar cells (PSCs). The numbers in the plot are
explained in Table 1. (b) PSCs with mesoscopic, normal-planar, and inverted-planar device structures. The mesoscopic structure contains a
mesoporous electron-transporting material (ETM) layer, where the perovskite can either be deposited on the ETM surface or filled in the
mesopores, and a blocking layer is usually required to protect the electron−hole recombination. In the planar configuration, structures are classified
as normal or inverted based on the type of charge collected on the transparent conductive oxide (TCO).

Table 1. PCEs Depicted in Figure 1a, Perovskite Composition, and Device Structuresa
no. PCE (%) perovskite composition device structure (coating procedure) instituted (year) ref
1 3.8 MAPbI3 liquid junction Toin U (2009) 3
2 6.5 MAPbI3 liquid junction SKKU (2011) 4
3 9.7 MAPbI3 solid-state mesoscopic TiO2 SKKU (2012) 1
4 10.9 MAPbI3 solid-state mesoscopic Al2O3 Oxford U (2012) 2
5 12 MAPbI3 mesoscopic TiO2 KRICT (2013) 6
6 14.1b MAPbI3 mesoscopic TiO2 (two-step) EPFL (2013) 11c
7 16 FAPbI3 mesoscopic TiO2 SKKU (2014) 7
8 17 MAPbI3 mesoscopic TiO2 (two-step) SKKU (2014) 8
9 17.9b (FAPbI3)0.85(MAPbBr3)0.15 mesoscopic TiO2 KRICT (2014) 12c
10 19.3 MAPbI3 normal planar, Y:TiO2 ETM UCLA (2014) 9
11 20.1b (FAPbI3)0.95(MAPbBr3)0.05 mesoscopic TiO2 KRICT (2014) 13c
12 20.4 MAPbI3 mesoscopic TiO2 SKKU (2015) 10
13 22.1b (FAPbI3)0.95(MAPbBr3)0.05 mesoscopic TiO2 KRICT (2016) 14c
14 22.7b (FAPbI3)0.95(MAPbBr3)0.05 mesoscopic TiO2, fluorene-terminated HTM KRICT (2017) 15c
15 23.7b FA0.92MA0.08PbI3 normal planar, SnO2 ETM ISCAS (2018) 16c
16 24.2b KRICT/UNIST (2019) 17c
17 25.2b KRICT/MIT (2019) 18c
a
The device structures are displayed in Figure 1b. bCertified value. c“Solar Cell Efficiency Tables” published in Progress in Photovoltaics: Research
and Applications. dInstitutes: SKKU = Sungkyunkwan University; KRICT = Korea Research Institute of Science and Technology; ISCAS =
Institute of Semiconductors, Chinese Academy of Sciences.

6.5%, was reported in 2011,4 the dissolution of the perovskite hybridization. The most studied organic−inorganic lead halide
dots adsorbed on the TiO2 surface in an iodide/iodine-based perovskites for PSCs are MAPbI3 (MA = methylammonium;
liquid electrolyte remained highly problematic. This drawback CH3NH3+) and FAPbI3 (FA = formamidinium; HC(NH2)2+),
was eventually resolved by replacing the liquid electrolyte with where mixed cations in the A site and/or mixed anions in the X
a solid-state hole-conducting material: spiro-MeOTAD site are also available. Starting with a PCE of 9.7% with pure
(2,2′7,7′-tetrakis(N,N-di-p-methoxyphenyl amine)-9,9′-spiro- MAPbI3 in 2012, a certified PCE of 25.2% was achieved in
bifluorene.1 2019 using a mixed cation and/or mixed anion composition.
Perovskite is originally the mineral name of calcium titanate Since 2009, high efficiency has been realized by two main
(CaTiO3), which is often applied to the class of materials approaches: effective coating and composition engineering. In
possessing the same type of crystal structure as CaTiO3.5 In Figure 1, the chronological progress in the PCE of PSCs is
the perovskite structure with the chemical formula, ABX3, the displayed. Noncertified PCE values are collected from the
A-site cation is coordinated to 12 X anions, forming a literature,1−10 and the certified data are collected from “solar
cuboctahedron, while the 6-fold-coordinated B-site cation has cell efficiency tables” reported in Progress in Photovoltaics.11−18
an octahedral geometry. The perovskite structure is unique By examining the PCE data in Figure 1a, it can be observed
among the known structures of close-packed oxides because it that the PCE increases steeply between number 3 and number
can accommodate a very large cation. Therefore, small organic 11, indicating that this big jump occurred in only three years,
cations, which are larger than inorganic cations, can participate from 2012 to 2014. Most of the achievements are based on a
in the perovskite structure, leading to organic−inorganic mesoscopic structure containing mesoporous TiO2 layers (see
B https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

the mesoscopic-filled device structure in Figure 1b). From employed primarily to remove the polar aprotic solvents used
2015 to 2019, the PCE increased rather slowly but for preparing the precursor solution, such as dimethyl sulfoxide
continuously and eventually reached 25.2%. Since 2015, (DMSO) and γ-butyrolactone (GBL), while in the Lewis
relatively high PCEs have been realized by changing the acid−base adduct approach, the polar aprotic solvent (DMSO)
material composition from MAPbI3 to FAPbI3 (Table 1), is used as the Lewis base instead of using the solvent to form
where small amounts of the MA cation and Br anion are an adduct, and the solvent, such as dimethylformamide
introduced to FAPbI3 to stabilize the α phase of FAPbI3 at (DMF), is removed by a selective diethyl ether solvent
ambient temperature. A PCE of 24.2% was certified in 2019 (DMF is selectively dissolved without dissolving DMSO).
(number 16 in Figure 1a) resulting from open-circuit voltage Currently, these technologies are extended to fabricate high-
(Voc) = 1.1948 V, short-circuit current density (Jsc) = 24.16 quality FAPbI3 films and have contributed to achieving
mA/cm2, and fill factor (FF) = 0.84,17 where Voc and FF are relatively high PCEs, which will be discussed in detail in
significantly improved as compared to a PCE of 20.1% certified section 3.3.
in 2014 (Voc = 1.059 V and FF = 0.77).13 In addition to In this review, the key challenges associated with achieving
composition engineering, interface engineering is likely to be high PCE in PSCs are described and discussed. Because a high
involved in the improvement of Voc and FF, as explained in the PCE is directly related to the device configuration, under-
section describing the importance of interface engineering.9 In standing the device physics and device structure engineering is
addition to perovskite compositional engineering, deposition of critical importance. Nevertheless, the materials and
(coating) methods play an equally important role in improving interfaces in the device are equally important for achieving
the PCE. In 2013, the first high PCE of 14.1% was achieved by maximum Jsc, Voc, and FF. In addition, an optimal fabrication
a two-step coating method,19 which led to the deposition of process must be developed, specifically for high-quality
organic halide on the predeposited PbI2 film. While the PCE perovskite films, because the perovskite light absorber is the
was improved to 17% by controlling the size of the perovskite key element in PSCs. A brief introduction to the optoelec-
cuboid in the two-step process,8 further improvement in the tronic properties of perovskite materials is considered as a
PCE was realized by a one-step procedure and perovskite prerequisite to designing the best PCE device.
composition engineering. Although the best PCE of 25.2% was
reported, the cell area of ca. 0.1 cm2 was considered overly 2. FUNDAMENTALS OF THE ORGANIC LEAD HALIDE
small for classification as an outright record.18 A PCE of 21.6% PEROVSKITE REQUIRED TO ACHIEVE HIGH PCE
was certified with a cell area of 1 cm2, which was measured
A solar cell is a device that converts absorbed incident light
based on a “settled power at fixed voltage” (SPFV) method,
into electrical energy. The optical and electronic properties of
where the device voltage was maintained at the nominal
light absorbers are fundamental for understanding their
maximum power point voltage of 1.05 V for 5 min.18
interaction with light, as well as the transport and collection
As listed in Table 1, the PCEs as high as 22% are mostly
of photogenerated charges. In addition, the recombination
derived from FAPbI3 in the mesoscopic or normal-planar processes must be adequately understood to maximize the
device structure. Moreover, because of the progress in coating charge-collection efficiency toward achieving the best PCE.
technologies, PCEs of over 25% can be achieved. Although the
two-step method is still useful for controlling the perovskite 2.1. Electronic and Optical Properties of Perovskites
grain size and morphology, the one-step procedure is more Because the optical properties of solids correlate with their
economical. In one-step spin-coating, kinetic control of the energy band structure, impurity levels, excitons, localized
crystal growth is of critical importance and can be achieved via defects, and lattice vibrations, it is important to study the
an intermediate formed by antisolvent dripping20 or Lewis optical properties of the organic−inorganic lead halide
acid−base adduct formation.21 In Figure 2, two approaches are perovskite. The exciton binding energy for the iodide-based
schematically illustrated, where the antisolvent method is perovskite materials commonly utilized in solar cells was
measured to be a few millielectronvolts at room temper-
ature.22,23 Therefore, the excited charge carriers move as free
carriers. As is well-known, a high efficiency of a PSC is related
to a high absorption coefficient and direct bandgap transition,
as well as a low effective mass, long diffusion length, and
relatively high carrier mobility of free electrons and holes. If
the organic lead halide perovskite is regarded as a semi-
conductor, the optical absorption in the perovskite can be
related to its electronic band structure and, thereby, its crystal
structure. First, we construct the molecular orbital of PbI2
based on the linear combination of atomic orbitals (LCAOs)
of Pb and I. Orbital energies can be calculated by
−Z 2
E = 2eff 13.6 eV ,24 where Zeff and n represent the effective
n
nuclear charge and principal quantum number, respectively.
Zeff was reported to be 13.40 and 11.61 for the 5s and 5p
Figure 2. Schematic illustration of the formation of high-quality orbitals of I, respectively, and 14.10 and 12.39 for the 6s and
perovskite films via the (a) antisolvent dripping method and (b) 6p orbitals of Pb, respectively; these values were obtained from
Lewis acid−base adduct method. An antisolvent is defined as a the orbital exponent expression: ζ = Zeff/n.25,26 The orbital
solvent in which the perovskite is slightly soluble. The figures are energies are, thus, calculated to be E(6p(Pb)) = −57.99 eV >
modified versions of the original figure in ref 21. E(5p(I)) = −73.33 eV > E(6s(Pb)) = −75.10 eV > E(5s(I)) =
C https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 3. (a) Schematic molecular orbital of PbI2 constructed by the linear combination of atomic orbitals (LCAOs) of Pb and I. The orbital
energies were estimated from the effective nuclear charge and the principal quantum number. (b) Schematic illustration of the valence band (VB)
and conduction band (CB) formed from the molecule (0D) state to the solid (3D) state of PbI2. Total density of states (DOSs) for (c) MAPbI3
and (d) FAPbI3. The patterned DOSs represent the upper part of the VB. (c,d) Reproduced with permission from ref 28. Copyright 2018 Elsevier.

Figure 4. Optical absorption for (a) GaAs and (b) MAPbI3. (c) Joint density of states (JDOSs) for MAPbI3 and GaAs. Reproduced with
permission from ref 30. Copyright, 2015 American Chemical Society. (d) Schematic illustration showing the indirect-to-direct transition, achieved
by applying a pressure of over 325 MPa. The weak indirect-bandgap transition was due to the spin−orbit coupling in the CB. Reproduced with
permission from ref 32. Copyright, 2017 Royal Society of Chemistry.

−97.68 eV. On the basis of the orbital energy diagram and 3b). The HOMO has more I 5p characteristics, and the
assuming that the bond angle of I−Pb−I in the gas phase is LUMO tends to have more Pb 6p characteristics. According to
90°, an approximate molecular orbital for PbI2 is constructed, ab initio effective core potential calculations,27 the electronic
as shown in Figure 3a. The overlap between the Pb 6s and I configuration of PbI2 is (a1)2(b2)2(a1)2(b2)2-
5pz orbitals produces σ-bonding and σ*-antibonding orbitals, (a1)2(b1)2(a2)2(b2)2(a1)2 based on 4 valence electrons in
and the I 5px and 5py orbitals result in nonbonding orbitals. (6s)2(6p)2 for Pb and 7 valence electrons in (5s)2(5p)5 for I,
The overlap between the Pb 6p and I 5p orbitals produces one and the HOMO is dominantly a combination of the hybrid 6s
σ-bonding (6pz−5pz) and two π-bonding (6px−5px and 6py− and 6pz(Pb) orbital and the 5pz(I) orbital. A significant Pb → I
5py) orbitals along with the corresponding σ*- and π*- charge transfer was expected based on the double-ζ valence
antibonding orbitals, where the π-bonding orbitals are (DZV) population analysis, which resulted in the electronic
expected to be more stabilized due to s−p mixing. The configurations of (6s)1.98(6p)1.25 for Pb and (5s)1.98(5p)5.40 for
highest occupied molecular orbital (HOMO) is, therefore, I due to a strong 6p(Pb)−5p(I) overlap. This eventually
composed of a σ-antibonding (Pb 6s−I 5p)* orbital, and the implies a direct transition from the VB to CB upon light
lowest unoccupied molecular orbital (LUMO) consists of σ- absorption. The density of states (DOSs) of MAPbI3 (Figure
antibonding (Pb 6p−I 5p)* and π-antibonding (Pb 6p−I 5p) 3c) and FAPbI3 (Figure 3d) were calculated based on cubic
orbitals, which are the bases for the valence band (VB) perovskite with lattice constants of 6.276 and 6.362 Å for
maximum and conduction band (CB) minimum when gas MAPbI3 and FAPbI3, respectively.28 The upper part of the VB
species are condensed into the 3D perovskite material (Figure between 0 eV and ∼−3 eV was confirmed to be mainly
D https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Table 2. Carrier Transport Properties, Including the Lifetime (τ), Diffusion Length (LD), Mobility (μ), and Carrier
Concentration, Depending on the Electrons (e) and Holes (h), Measured under Illumination Using the Hall Effect
Measurement (the Carrier Type Was Not Resolved)a
LD (μm) μ (cm2/(V s)) carrier conc (cm−3)
material τ (ns) e h e h e h
MAPbI3b 40−10 000 2−8 1.8−12 2−40 10−44 7.6 × 1010
(FA1−xMAx)Pb(I1−xBrx)3b 44−40 000 1.7−23 1.8−34 7−26 10−28 8.3 × 1011
a
Reproduced with permission from ref 33. Copyright, 2019 Springer Nature. bPolycrystalline thin film.

composed of the I 5p orbital along with minor contributions short carrier lifetimes is exhibited by direct-bandgap semi-
from the Pb 6s orbital, while the Pb 6p orbital contributed conductors, and weak absorption with long carrier lifetime is
dominantly to the lower part of the CB. exhibited by indirect-bandgap semiconductors), could be
Regarding the origin of the high absorption coefficient (α) explained by the presence of the weak indirect-bandgap
of lead halide perovskite, the optical transition type (direct or characteristics. An indirect-to-direct bandgap transition was
indirect transition) and the time-dependent transition realized by applying a hydrostatic pressure of over 325 MPa on
probability should be considered. Upon light absorption, the MAPbI3 (Figure 4d), resulting in the doubling of the radiative
absorption coefficient is proportional to the transition rate efficiency and an increase in the carrier lifetime.32
(Wi→f) from the initial state (|i⟩), which is associated with the 2.2. Parameters Affecting Charge Transport, Collection,
VB, to the final state (⟨f |), which is associated with the CB, And Recombination
according to Fermi’s Golden rule.29 According to this rule, the
transition probability depends on the transition matrix (Mif = As a working principle in solar cells, photogenerated electrons
⟨f |H′|i⟩, H′ is Hamiltonian) and the joint density of state and holes (carriers) should be separated, transported, and
(JDOS) at a given photon energy (hc/λI), which is simplified collected in electrodes, which result in external quantum
in eq 1: efficiency (EQE). Hence, carrier transport and recombination
are extremely vital in determining the overall photovoltaic
4π 2 performance. In a solar cell, the majority carriers determine the
Wi → f = |Mif |2 ·JDOS overall device architecture, such as the p−n or p−i−n junction,
h (1)
width of the depletion region, and bulk series resistance, while
The transition matrix depends on the coupling of the initial the minority carriers directly influence the recombination
state and the final state, and transition occurs rapidly when the lifetime (τ), diffusion length (LD), and recombination
strength of this coupling increases. The JDOS depends on the coefficients (kn).33 The conventional standard Hall measure-
atomic orbitals because it is related to the DOSs of the CB and ment only provides information on the majority carriers. In
the VB.30 As displayed in Figure 4a,b, optical transition occurs 2019, Shin’s research group in KAIST collaborated with IBM
from the (6s(Pb)−5p(I))* antibonding orbital to the (6p- to provide more information on the mobility of electrons and
(Pb)−5p(I))* antibonding orbital. As compared to the optical holes in ambipolar perovskite materials, (FAP-
transition in direct bandgap GaAs from the As p orbital to the bI3)1−x(MAPbBr3)x. In Table 2, the carrier transport proper-
Ga s orbital, the p → p transition in perovskite is more efficient ties, including τ, L D , the mobility (μ), and carrier
because the CB with the p character is less dispersive than that concentration, are presented based on photo-Hall measure-
with the s character from the viewpoint of JDOS, as shown in ments.33 Because the measurements were carried out under
Figure 4c.30 The direct bandgap with the p → p transition is illumination, the parameters can be applied under the
highly suitable for achieving a high absorption coefficient. operational condition. When comparing the highest values of
When considering that the CB in semiconductors is influenced the transport parameters between MAPbI3 and the cation/
by metal cations, elements with lone pair electrons are anion mixed perovskite (note that FAPbI3 is dominant in the
expected to deliver the p character at the CB. Thus, In1+, mixed perovskite), most of the parameters are similar;
Ge2+, Sn2+, Pb2+, Sb3+, and Bi3+ are expected to exhibit high however, τ, LD, and the carrier concentration for the mixed
absorption coefficients if the compounds with these metal perovskite are slightly higher than those for MAPbI3 except for
cations have an optical direct bandgap. How do we identify μ, implying that the photovoltaic performance of the FAPbI3-
direct- or indirect-bandgap materials? It is well-known that based mixed perovskite might be better than that of MAPbI3.
spectroscopic measurements, such as a simple Tauc plot from a Next, we discuss how these parameters affect the photovoltaic
UV−vis spectrum, together with theoretical calculation, can performance. For instance, LD is considered when determining
provide an answer to this question. Facilitating an indirect-to- the device structure, whether mesoscopic or planar (Figure 1).
direct transition by material engineering is another approach to Because LD is proportional to the diffusion coefficient (D) and
finding new light-absorbing materials with high absorption τ, i.e., L D = D τ , the diffusion coefficient for electrons (De)
coefficients. For instance, Cs2AgBiBr6 perovskite was reported and holes (Dh) can be estimated from LD and τ. The minimum
to have an indirect bandgap, which could be changed to a required carrier diffusion coefficient (Dm) was derived with the
direct bandgap when Bi3+ was alloyed with In3+ and/or Sb3+.31 assumption that the absorbance of the perovskite film is greater
It was reported that a strong spin−orbit coupling (SOC) in than 1 (more than 90% of the incident light is absorbed), the
MAPbI3 can influence the optical transition, where a weak charge injection yield is greater than 90%, and the charge
indirect bandgap transition was observed due to Rashba collection yield is 100%. If charge injection is limited by charge
splitting in the CB.32 The strong absorption and long charge- 10
carrier lifetime for MAPbI3, which could not be achieved in diffusion in the perovskite, Dm can be estimated by Dm = α2 τ
conventional semiconductors (where strong absorption with (α represents the absorption coefficient).34 According to the
E https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 5. (a) Grain-size-dependent mobility for MAPbI3. The filled squares with error bars represent the measured data, and the blue solid line
represents the best-fit result obtained using a Kubo relation (see ref 38 for more details). The dotted line represents the fit result with two data
points. Reproduced with permission from ref 38. Copyright, 2016 American Chemical Society. (b) Mobility as a function of the bromide fraction
(x) in FAPb(BrxI1−x)3 and Cs0.17FA0.83Pb(BrxI1−x)3. The arrow indicates that the partial substitution of FA with Cs ions improved the carrier
mobility. Reproduced with permission from ref 35. Copyright 2017 American Chemical Society.

Figure 6. (a) Schematic illustration of the recombination process, where the first-order SRH, second-order radiative, and third-order Auger
recombinations are present. The interface recombination is related to the presence of an electron transport layer (ETL). (b) The recombination
rate, dn/dt, is given as a function of the charge density, n, for MAPbI3. Reproduced with permission from ref 41. Copyright 2016 Springer Nature.
(c) Effect of nonradiative recombination on the current−voltage curves, and (d) difference in the quasi-Fermi level between the CB and the VB
depending on the interfaces. Reproduced with permission from ref 42. Copyright 2017 Springer Nature.

estimation, a mesoscopic structure containing mesoporous crystal, etc.) have been reported,35,36 where a single crystal
TiO2 is recommended for MAPbI3 because De is lower than exhibited a higher μ than a polycrystalline thin film. This is
Dm, although Dh is higher than Dm. For GaAs, the planar reasonable when considering that LD increased with the
structure is preferable because De and Dh are much higher than perovskite grain size.37 Nevertheless, it cannot be generalized
Dm.32 D can also be evaluated from μ via the following Einstein because of experimental uncertainty. Excluding the intrinsic
relation: μ = eD/(kBT), where kB is the Boltzmann constant, factors limiting μ, we investigate the extrinsic factors affecting
which indicates that a high mobility can lead to a high LD. μ. Evidently, the grain size influences the carrier mobility, as
Charge-carrier mobility values depending on the perovskite mentioned previously.38 Figure 5a shows that the carrier
compositions and architectures (film, polycrystalline, single mobility increases with the grain size; however, it reaches a
F https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 7. (a) Perovskite crystal structure. For photovoltaically interesting perovskites, the large cation (A) is usually the methylammonium ion
(CH3NH3), the small cation (B) is Pb, and the X anion is a halogen ion (usually I, although both Cl and Br are also receiving attention). For
CH3NH3PbI3, the cubic phase forms only at temperatures above 330 K. Reproduced with permission from ref 48. Copyright 2015 Elsevier.
Schematic structural representation of MAPbI3, where MA = methylammonium in the (b) pseudocubic and (c) tetragonal phases. The thick
dashed lines indicate the unit cell, which contains C (black), N (green), H (white), I (violet), and Pb (light gray) atoms; the highest symmetry
phase for MAPbI3 and its related materials is the cubic (Pm3̅m) lattice phase, with sequential transitions lowering the symmetry typically through
octahedral tilting. The tilting collectively leads to the transition from the cubic to tetragonal phase. Reproduced with permission from ref 50.
Copyright 2013 American Chemical Society.

saturation point at grain sizes larger than 100 nm. Energetic without nonradiative recombination because the Voc and FF
disorder was reported to change the carrier mobility, where values approach the Shockley−Queisser limit values of 1.33 V
structural changes with low crystallinity in FAPb(Br1−xIx)3 and 0.91, respectively.42 Because Voc is logarithmically
were observed at x ≈ 0.3−0.5, at which point μ was proportional to the photoluminescence external quantum
significantly reduced.39 This challenge could be overcome by yield (EQYPL), i.e., Voc = Voc,rad + kT/q[ln(EQYPL)], Voc
introducing the Cs ion, which eliminates the mobility gap,40 as reaches a theoretical value when EQYPL = 1 for only radiative
shown in Figure 5b. recombination. For nonradiative recombination, EQYPL will be
2.3. Recombination in Perovskite Materials and Solar Cells
<1, and thus Voc is expected to be lower than Voc,rad. For
instance, for EQYPL = 10−3 at 25 °C, Voc will be ca. 173 mV
The recombination of charge carriers in semiconductors can be lower than Voc,rad (kT/q = 25 mV at 25 °C). In Figure 6d, Voc,
simply expressed by the following relation: dn/dt = −k1n − represented by the difference in the quasi-Fermi level between
k2n2 − k3n3, where n is the electron charge-carrier density, k1 is the CB (FCB) and VB (FVB), tends to decrease as EQYPL
the first-order Shockley−Read−Hall (SRH) trapping (non- decreases, which is caused by the creation of interfaces by the
radiative) rate constant (also called monomolecular recombi- electron transport layer (ETL) and/or hole transport layer
nation), k2 is the second-order band-to-band (radiative) (HTL). This clearly indicates that interface engineering is of
recombination rate constant (also termed “bimolecular critical importance in increasing the photovoltaic performance
recombination” due to the reaction of e + h → γ), and k3 is toward achieving high theoretical efficiency.43,44 As one of the
the third-order Auger (nonradiative) recombination rate factors responsible for the nonradiative recombination, the
constant.40 Because monomolecular recombination is related lattice strain was reported to enhance nonradiative recombi-
to traps, the SRH recombination is also termed “trap-assisted nation,45 which highlights that structural modification also
recombination”. The recombination rate constants were plays an important role in the modulation of the recombina-
estimated to be k1 = ∼107 s−1 (τ = ∼100 ns) for SRH, k2 = tion process.
∼10−10 cm3 s−1 for radiative recombination, and k3 = ∼10−28
cm6 s−1 for the Auger recombination, for both MAPbI3 and 3. KEY CHALLENGES IN ACHIEVING HIGH
FAPbI3.40 The trap-assisted recombination (SRH) was EFFICIENCY
reported to be accelerated by impurities and crystal disorder.39
In Figure 6a, the recombination process is schematically 3.1. Structure Design
displayed from the electron point of view. The dependence of 3.1.1. Crystal Structure. MAPbI3 has a crystal structure of
the recombination rate (dn/dt) on the light intensity, perovskite with a general chemical formula of ABX3 in a simple
represented by the charge-carrier density, is displayed in cubic lattice.46 The space group is Pm3̅m. In principle, any
Figure 6b, where at low light intensities, the recombination combination of compatible ions can form a stable perovskite
kinetics is dominated by the monomolecular process attributed crystal, where the tolerance and octahedral factors are satisfied
to the unfilled trap states, whereas the bimolecular kinetics within an empirically determined range, depending on the
dominate at high light intensities because the traps are filled, ionic radii of the constituents (this will be discussed in more
and consequently, there is a substantial number of free detail in section 3.2.).47 The anion corner-shared 3D network
electrons and holes.41 At very high light intensities or carrier of the PbI6 octahedra can be viewed in Figure 7a.48 As a simple
densities over 1018 cm−3, Auger recombination appears to be cubic lattice, the Pb ions are centered in the octahedral, and
significant. Here, we discuss the effect of the nonradiative the MA ions are located between the octahedral building
recombination on the photovoltaic performance. The trap- blocks.49 The highest symmetry phase for MAPbI3 and its
assisted SRH recombination tends to decrease both Jsc and Voc, related materials is the cubic (Pm3̅m) lattice, with sequential
as compared to an ideal device, which allows only radiative transitions lowering the symmetry typically through octahedral
recombination, as shown in Figure 6c. Voc is expected to tilting. The tilting collectively leads to the phase transition
decrease further when interface recombination exists. A PCE of from cubic to tetragonal. (Figure 7b,c)50 The collective
30.5% can be achieved from an ideal single-junction device rotation of the PbI6 octahedra around the c axis results in
G https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

the tetragonal space group, I4/mcm, with relatively close in the formation of the electronic structures; rather, they
packing within the ab plane. Further transition to an stabilize the structure of the perovskite and change the lattice
orthorhombic phase (space group: Pnma) is accompanied by constant.60,61 As a result, MAPbI3 and its related materials are
a tilting of the PbI6 octahedra out of the ab plane. MAPbI3 direct-bandgap semiconductors with strong band optical
transforms from the cubic to the tetragonal phase at absorption and luminescence. Essentially, the valence band
temperatures below 315−330 K, and the tetragonal to the maximum (VBM) and conduction band minimum (CBM) lie
orthorhombic phase at ∼160 K. These transitions need to be at the same point in the reciprocal space (k-space). Note that
critically considered when assessing charge-carrier dynamics recent theoretical calculations suggest that there may be subtle
because they may alter the electronic band structure and, yet important exceptions that arise in noncentrosymmetric
consequently, the optoelectronic properties of the material. crystal structures in MAPbI3 due to SOC.56,62 The Rashba
The group of hybrid perovskites exhibits crystallographic spin-splittings with CBM are not located at the same high-
phase transitions with varying temperatures and pressures. symmetry points of the Brillouin zone (R point); instead, they
Heat capacity measurements along with temperature-depend- are slightly shifted. In other words, the CBM and VBM are not
ent X-ray diffraction (XRD) studies revealed a crystallographic located at the same points, having different Rashba momentum
phase transition from cubic to tetragonal at temperatures of offsets.63,64 This result indicates the formation of an indirect
330.8, 236.1, 177.2, 238, and 403 K for MAPbI3, MAPbBr3, bandgap, which significantly reduces the radiative recombina-
MAPbCl3, FAPbBr3, and CsPbBr3,51 respectively. The cubic tion rate. Conversely, because the indirect bandgap is generally
structure contains only one formula unit (Z = 1); therefore, the only a few tens of millielectronvolts smaller than the direct
noncentrosymmetric MA ions must be randomly oriented to optical transition gap, it does not significantly affect the
satisfy the Oh symmetry. As the temperature decreases, absorption spectrum.65 Furthermore, the direct bandgap
tetragonal and orthorhombic phases are stabilized with characters in the strong optical absorption and indirect
accompanying ordering of MA molecular ions.52 The structural bandgap characters in the slow recombination are responsible
transition from the cubic to tetragonal phase occurs due to the for the high efficiencies of MAPbI3 solar cells. Both Pb and I
reorientation of MA+, as observed by NMR studies, where the are heavy ions in MAPbI3, and therefore, both the VB and CB
number of disordered states of MA+ was lowered from 24 in exert considerable relativistic effects; that is, for a quantitative
the cubic phase to 8 in the tetragonal phase. This molecular treatment of the electronic band structure, the SOC must be
ordering is completed in the orthorhombic phase such that at included. SOC causes a decrease in the electronic bandgap due
161.8, 149.8, 171.5, 150, and 385 K for MAPbI3, MAPbBr3, to the splitting of the gap. Prediction of the bandgap by a state-
MAPbCl3, FAPbBr3, and CsPbBr3, respectively, an ortho- of-the-art quasiparticle self-consistent GW with SOC correc-
rhombic Pnma space group is formed. Some studies have tions is the only method that can accurately predict the
reported the crystal structure of MAPbX3 (X = I, Br, Cl) to electronic structure of MAPbI3. The calculated direct bandgap
comprise noncentrosymmetric space groups, such as I4cm and of MAPbI3 is Eg ∼ 1.67 eV, which is in good agreement with
Pna21 for the tetragonal and orthorhombic phases, respec- the measured value of 1.61 eV by room-temperature
tively. Nandi et al. refined the high-resolution synchrotron photoluminescence (PL). The computed bulk ionization
XRD data both in centrosymmetric and noncentrosymmetric potential and electron affinity are 5.7 and 4.1 eV,
space groups and reported relatively good reliability parameters respectively.66
in the centrosymmetric structure.53 Irrespective of the 3.1.2. Defect Structure. Thus far, MAPbI3 and its related
temperature-dependent phase transitions, hybrid perovskites materials have mainly been synthesized by a low-temperature
exhibit crystallographic phase transitions with applied pressure. solution process. This process inevitably produces many
Reportedly, in the literature, the crystal structure of MAPbI3 is different types of defects, including thermodynamically stable
tetragonal (I4/mcm) at ambient temperature and pressure and point defects, interfaces, surfaces, and grain boundaries
has a distorted orthorhombic phase with the space group Imm2 (GBs).67 Generally, the Voc deficit arises from a trap-assisted
when the pressure is above 0.3 GPa. Further, reversible electron−hole nonradiative recombination. The Voc decrease
amorphization was realized for this system when a pressure of in PSCs is rather small compared to that in other inorganic
10 GPa was applied.54 However, MAPbBr3 is in the cubic semiconductors, indicating that defects play a relatively minor
phase with space group Pm3̅m at ambient pressure, and has a role in PSCs. Nonradiative recombination and carrier
cubic space group (Im3̅) when a pressure of 1.7 GPa is scattering are often caused by defects that generate deep-
applied.55 level states in the bandgap. Most structural defects, including
Many theoretical studies have improved our fundamental surfaces and GBs, occur in deep-level states.68 Unusually, deep-
understanding of electronic structures.56,57 In contrast, few level defects with high formation energy are rare, whereas
fundamental and experimental studies have been reported.58 In shallow trap states in the bandgap are abundant in MAPbI3,
2014, the electronic structures of TiO2/MAPbI3 interfaces resulting in defect-tolerant materials. These unique defect
were investigated using hard X-ray photoelectron spectroscopy. properties are also attributed to the strong Pb lone-pair s−
The simulated DOS of the VB describes the experimental halogen p antibonding coupling and the ionic characteristics, as
spectra adequately and shows that the outermost levels consist mentioned in the previous section.69,70 Understanding and
of Pb and I orbitals.59 controlling defects to suppress nonradiative charge recombi-
As discussed earlier, MAPbI3 exhibits an inverted band nation are essential for further improvement of the PSC
structure compared with conventional inorganic semiconduc- efficiency.69,71
tors, and the MA+ ion contribution to the electronic states is First-principles calculations have been used to study the
far from the band edges. In other words, occupied molecular point defect properties of MAPbI3 and its related materi-
states are found well below the top of the VB, and empty als.30,72−75 The growth of the MAPbI3 phase can occur only in
molecular states are found well above the bottom of the CB. a long and narrow range of chemical potential, indicating a
Consequently, the organic molecules do not play a direct role small experimental window in terms of thermodynamic
H https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

variables, e.g., precursors, pressure, and temperature. All


possible intrinsic point defects include three vacancies (VMA,
VPb, VI), three interstitials (MAi, Pbi, Ii), and four antisite
substitutions (MAI, PbI, IMA, IPb). All the vacancy defects and
most interstitial defects, such as MAi, VPb, MAPb, Ii, VI, and
VMA, exhibit rather shallow transition energy levels. The
shallow level defects of both, donors and acceptors, have
comparable low defect formation energies resulting in n- and
p-type electronic systems in MAPbI3.76 Theoretical bipolar
conductivities have already been confirmed by doping
experiments.77 In MAPbI3, the dominant donors, MAi and
VI, and acceptors, VPb and MAPb, have comparable formation
energies; therefore, both p- and n-type dopings are possible.
The shallow acceptors originate from the covalent coupling
between the Pb lone-pair s orbital and I p orbitals, which
increase the VBM so that the acceptors are generally shallower
than those in the cases without strong sp coupling. In contrast, Figure 9. Total and projected densities of states (DOS) of pristine
the shallow donors originate from the high ionicity. The MAPbI3 (top panel) and MAPbI3 with the I interstitial defect
calculated transition energy levels of the point defects in (bottom panel). The insets show the charge densities of the (a) VBM
MAPbI3 and MAPbBr3 are shown in Figure 8.78 For the and (b) CBM in pristine MAPbI3, and the (c) VBM, (d) CBM, and
(e) trap state in MAPbI3 with the I interstitial defect. Reproduced
with permission from ref 71. Copyright 2019 American Chemical
Society.

interstitial defect (bottom panel), showing I interstitials near


the VBM. The recombination of the trapped holes with the CB
electrons is several times slower than that of the VB holes with
the CB electrons. These factors can lead to charge reduction
and energy losses in PSCs with I interstitial defects and,
consequently, contribute to the high efficiency of the PSCs.
Pb vacancies, VPb, also contribute to the defect chemistry of
MAPbI3 with a low defect formation energy. VPb is less stable
in MAPbBr3 than in MAPbI3. VPb is stable in the −2 charge
state in MAPbI3, showing comparable defect density with
those of Ii− and MAi+. The oxidation of VPb2− with two holes
leads to the oxidation of lattice halides with the formation of Ii+
and an I vacancy. The effect of VPb on the charge-carrier
Figure 8. Thermodynamic ionization levels of most stable defects in dynamics in MAPbI3 was studied using the Redfield theory
MAPbI3 and MAPbBr3. Reproduced with permission from ref 78. formalism, and it was experimentally confirmed that VPb
Copyright, 2019 John Wiley and Sons. increases the nonradiative relaxation time of the excited
electrons.79 For FAPbI3, the effect of VPb was also investigated
simulated MAPbBr3, one can notice quite similar defect theoretically and experimentally; it was concluded that the
chemistries in the two perovskite materials. The defects that effect was similar to that in MAPbI3.80,81 Although PSCs based
generate deep levels are IMA, IPb, Pbi, MAi, and Pbi, which are on FAPbI3 with mixed A-site cations of MA and Cs exhibited
mostly cation or antisite defects, except for Pbi, having high much higher PCEs, only a few first-principles calculation
formation energies. The most stable defects in both MAPbI3 studies of intrinsic defects in FAPbI3 can be found in the
and MAPbBr3 are MA and halide interstitials, i.e., MAi+ and literature.82 In this study, it was calculated that defects VFA,
Ii−/Bri−. MAi is not covalently bonded to the Pb−I framework FAI, and IFA show much lower formation energies in FAPbI3
and does not create additional gap states. Thus, MAi defects than in MAPbI3. Antisites FAI, and IFA create deep-level states
are stable only in the positive charged states within the Fermi in the bandgap and can act as recombination centers, resulting
level range spanning the perovskite bandgap, and they exhibit a in reduced carrier lifetimes and low Voc values in FAPbI3-based
shallow (+/0) transition close to the CBM. In the physical cells. They also found that by mixing the cations of MA and FA
image, the cation occupies the interstitial positions on the in perovskites, the formation of these defects can be
facets of the Pb−X6 (X = I, Br) octahedra, with associated substantially suppressed, which would lead to enhanced PCEs.
densities of ∼1012 cm−3. Ii+ exhibits relatively high stability, i.e., The formed halide vacancies may couple with metal
a lower defect formation energy, compared to Bri+, leading to a vacancies through the formation of Schottky defects. As
sizable defect density of ∼1010 cm−3. The relatively high reported by Walsh et al., these defect couples have low
stability of positive I interstitials is mainly due to the lower formation energies and do not introduce electronic disorders.83
electronegativity of I compared to Br. Interestingly, an The Schottky formation energy of 0.14 eV per defect is
interstitial I atom introduces deep trap states near the VBM, remarkably low and corresponds to an equilibrium vacancy
and holes are trapped in the states; the trapped holes exhibit concentration of 0.4% at room temperature. In comparison,
surprisingly long lifetimes. Figure 9 shows the calculated DOSs the reported Schottky formation energy for BaTiO3 is 2.29 eV
of pristine MAPbI3 (top panel) and MAPbI3 with the I per defect, which results in parts-per-million (ppm) levels of
I https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

equilibrium vacancy concentrations. The Sargent group screening and compensation of the charged defects in the
provides a complete electronic band diagram of a single-crystal MAPbI3 lattice. The traps are energetically shallow and
MAPbI3 using the in-gap electronic state spectrum. The high electrostatically screened, which minimize their roles in
mobility and long diffusion length of both electrons and holes nonradiative recombination processes.90,91
are directly measured, as well as the extremely low 3.1.3. Microstructure. Highly efficient PSCs are generally
concentration (nt(VB) = 3 × 1010 cm−3; nt(CB) = 3 × 1011 produced by simple solution processes. The simple solution
cm−3) with ambipolar charge carriers.84 The halide perovskites method guarantees low-cost photovoltaic devices. Spinning of
show an ultralow density of trap states even at low synthesis the precursor solution often results in polycrystalline thin films
temperatures. The most easily formed defects in perovskites on transparent, conducting substrates. Polycrystalline Si,
have little influence on the charge-carrier lifetimes, ration- Cu(In1−xGax)Se2 (CIGS), CdTe, and PSCs are all parts of
alizing the high efficiencies of PSCs. polycrystalline thin-film solar cells. The PCE and other solar-
Conventional semiconductor technologies are centered on cell performance parameters, such as stability, are largely
the production of perfect crystals to reduce defects, as dependent on the microstructure of the thin films.
exemplified by single-crystalline Si with remarkably low defect A GB is a critical defect in polycrystalline functional
densities. However, the synthesis of MAPbI3 by solution materials. GBs can affect the mechanical, chemical, and
processing, at near room temperature, inevitably results in a electrical properties of a material. The latter is important for
high density of defects, which may be detrimental to its solar cells because GBs modify the charge-carrier transport
performance. Whereas nonradiative charge recombinations properties and, thus, the performance.92 In GBs, a lattice
involving defect states are significant in traditional semi- discontinuity, incomplete and/or broken bonds resulting in
conductors, the reported charge-carrier lifetimes are excep- dangling bonds, and depletion and/or segregation of impurities
tionally long in perovskites, ranging from a few nanoseconds to can be expected. The defects, dislocations, misplaced atoms
hundreds of nanoseconds,85 indicating that the defect states in (interstitials), vacancies, distorted bond angles, and bond
perovskites are indeed benign electrically. In this context, distances at GBs can act as recombination centers and/or trap-
defect-tolerant MAPbI3 materials are still not fully understood. states for the charge carriers. Depending on their trap states
According to the original concept of defect tolerance, the and/or centers, the GBs in a polycrystalline thin-film
presence of an antibonding upper VB and a bonding lower CB, photovoltaic material play an important role that could be
as in MAPbI3, signifies that dangling bond defects would be either benign or detrimental to the cell performance. In
repelled quantum mechanically into the continuum bands, polycrystalline Si thin-film solar cells, the dangling bonds at the
leaving the bandgap clean and allowing for the formation of GBs create deep-level states within the bandgap and act as a
shallow defects. The second point to consider is the screening nonradiative recombination center. Reducing the density of
in the dielectric medium. If charged defects are present in the dangling bonds in a polycrystalline Si absorber by hydrogen
host material, their influence should be minimized. The passivation is a critical microstructure engineering solution for
dielectric constant represents the ability of a material to screen GBs.93 Meanwhile, deep-level states are not found at the GBs
an electrostatic perturbation. The higher the dielectric constant of CuInSe2 absorber films. GBs are calculated to be electrically
of a material, the more effective the screening, and thus, the benign.94 Many experimental studies have also demonstrated
higher the defect tolerance of the material. The dielectric the beneficial effects of the GBs in CZTS, CZTSSe, CIGS, and
constants of MAPbI3 are approximately three times higher than CdTe films, toward enhancing minority carrier collections. The
those of other thin-film photovoltaic materials, such as CdTe GBs in these solar absorbers are electrically charged by
and Cu2ZnSnS4.86,87 The last factor under consideration is the impurities or different types of vacancies, and they create a
carrier effective mass. A low mass favors free charge carriers downward band bending toward the GBs in mostly p-type
and can support high carrier mobility and electrical absorbers. Thus, the photogenerated electrons and holes near
conductivity. The spatial localization of electrons and holes the GBs will be easily separated by the built-in field at the GBs,
at charged defect sites should be avoided as they retard the and the recombination rate of these carriers will be lower at the
transport of charge carriers and are associated with thermal GBs.95 GB engineering is, therefore, inevitable for high-
energy losses. The effective mass is also a critical factor in efficiency solar cells.96
avoiding the formation of small polarons. There is competition Many studies on the GB in polycrystalline semiconducting
between the kinetic energy of the free carriers and the potential organic−inorganic metal halides for PSCs have been reported;
energy gained by localization in the lattice. Typically, metal however, the exact role of GBs is yet to be clarified.97 One
oxides often have a high hole effective mass due to the aspect is clear though, that the trap density at perovskite
localization of the O 2p orbitals that form the VB; this leads to surfaces and GBs is much higher than in bulk because the
the formation of small polarons, which make it difficult to find observed carrier diffusion length is much greater in single
p-type oxides. crystals than in polycrystalline MAPbI3 films. Numerous
In MAPbI3, the unique electronic structure results from theoretical as well as experimental studies have reported
corner-sharing in the PbI63− octahedral framework with overall controversial results regarding the effect of GBs on the
stoichiometric PbI3− and a sublattice of MA+ cations in the perovskite photovoltaic performances.98 First-principles calcu-
cuboctahedral voids, as discussed earlier. The framework with lations predicted that the GBs in MAPbI3 are benign simply
the corner-sharing octahedra is overly intrinsically soft and because no deep trap states were found. Haruyama et al.
exhibits structural dynamics. The relatively light electrons and/ systematically studied the thermodynamic properties of
or holes in the CB are rapidly protected to form a large polaron MAPbI3 surfaces and concluded that the PbI2-terminated
with phonons.88 In addition, the large polaron can screen the surface does not introduce subgap states.99,100 Yin et al.
electrostatic potential drastically and, thus, reduces its reported GBs with intrinsically benign properties by first-
scattering with charged defect sites.89 Consequently, the high principles calculation. GBs such as Σ5 (310) and Σ3 (111) do
polarizability in the soft and ionic lattice can enable adequate not generate deep states in the bandgap, making polycrystalline
J https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 10. HOMO and LUMO charge density in (a) pristine MAPbI3, (b) Σ5 (012) GB, and (c) Cl-doped Σ5 (012) GB; the VBM and CBM are
delocalized in (a), and become localized in inorganic atoms in the vicinity of the GB in (b) and (c). Doping with Cl atoms pushes the charge
density away from the GB. DOSs of the (d) pristine MAPbI3, (e) Σ5 (012) GB, and (f) Cl-doped Σ5 (012) GB obtained using optimized zero-
temperature geometries. The zero energy is set to the Fermi level. Reproduced with permission from ref 103. Copyright 2019 American Chemical
Society.

halide perovskite thin films behave single-crystal-like.101 The Many scanning probe techniques have been employed to
same research group further studied and reported that the Σ5 investigate the local properties of perovskite microstructure
(310) GB forms shallow defects states close to the VBM. They conjunctions. In particular, atomic force microscopy (AFM)
also showed that the structural disorder at GBs, which induces and Kelvin probe force microscopy (KPFM) have been
the formation of Pb−Pb bonds, creates no subgap states employed to explore the local electronic structures of
because of the relatively large distance between the two Pb2+ perovskite photovoltaic absorbers. Both techniques typically
ions. The shallow trap states are not as critical as the deep provide information based on the measurement of the surface
levels; possibly, the states may increase the effective mass of potential differences and local conductivity. Yun et al.
holes and/or trap the holes to decrease the hole mobility, investigated the effect of high-density GBs in the presence of
thereby inducing nonradiative SRH recombination.102 These a MAPbI3 absorber by KPFM and conductive-AFM (c-AFM).
benign GBs can, in part, explain the long carrier diffusion Their KPFM results showed a relatively low contact potential
length and high efficiency reported with lead halide perov- difference (CPD) at the GBs under the dark condition,
skites. implying that a downward band bending occurred in the GBs.
Meanwhile, another theoretical study involving a time- The higher current collection observed by c-AFM near the
domain atomistic simulation via nonadiabatic molecular GBs is consistent with the KPFM results, which indicates that
dynamics combined with the density functional theory photogenerated carriers are more efficiently separated and
(DFT) showed that while GBs may act as major recombina- transported along the GBs.107 Chen et al. also employed
KPFM to study the CPD between and near the GBs. They
tion sites in a pure iodide material, they can be
found that the band bending near the GBs and the band
passivated.103−105 By breaking the perovskite crystalline
bending of the grains to the GBs switched from downward to
periodicity, GBs create localized states, which overlap less
upward when the organic species were released to form PbI2.
than the CBM and VBM in the pristine perovskite (Figure
This indicates that PbI2 could effectively passivate the GBs and
10).103 These studies demonstrated that GBs strongly prevent charge recombination at the GBs.108 Scanning
accelerate the electron−hole recombination relative to the tunneling microscopy (STM) combined with scanning
bulk MAPbI3, while the introduction of substitutional Cl tunneling spectroscopy (STS) was also employed to provide
dopants at the boundary reduces the recombination. Once the local electronic DOSs with spatial resolution at the
relatively light Cl ions substitute relatively heavy I ions at the nanometer scale or even atomic scale. M.-C. Shih et al. utilized
GB, they introduce high-frequency vibrations and accelerate a light-modulated STM technique to reveal the correlation
the quantum coherence loss. Furthermore, in contrast to I, the between the nanoscale compositional distribution and the
Cl dopants do not contribute to the hole wave function; interfacial electronic structures of the heterointerfaces in PbI2-
consequently, they eliminate some of the wave function passivated MAPbI3 films. They revealed the important role of
overlap, reduce the nonadiabatic electron−phonon coupling, the optimum PbI2 passivation layers in charge separation and
and restore the value of the material close to that of the recombination at the perovskite crystal grains.109 In contrast,
pristine MAPbI3. It is suggested that charge recombination Chu et al. conducted the quantitative nanoscale photo-
depends on the type of GBs created and that annealing of conductivity imaging of MAPbI3 thin films by light-stimulated
perovskite films can cure GB defects, which would, in turn, microwave impedance microscopy. The microwave signals
eliminate midgap states and stabilize the GBs, thereby were largely uniform across grains and GBs, suggesting that the
prolonging the charge-carrier lifetime.106 microstructures do not cause strong spatial variations in the
K https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 11. I−V curves for a series of different voltage scanning rates measured on the (a) grain of a MAPbI3/TiO2/FTO/glass sample by AFM and
(b) an Au/MAPbI3/TiO2/FTO/glass structure using macroscopic probes in the nonilluminated state. (c) Topography and (d) corresponding
current image of the MAPbI3 thin films obtained by applying −1.5 V to the AFM tip. (e) Current image overlaid on a 3D topography image of the
white boxed region, as indicated in (c) and (d). (f) Schematic representation of the origins of I−V hysteresis, i.e., ionically active GBs. Reproduced
with permission from ref 111. Copyright 2017 John Wiley and Sons.

photoresponse. The measured photoconductivity and lifetime among grains and uniform across the GBs and interiors.
are strongly affected by bulk properties, such as the Confocal fluorescence microscopy was suggested as a powerful
crystallinity.110 Using temperature-dependent PFM and scan- tool to investigate the spatially as well as the temporally
rate-dependent nano/macroscopic I−V measurement, ion resolved charge-carrier recombination dynamics, providing
migration through GBs was determined as the primary origin insight into the potential sources of nonradiative loss in
of the I−V hysteresis. The relatively high current passing perovskite. M. Yang et al. examined GBs in high-quality
through the grains was measured as the voltage scan rates were micrometer-sized MAPbI3 thin films using high-resolution
increased, by local I−V measurement using a c-AFM, and confocal fluorescence-lifetime imaging microscopy with kinetic
opposite I−V behaviors were observed through the GBs, as modeling of the charge-transport and recombination processes.
shown in Figure 11.111 Further investigations are required to They found that the GBs generally display lower luminescence
elucidate the local intrinsic properties of GBs. intensities than the grain surfaces and/or grain interiors;
The consensus is that charge-carrier trapping is the additionally, the lifetimes at GBs are not poorer than those of
dominant recombination pathway under typical operating the grain surfaces/interiors, suggesting that the GBs do not
conditions owing to the slow recombination of free charge dominate nonradiative recombination.116 Through transient
carriers despite the low trap density. None of the spectroscopic reflection spectroscopy, the same group also observed that the
techniques have provided information on the spatial locations charge-carrier lifetime is mainly limited by surface recombina-
of the trap states or their associated chemical/physical tion.117 Moreover, Snaider et al. employed transient absorption
nature.112 Spatially resolved measurements, mostly based on microscopy (TAM) to directly measure the carrier transport
PL imaging, combined with microstructural characterization across GBs in hybrid organic−inorganic perovskite thin films
techniques, such as confocal microscopy, scanning electron with 50 nm spatial as well as 300 fs temporal resolutions.118
microscopy (SEM), and AFM, are required to study the While carrier transport across GBs is limited to some extent,
microstructural contribution in highly efficient PSCs. M. J. the carrier diffusion constants are only reduced by approx-
Simpson et al., in their early work, utilized a combination of PL imately a factor of 2 when the grain sizes are >1 μm. For a film
and femtosecond transient absorption microscopic techniques with a grain size on the order of 200 nm, carrier transports
to map the trap states by spatial resolution.113 The localized across multiple GBs have been observed within a time window
optical response of the MAPbI3 films, obtained by spatial of 5 ns, although they were not spatially uniform. Their results
resolution and correlated emission/absorption measurements, demonstrated how grains oriented between grains affect carrier
was used.114 Their results strongly implied the existence of transport. Adequately large grains, arranged in the depth
other factors, beyond GB-related nonradiative recombination direction of the film, can also play a critical role in high-
channels, which led to significant intrafilm optical hetero- performance PSCs. The preferential orientation of tetragonal
geneities in the perovskite films. The effect of the significant perovskite grains can thus enhance the charge-carrier lifetime,
morphological heterogeneity of the polycrystalline perovskite resulting in superior photocurrent and high electric potentials
films on their photovoltaic performance, determined by PL in PSCs.119,120 Kim et al. fabricated highly preferred-oriented
microscopy, was also reported by W. Tian et al.115 They and enlarged perovskite grains by pressure-induced crystal-
studied the carrier lifetime, intra- and intergrain carrier lization with low-angle GBs, which led to improved solar cell
diffusion length, and carrier extraction efficiency with performance.121
submicrometer spatial resolution. Contrary to the common In contrast, de Quilettes et al. observed PL quenching in the
belief concerning the deleterious effect of the morphological GBs of nonstoichiometric organic−inorganic perovskites,
heterogeneity on the carrier lifetimes and diffusivities, in neat MAPbI3(Cl), by confocal fluorescence microscopy in con-
MAPbI3(Cl) polycrystalline films, the local carrier diffusivities junction with SEM to spatially resolve the PL decay dynamics.
in different grains are all surprisingly high. Furthermore, the Although the PL intensities and lifetime varied between
local carrier lifetimes are long and surprisingly homogeneous different grains in the same film, the GBs were relatively dim
L https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 12. (a) ABX3 cubic perovskite crystal structure and the corresponding lengths expressed by the ionic radius. (b) Band structure of
CH3NH3PbI3. (c) Total DOS (top panel) and contributions from CH3NH3+, Pb, and I partial DOSs (bottom panels), respectively. The zero in
DOS refers to the VBM. The Pb partial DOS is enlarged by five times for a clear indication of the s orbital contribution. Reproduced with
permission from ref 72. Copyright 2014 American Institute of Physics.

and exhibited fast nonradiative decay.122,123 The GBs in their and not the physicochemical interactions between the
study are not benign, as has been suggested previously. constitutive ions, it provides useful insights for estimating the
Chemical treatment with pyridine could activate previously structural formability and intrinsic stability of oxide perovskite
dark grains. Ham et al. also reported the acceleration of charge- materials. For hybrid perovskite materials, in fact, the
carrier recombination at the GBs of MAPbI3 films prepared application of the tolerance factor is relatively challenging. A
from both stoichiometric and nonstoichiometric precursor conventional concept concerning ionic radii assumes that the
solutions. The average charge PL lifetime is relatively longer at ion is a solid sphere, which is reasonable in the case of
the grain interiors than at the GBs. The relatively short PL inorganic oxide perovskite materials with high bond ionicity.
lifetime was found to be due to trap-assisted nonradiative Unfortunately, the organic cations in perovskite are not
recombination, which is dominant at the GBs.124 By removing spherically symmetric, and they form hydrogen bonds with
these nonradiative recombination pathways to obtain more the BX6 octahedra. Furthermore, typical perovskite materials
uniform brightness with high emissivity across all the grains based on the large-sized iodide would have lower bond ionicity
and GBs, efficiencies close to the theoretical efficiency of PSCs than conventional oxide perovskites, making it challenging to
can be achieved. Although many investigations have estimate the effective ionic radius of the constitutive ions. In
demonstrated that GBs could enhance the photovoltaic fact, we noted that in published studies (including our study),
performances by acting as charge-carrier dissociation sites, different ionic radii of organic cations were utilized to estimate
thus leading to an increase in the photocurrent density, the the tolerance factor. Consequently, intensive efforts have been
passivation of GBs, which can prevent negative behaviors such devoted to revising the ionic radii to obtain a more realistic
as nonradiative recombination, has been conducted by many tolerance factor.125−127 In this review, we adopt the effective
researchers. ionic radii of organic cations estimated by Kieslich et al.125
3.2. Materials Chemistry They hypothesized the rotational freedom of the organic
cations around the center of mass to apply the rigid sphere
As briefly mentioned in section 3.1, the structural formability model, which appears to be reasonable based on the
of the 3D ABX3 perovskite can be estimated by considering the experimental observation. The effective ionic radius is
tolerance factor. The tolerance factor defines the geometrical estimated by the following equation:
criteria for the formation of an ideal cubic perovskite structure.
From Figure 12a, the hypothetical cubic crystal structure rAeff = rmass + rion (3)
displays the lengths of all the yellow lines as identical and equal
where rmass is the distance between the center of mass of the
to 1 of the length of the blue line. With the assumption that molecule and the atom with the largest distance to the center
2
the distance between two ions is the same as the sum of the of mass, excluding the hydrogen atom; rion is the corresponding
ionic radii, we can derive an expression for the tolerance factor ionic radius of this atom. For inorganic ions, we utilized the
(t), as follows: Shannon effective radii for the discussion.128 The ionic radii of
common ions are summarized in Table 3.
rA + rX The optoelectronic properties of semiconductor materials
t=
2 (rB + rX) (2) are generally dependent on the electronic structure of the band
edges. The band edge state of ABX3 perovskite materials
where rA, rB, and rX are the ionic radii of A, B, and X ions, mainly originates from the orbitals of B and X ions. For the
respectively. In principle, when the tolerance factor is 1, the APbI3 perovskite, as mentioned previously, the CB edge of the
ideal cubic perovskite crystal structure is formed. Empirically, perovskite material mainly originates from a Pb 6p orbital with
material compositions with tolerance factors ranging from 0.8 negligible coupling with I, whereas the VB edge of these
to 1.0 were found to be able to crystallize into perovskite- materials originates from a strong antibonding coupling
structured materials at room temperature.125 Below 0.80, other between a fully occupied 6s orbital of Pb2+ and a I 5p orbital
structures, such as ilmenite-type (FeTiO3), are stable, whereas (Figure 12b,c).72 The A-site cations were found to contribute
at tolerance factors greater than 1, hexagonal structures with to the deep energy levels in the CB and VB; thus, they do not
face-sharing octahedra are formed. Although the tolerance directly contribute to the optoelectronic properties of the
factor only considers the geometrical criteria of the materials perovskite materials. Different from conventional semiconduc-
M https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Table 3. Ionic Radii of Commonly Used Ions in ABX3 PbI6 octahedra is relatively weak; therefore, it undergoes
Halide Perovskites125,128 degradation at relatively low temperatures (<200 °C), which
can cause the thermal degradation of the materials and devices.
radius B radius X radius
A cations (pm) cations (pm) anions (pm) Therefore, compositional engineering and, thus, modifica-
tion of the chemical bonding nature critically impact not only
ammonium [NH4]+ 146 Pb2+ 119 Cl− 181
the optoelectronic properties of the materials and the PCE of
methylammonium 217 Sn2+ 69 Br− 196
[(CH3)NH3]+ the devices but also their thermodynamic and environmental
azetidinium 250 I− 220 stability. In this section, we review the critical roles of each
[(CH2)3NH2]+ constitutive ion in this regard and summarize the successful
formamidinium 253 strategies developed to date.
[NH2(CH)NH2]+
3.2.1. A Site(Cation)-Modified Perovskites. Although
dimethylammonium 272
[(CH3)2NH2]+ the A-site cation does not directly contribute to the band edge
ethylammonium 274 states of perovskite materials, it serves as a geometrical spacer
[(C2H5)NH3]+ for the corner-sharing PbI6 octahedra in a 3D perovskite to
guanidinium [C(NH2)3]+ 278 prevent distortion. Therefore, depending on the size and
tetramethylammonium 292 symmetry of the A cation, the bond distance and angle
[(CH3)4N]+ between the Pb2+ and halide ions can be changed to ultimately
Cs+ 167 affect the distribution of the band edge state.129 In fact, the
Rb+ 152 bandgap of APbI3 perovskite was found to vary from ∼1.7 to
K+ 137 1.5 eV depending on the size of the A cation. Nevertheless,
compared to the cases of the B- and X-site ions, the
tors, such as cadmium chalcogenides, whose bandgap is formed compositional engineering of the A cation does not
between the bonding (σ) and antibonding (σ*) orbitals, the significantly affect the optoelectronic properties of the
bandgap of perovskite materials is formed between the perovskite materials; this provides an opportunity to fine-
antibonding states; therefore, the majority of the structural tune the crystal structure of the materials.
defect states are formed inside the energy band, or they only 3.2.1.1. MAPbI3. Historically, the MA-based perovskite
constitute the shallow trap states, which do not significantly material was first explored for photovoltaic applications. With
deteriorate the charge-carrier transport properties. This defect- an ionic radius of 217 pm and a corresponding tolerance factor
tolerant property of perovskite materials was found to promote of 0.91 (Tables 3 and 4),125 it crystallized as a tetragonal
the optoelectronic properties of low-temperature-solution- perovskite at room temperature, whose octahedra were slightly
derived perovskite thin films compared to conventional highly distorted from the cubic structure (Figure 13a,c).46 At elevated
crystalline inorganic semiconductors obtained by delicate high- temperatures (54−57 °C), the MA-based perovskite material
temperature vacuum processes. was found to undergo a phase transition to a cubic structure.
The chemical bonding nature of perovskite materials is The bandgap of the single-crystal MAPbI3 was measured to be
found to be considerably more ionic than that of conventional 1.51 eV. The effective masses of the charge carriers in MAPbI3
semiconductors. This characteristic ionic bonding property have been obtained either using first-principles calculations of
enables the easy dissolution of perovskite materials in a variety the band curvature or by magneto-absorption measurements,
of polar aprotic solvents for further processing via solution- where the low effective masses for both electrons (me*) and
based coating methods at low temperatures. However, the holes (mh*) (around 0.1−0.2 m0) were correlated with the
highly polar ionic nature of the bonding, in turn, causes it to be high charge-carrier mobility (1−70 cm2/(V s) for polycrystal-
easily degraded by moisture, oxygen, and polar solvents, line films).23,35,130 The organic MA cations in the octahedral
resulting in relatively low environmental stability of the cage were found to reorient dynamically within a very short
materials compared to conventional inorganic semiconductors. time. At room temperature, dipolar MA cations reorient
The low stability of the perovskite materials in liquid between the faces, corners, or edges of the pseudocubic lattice
electrolytes posed a serious hindrance to further development cages of corner-sharing octahedra with a residence time of ∼14
until the introduction of solid-state PSCs.4 Furthermore, the ps.131 The dynamic reorientation of the organic cation is
secondary hydrogen bonding between the organic A cation and suggested to stabilize the energetic charge carriers by the

Table 4. Crystal Structures, Along with Optoelectronic and Physical Properties of MAPbI3, FAPbI3, and CsPbI3 Perovskites

mobility diffusion length


effective mass129 (cm2/V s)33 [μm]33
ionic activation
radius of energy for
the A the first cation
cation tolerance crystal structure bandgap mass loss reorientation
[pm]125,128 factor (300 K) (eV) e h e h e h (kJ/mol)143 rate (ps)142
MAPbI3 217 0.91 tetragonal 1.55 0.19me 0.25mh 2−40 10−44 2−8 1.8−12 93 ± 8 108 ± 18
(cubic, (300 K)
>330 K)
FAPbI3 253 0.99 hexagonal/cubic 1.48 0.18me 0.23mh 35 6.6 115 ± 3 8.7 ± 0.5
(metastable)
CsPbI3 167 0.84 orthorhombic 1.73 650 ± 90
cubic
(metastable)

N https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 13. Crystal structure of (a) tetragonal MAPbI3 and (b) cubic FAPbI3 perovskites. (c) molecular structure and ionic radii of the MA+ and
FA+ ions. (d) DOS and (e) absorption coefficients of the MAPbI3 and FAPbI3 perovskites. Reproduced with permission from ref 129. Copyright
2014 American Chemical Society. Reproduced with permission from ref 7. Copyright, 2014 John Wiley and Sons.

formation of a large polaron, which, in turn, enhances the cubic FAPbI3 film tends to undergo an accelerated phase
lifetime of hot carriers as well as band edge carriers.132,133 transition to the nonperovskite phase in humid environments.
However, this dynamic disorder of the A cation highlights the Nevertheless, the cubic FAPbI3 phase film was found to have a
relatively weak bond between the A cation and PbX6 relatively lower bandgap of 1.48 eV compared to that of the
octahedra. Consequently, the degradation of MAPbI3 under MAPbI3 thin film (1.57 eV), while the electronic band
illumination or at elevated temperatures is typically triggered structure and, thus, the absorption coefficient are comparable
by the volatilization of the MA cation. (Figures 13d,e).136 Furthermore, the charge-carrier lifetime
3.2.1.2. FAPbI3. Crystal growth and defect engineering, and diffusion length in the FAPbI3 thin film were found to be
which will be discussed in a later section, have enabled the much longer than those in the MAPbI3 film. Through solid-
formation of high-quality MAPbI3 films to achieve PCEs state NMR measurements, the reorientation rate of the FA
exceeding 20%.10 However, the high PCE of MAPbI3 PSCs is cation in the lattice was found to be much higher (8.7 ± 0.5
limited by the relatively large bandgap and low charge-carrier ps) than that of the MA cation (108 ± 18 ps), enabling a
mobility, which limit the photocurrent and photovoltage of the superior charge-carrier stabilization capability.141,142 However,
devices. After a few years of study, an FA cation-based FAPbI3 the low phase stability of the FAPbI3 perovskite limits the
perovskite material was proposed as an alternative to overcome further enhancement of the PCE in devices based on pure
the drawbacks of the MAPbI3 perovskite. FAPbI3 crystallizes FAPbI3 perovskite films (PCEs exceeding 19% were rarely
into a nonperovskite hexagonal phase at room temperature achieved).
with a bandgap of 2.43 eV.134 The hexagonal phase is 3.2.1.3. CsPbI3. As a replacement for the volatile organic A
converted to a pure cubic phase upon annealing at temper- cations, the pure inorganic CsPbI3 perovskite has also been
atures above 150 °C, and the cubic perovskite phase is explored. Different from organic cations exhibiting weak
maintained after cooling to room temperature (Figure secondary hydrogen bonding with the perovskite lattice, Cs
13b).7,135−137 The spontaneous crystallization of the non- forms a strong primary chemical bond with the perovskite
perovskite hexagonal phase at room temperature is probably lattice. The strong chemical bond between the Cs cation and
related to the larger ionic radius of the FA cation (253 pm, perovskite lattice is supported by the considerably high
tolerance factor = 0.99) compared to that of the MA cation activation energy required for the initiation of the thermal
(Figure 13c, Tables 3 and 4),125 resulting in intrinsic lattice degradation (650 ± 90 kJ/mol) of CsPbI3.143 However, the
strain, which induces the distortion of the corner-sharing relatively small ionic radius of the Cs cation (167 pm) results
octahedra constituting the cubic perovskite lattice to the face- in a tolerance factor of 0.81, effectuating the low thermody-
sharing octahedra. Although the cubic perovskite phase was namic stability of the perovskite phase at room temperature. In
formed upon annealing at relatively high temperatures, the fact, the CsPbI3 film formed by solution processes was found
phase was found to be metastable at room temperature with an to exist in a nonperovskite orthorhombic phase with a wide
activation energy barrier of approximately 0.6 eV for reverse- bandgap (2.82 eV) at room temperature; this phase trans-
phase conversion to the hexagonal phase.138 Owing to the formed into a cubic perovskite phase at temperatures above
relatively high potential barrier, the as-formed cubic FAPbI3 300 °C.144 However, the cubic CsPbI3 film is very unstable
film was stable under inert atmosphere conditions. However, under ambient conditions.145 Consequently, it tends to
the reverse-phase conversion was found to be significantly undergo a phase transition to the orthorhombic phase after a
accelerated upon exposure to moisture.139 Because of the few minutes. The thermodynamic stability of the cubic phase
highly polar ionic bonding, water vapor can attack the crystal was found to be improved by additive engineering. Eperon et
surface of the perovskite phase and induce a significant number al. found that the addition of hydroiodic acid (HI) into the
of surface defects, which reduce the energy barrier for precursor solution could facilitate the formation of the cubic
nucleation of the nonperovskite phase.140 Consequently, the perovskite phase at 100 °C with a bandgap of 1.73 eV.144 The
O https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 14. (a) Absorbance at 630 nm and (b) X-ray diffraction (XRD) patterns of the FA1−xCsxPbI3 films coated on glass. The film was dried in a
vacuum without heat treatment. Insets of (a) show the photographs of the films depending on the Cs content. (c) Differential scanning calorimetry
(DSC) curves for x = 0 and 0.10 in FA1−xCsxPbI3. Reproduced with permission from ref 139. Copyright 2014 John Wiley and Sons.

addition of HI into the precursor solution was found to induce degradation of the photo and thermal stabilities of the film.
the formation of relatively small grains with a slightly strained The FA cation possesses an additional proton (4) compared to
lattice, which enables the stabilization of the cubic phase at the MA cation (3), which enables the formation of enhanced
lower temperatures (approximately 100 °C). The enhanced hydrogen bonding with the perovskite lattice,129 and the
thermodynamic stability with reduced crystal size was also resonance characteristics of the C−N bonds in the FA cation
observed for CsPbI3 quantum dots, which enabled the further stabilize the proton to avoid the generation of HI.139 In
achievement of stable CsPbI3 quantum dot solar cells with fact, a relatively high activation energy for the initiation of the
efficiencies exceeding 13%.146,147 The surface and strain thermal degradation was observed for FAPbI3 (115 ± 3 kJ/
energies were found to be the origin of the enhanced mol) compared to that for MAPbI3 (93 ± 8 kJ/mol).143
thermodynamic stability with the reduced crystal size. A However, the partial replacement of the FA cation with the
similar approach was also applied to fabricate stable FAPbI3 MA cation led to the degradation of the beneficial photo and
quantum dot solar cells.148 Recently, the development of thermal stabilities of the FAPbI3 perovskite films. For example,
various additives has enabled noticeable improvement in the the FA0.5MA0.5PbI3 film undergoes much-accelerated degrada-
PCE. Both zwitterion and 2D perovskite additives induce small tion (complete bleaching after about 8 h of illumination) under
grains to stabilize the cubic phase, resulting in efficiencies 1 sun illumination compared to the FAPbI3 film (complete
higher than 11%.149,150 Furthermore, the passivation of the bleaching after about 16 h).139
crystal surface by organic-halide treatment significantly This problem was overcome by the incorporation of the
improves the optoelectronic properties of CsPbI3 films, relatively small Cs cation instead of the MA cation.128 Owing
which results in efficiencies as high as 18%.151,152 However, to the strong primary chemical bond with the perovskite
the PCE of CsPbI3 PSCs is still limited due to its relatively lattice, the incorporation of the Cs cation does not degrade the
high bandgap of ∼1.7 eV. It can possibly be applied to the top stability of the FAPbI3 films. Lee et al. first incorporated the Cs
subcell of tandem solar cells with appropriate bandgap cation in FAPbI3 films and observed the formation of the cubic
engineering. perovskite phase at room temperature.139 By replacing 10% of
3.2.1.4. Mixed A-Cation Perovskites. Because the low the FA cation with the Cs cation (FA0.9Cs0.1PbI3), seeds of the
environmental stability of FAPbI3 is attributed to the overly cubic FAPbI3 perovskite phase were formed at room
large FA cation, which needs to be fitted into the temperature,139 promoting the formation of the cubic
cuboctahedral site, the partial replacement of the FA cation perovskite film at relatively low temperatures.156 The
with the relatively small MA cation was found to stabilize the contraction of the unit cell volume with the addition of the
cubic perovskite phase at room temperature. 153 The Cs cation indicates the successful incorporation of the cation
incorporation of the MA cation (<50 mol %) enables the into the lattice, resulting in the effective tuning of the tolerance
formation of the cubic perovskite phase at room temperature factor, which would facilitate the formation of the cubic
and significantly enhances the PL lifetime of the perovskite perovskite phase at room temperature (Figure 14a,b).139,157
films, resulting in improved photovoltaic performance.153,154 Consequently, the endothermic peak observed for pure
The significantly enhanced PL lifetime might be related to the FAPbI3, which corresponded to the phase transition at 106.5
enhanced crystallinity of the films, as a consequence of the °C, disappeared upon addition of the 10 mol % Cs cation
released intrinsic strain, enabling long-range ordering of the (Figure 14c). Similar to the FA/MA mixed-cation perovskite,
crystals.155 Although the incorporation of the MA cation enhanced crystallinity was also observed with the substituted
enhances the phase stability of the perovskite film, it causes the cation, which resulted in reduced defect density and enhanced
P https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

photovoltaic performance. Contrary to the FA/MA case, the Additionally, DMA enhanced the moisture stability of the film
incorporation of the Cs cation was found to also be beneficial by hindering the penetration of water molecules toward the
for the environmental stability of the films. Enhanced stability perovskite lattice, and impeded the migration of ionic species
under moisture and illumination was achieved with the Cs in the film. Ferdani et al. investigated the ion transport
cation, which contributed to enhancing the stability and properties in a variety of mixed-cation systems. They found
reproducibility of the solar cells.136,139,157,158 that the activation energy for ion transport in a mixed-cation
With the development of alternative cations, complex system increases due to the modified bonding nature and steric
multication systems have also been explored to enhance the inhibition of ion transport.165 In fact, with the incorporation of
performance and stability of the films and devices. In addition 5 mol % of a variety of different cations (Rb, Cs, azetidinium,
to the improved structural formability achieved by tuning the FA, DMA, and GA) in MAPbI3, the activation energy for ion
tolerance factor, it was suggested that increasing the perovskite transport was increased. Notably, ion transport is critical to the
complexity by the addition of inorganic elements can enhance operational stability of the PSC;166 this indicates that the
the entropy of mixing to stabilize the perovskite phase.158 In mixed-cation system is probably beneficial for enhancing not
fact, Park et al. fabricated MAPbI3 perovskite films with only the environmental and intrinsic stability of the perovskite
enhanced quality using the relatively small Rb cation (ionic materials but also the operational stability of the solar cell
radius = 152 pm),128 which reduced the tolerance factor to a devices.
value further away from 1.0.159 Saliba et al. incorporated Cs Notably, most of the certified record PCEs were achieved
cations into MA/FA systems to fabricate a triple-cation using devices based on mixed MA/FA-based perovskite
system.160 The addition of the Cs cation enhanced the phase materials (Table 1). Nonetheless, there is a recent publication
purity as well as the thermal stability of the perovskite film. As on high certified PCEs over 22% achieved using devices based
a result, a stabilized PCE of over 21% was achieved with an on MA/FA/Cs-based perovskite materials.167 While the
enhanced operational lifetime of the device. They further utilization of organic A cations was reported to be advanta-
extended their approach by adding the Rb cation with an ionic geous in terms of polaron formation,132 it was also recently
radius of 166 Å (the composition of the A-site cation is reported that the utilization of the multication system
RbxCsyMAzFA1−x−y−z). The addition of Rb further reduced the incorporating Cs is beneficial for preventing carrier reflection
defect density and charge-transporting capability of the films, at GBs.168 Furthermore, the A cation composition can affect
resulting in a stabilized PCE of 21.6% with a Voc of 1.24 V, the local strain and, consequently, the crystallization and
corresponding to a potential loss of 0.39 V, considering the stability of the thin films.165,169 Therefore, in-depth and
bandgap of 1.63 eV.161 Furthermore, the device demonstrated systematic studies are required to unravel the effects of the A
excellent thermal stability, where over 95% of the initial PCE cation composition on the photovoltaic performance of the
was maintained after 500 h of operation at 85 °C under films.
continuous light illumination. 3.2.2. B Site(Cation)-Modified Perovskites. Most
Recently, more diverse organic cations are being explored as reported high-efficiency solar cells incorporate Pb-based
alternative constituents for the A site. The utilization of a perovskite materials. Because of the toxicity issue associated
guanidinium (GA) cation was first suggested by De Marco et with Pb, however, intensive efforts have been devoted to the
al.162 GA has a much larger ionic radius of 2.78 Å than MA and development of alternative B-site cations. Thus, far, only Sn-
FA cations (corresponding to a tolerance factor of 1.04);125 based perovskites have demonstrated the potential to replace
therefore, pure GAPbI3 cannot form the perovskite phase. Pb-based perovskites to some degree. Sn has a valence-shell
They incorporated GA cations into the MAPbI3 perovskite to electron configuration identical to that of Pb (4 valence
partially replace MA cations; the GA cations were found to electrons in its s and p orbitals, s2p2); this probably allows the
drastically enhance the PL lifetime of the perovskite film. The Sn-based perovskites to render optoelectronic properties
enhanced charge-carrier lifetime was also observed in the comparable to those of Pb-based perovskites.145 Sn2+ has a
corresponding devices, resulting in the enhancement of the much smaller ionic radius of 69 pm compared with that of Pb2+
photovoltage, which, in turn, improved the device perform- (119 pm), resulting in different band structures and
ance. Through solid-state NMR measurements, it was found corresponding bandgaps of the materials. The bandgap of
that the reorientation rate for the GA cation in the Sn-based iodide perovskites was found to be smaller than that
GA1−xMAxPbI3 system is much higher (≤18 ± 8 ps) than of Pb-based iodide perovskites. For example, the bandgaps of
that for the MA cation (113 ± 25 ps) at 300 K,141,142 which CsSnI3, MASnI3, and FASnI3 were measured to be 1.3−1.4
probably helps to stabilize the charge carriers for enhancing the eV.170−172 Although Sn-based perovskite materials have more
PL lifetime. In addition to the improved tolerance factor of ideal bandgaps for single-junction solar cells than Pb-based
MAPbI3 due to the incorporation of the relatively large GA perovskite materials, the reported PCEs of solar cells based on
cation, the GA cation forms larger number of hydrogen bonds pure Sn-based perovskites is still much lower than those of the
with the PbI6 octahedral cage, which contribute to the reduced devices based on Pb-based perovskite materials (PCE <
formation enthalpy (ΔH), which, in turn, stabilizes the 12%).173 The poor performance of Sn-based PSCs is attributed
perovskite phase.163 Because of the enhanced optoelectronic to the instability of Sn2+ due to its intrinsically low redox
properties and stability, Jodlowski et al. achieved the maximum potential (E0), i.e., Sn2+/Sn4+ (E0 = +0.15 V), compared to that
PCE of 19.21% with a significantly improved operational of Pb2+, i.e., Pb2+/Pb4+ (E0 = +1.67). Consequently, a
lifetime in a solar cell device based on MA0.86GA0.14PbI3. significant proportion of Sn2+ ions is oxidized to Sn4+ during
Similarly, Chen et al. replaced the MA cation with the the fabrication of Sn-based perovskite films, resulting in a high
dimethylammonium (DMA) cation (ionic radius = 272 Å, defect density and undesired p-type doping. Hence, intensive
tolerance factor = 1.03).164 The incorporation of DMA studies have been conducted to stabilize pure Sn-based
allowed the MAPbI3 lattice to expand and, thus, form the perovskites considering that the low stability of Sn2+ hinders
cubic phase, probably due to the increased tolerance factor. the progress of Sn-based PSCs.
Q https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 15. Photographs and UV−vis absorption spectra of MAPb(I1−xBrx)3. (a) UV−vis absorption spectra of the FTO/bl-TiO2/mp-TiO2/
MAPb(I1−xBrx)3/Au cells measured using an integral sphere. (b) Photographs of 3D TiO2/MAPb(I1−xBrx)3 bilayer nanocomposites on FTO glass
substrates. Reproduced with permission from ref 183. Copyright, 2013 American Chemical Society. (c) UV−vis absorbance spectra of the
FAPb(I1−xBrx)3 (top panel) and FA0.83Cs0.17PbI(1−xBrx)3 films (bottom panel). a.u.: arbitrary units. Reproduced with permission from ref 189.
Copyright, 2016 American Association for the Advancement of Science.

Perovskite materials with a partial substitution of Pb with Sn rB


μ=
were found to possess superior stability than pure Sn-based rX (4)
perovskite materials. Mixed Sn−Pb perovskite materials have
low bandgaps around 1.25 eV, which is ideal for the bottom where rB and rX are the ionic radii of the B-site cation and X-
subcell of tandem solar cells. The suppression of Sn4+ and the site anion, respectively. The BX6 composition with μ > 0.41
reduction of the defect density enabled PSCs based on Sn−Pb can form stable BX6 octahedra. The most common X anions
mixed compositions to deliver markedly enhanced PCEs. For are halide anions, including Cl−, Br−, and I−. With 8 valence
example, Zhao et al. achieved a PCE of 18.4% by incorporating electrons, the ionic radii of Cl−, Br−, and I− are 181, 196, and
2.5% Cl into a thick (FASnI3)0.6(MAPbI3)0.4 perovskite 220 pm, respectively. With the ionic radius of Pb2+ (119 pm),
film.174,175 Tong et al. utilized a GuaSCN additive to further the octahedral factors of APbCl3, APbBr3, and APbI3 are
enhance the charge-carrier lifetime in the (FAS- calculated to be 0.66, 0.61, and 0.54, respectively.
nI3)0.6(MAPbI3)0.4 film and achieved a PCE of 20.2%.176 As the ionic radius of the halide anion decreases, the bond
Yang et al. achieved a PCE of 20.3% using a device based on length between the Pb and halide anion decreases;
FA0.5MA0.45Cs0.05Pb0.5Sn0.5I3, attributed to the low trap density consequently, the overlap of the wave function increases to
and long electron diffusion length (2.72 ± 0.15 μm), with the enhance the coupling of atomic orbitals, which, in turn,
addition of 0.03 mol % Cd2+.177 Recently, Lin et al. achieve a increases the bandgap of the material.56 For example, the
PCE as high as 21.1% with a device based on a bandgap of MAPbX3 increases from 1.58 eV (MAPbI3) to 2.28
eV (MAPbBr3) to 2.88 eV (MAPbCl3) in accordance with the
MA0.3FA0.7Pb0.5Sn0.5I3 perovskite by adding metallic Sn into
substituted halide anion.179,180 In terms of photovoltaic
the precursor solution to suppress the formation of Sn4+.178
application, iodide-based perovskite materials are desirable
The highly efficient Sn−Pb PSCs were successfully applied to
because they have the lowest bandgap, which corresponds to a
all-perovskite tandem solar cells, and a PCE of 24.8% was
broad light-absorption range. In terms of structural stability
obtained. Owing to the extended bandgap tunability toward and chemical bonding nature, MAPbI3 is the least stable
the lower end, mixed Sn−Pb-based PSCs have the potential to material because of the following: (i) the tolerance factor of
surpass pure Pb-based PSCs in terms of the PCE, if an effective MAPbI3 (t = 0.91) deviates further away from 1.0 compared to
methodology for crystal growth and defect engineering is those of MAPbBr3 (t = 0.93) and MAPbCl3 (t = 0.94), and (ii)
established. the relatively low electronegativity (χ) of I− (χ = 2.66)
3.2.3. X Site (Anion)-Modified Perovskites. Because X- compared to those of Br− (χ = 2.96) and Cl− (χ = 3.16), which
site anions directly contribute to the band edge states of corresponds to a relatively weak bond with the Pb2+ ion.181,182
perovskite materials, the substitution of the X-site anion can Accordingly, the formation enthalpy of MAPbI3 was higher
significantly alter the optoelectronic properties of perovskite than those of MAPbBr3 and MAPbCl3 as per calorimetric
materials. The tolerance factor determines whether or not the measurements and DFT calculations. The partial substitution
A cation can fit into the cuboctahedral site formed in the of I− with Br− or Cl− was, therefore, utilized to fine-tune either
corner-sharing BX6 octahedra, and another factor needs to be the optoelectronic properties or the chemical stability of the
considered for the structural formability of the BX6 octahedra. perovskite films.
The octahedral factor is devised for this purpose. The radius of 3.2.3.1. Br/I Mixed Perovskites. The partial substitution of
the octahedral hole (rhole) within the six closely packed X ions I− with the relatively small Br− ion to fabricate MAPb-
is approximately 0.41rhole; hence, a B-site cation smaller than (I1−xBrx)3, changed the average Pb−X bond distance and,
0.41rhole cannot coordinate octahedrally without overlapping consequently, the bandgap of the perovskite thin films (Figures
the surrounding anions. From this relationship, the octahedral 15a,b).56 The bandgap of the films increases with respect to x
factor is expressed as according to the following equation:183
R https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Eg (x) = 1.57 + 0.39x + 0.33x 2 (5) the PL spectrum under prolonged illumination was not
observed. Conversely, the PL spectrum of the MAPb(I1−yBry)3
Because of the relatively small ionic radius of Br−, the tolerance films exhibited a red-shift due to the segregation of the I-rich
factor of MAPb(I1−xBrx)3 increases with x; consequently, the phases.189 The perovskite film with an optimized composition
tetragonal crystal structure of MAPbI3 changes to a cubic or (FA0.83Cs0.17Pb(Br0.4I0.6)3 exhibited a charge-carrier mobility
pseudocubic structure when x is greater than 0.2.183 Because of (μ) of 21 cm2 V−1 s−1 and diffusion length of 2.9 μm. Owing to
the substantial increase in the bandgap with the incorporation the high bandgap (1.74 eV), the perovskite film with this
of the Br− ion (approximately 6 mol %), only a small amount composition is suitable for the top subcell in a tandem solar
of Br can enhance the PCEs owing to the improved FF and cell (tandem devices are discussed in more detail in later
Voc. However, a further increase in the Br content will reduce sections).189 Rehman et al. investigated the correlation
the short-circuit current density and, consequently, the PCE of between the composition-dependent crystallinity, optoelec-
the devices due to the relatively high bandgap. The improved tronic properties, and photostability.190 They found that the
FF was attributed to the reduced series resistance of the evolution of the charge-carrier mobility and lifetime in mixed-
perovskite layer, which might be related to the crystallinity of cation and mixed-halide perovskite films closely correlated with
the film. The more ideal tolerance factor possibly improved the the crystallinity of the films, indicating that composition-
long-range ordered growth of the perovskite crystals, thereby dependent crystal growth behaviors have crucial effects on the
promoting the charge-carrier transport properties. Regardless optoelectronic properties of the films. This is probably related
of the PCE drop, the improved ambient stability of the device to the composition-dependent structural stability of perovskite,
with a high Br content was noticeable. The device with a high which will ultimately affect the energetics of crystal nucleation
Br content (x = 0.20, 0.29) was considerably more stable than and growth (discussed in more detail in the later section).
those with low Br contents (x = 0, 0.06) under ambient 3.2.3.2. Cl-Incorporated Perovskites. At the initial stage of
conditions, with the relative humidity exceeding 50%. This is the development, incorporating Cl− ions into an iodide-based
probably related to the high electronegativity of the Br− ion, perovskite materials has been under debate for quite some
which enables a stronger bond with the Pb and MA ions, time. Although the ionic radius of the Cl− ion (181 pm) is
compared to the case with the I− ion. smaller than that of the I− (220 pm) ion, MAPbI3−xClx
Despite the beneficial effects of Br incorporation into the perovskite thin films fabricated from precursor solutions
MAPbI3 perovskite, the phase homogeneity of the mixed halide incorporating Cl− showed an absorption onset comparable
perovskite was found to be degraded under light illumination. with that of MAPbI3 films (i.e., no noticeable change in the
In halide perovskite materials, the formation enthalpies of bandgap of the film was observed). For example, MAPbI3−xClx
point defects were estimated to be much lower (<0.3 eV per perovskite films fabricated from a solution containing 3 mmol
defect) than those of conventional oxide perovskite materials of MAI and 1 mmol of PbCl2 showed an absorption onset
(2.29 eV per defect for BaTiO3).83 In particular, the formation around 800 nm, which is almost identical to that of MAPbI3
enthalpy of halide vacancies was calculated to be as low as 0.08 films.2 Despite the identical absorption properties, however,
eV per defect (iodide vacancy in MAPbI3, VI•), indicating the the significant difference in the electrical properties of the films
presence of a substantially high density of halide vacancies in indicated that Cl− plays a crucial role in the determination of
the film (estimated to be 2 × 10 20 cm −3 at room the charge-carrier behaviors. For instance, both the electron
temperature).83 Furthermore, these defects tended to migrate and hole diffusion lengths in the MAPbI3−xClx film (L = 1069
inside the lattice, and the activation energy for the migration ± 204 nm for electrons, and L = 1213 ± 243 nm for holes)
was 0.58 eV for VI•,184 which further reduces under light were almost 10 times those in the MAPbI3 films (L = 129 ± 41
illumination.185 Because of active halide migration in mixed nm for electrons, and L = 105 ± 32 nm for holes).85
halide perovskites, the segregation of different crystalline Consequently, the roles of the Cl− ion in iodide-based
phases was observed under illumination, which hindered perovskite films have been intensively investigated.
effective charge-collection efficiency and operational stability of The added Cl− ions were found to influence the
the devices.186−188 crystallization of the perovskite films. During the formation
The phase segregation in the MA-based mixed halide of the MAPbI3 perovskite, the Cl− ion in the solution forms
perovskite was mitigated by replacing the MA cation with an intermediate MAPbI3‑xClx phases, which are finally converted
FA or Cs cation. As discussed in the previous section, FA and to MAPbI3 via the volatilization of MACl from the film. The
Cs cations have relatively strong interactions with the formation of the intermediate phase promotes the growth of
surrounding PbX6 octahedra cage; therefore, they can be large grains with relatively high crystallinity, thereby enhancing
expected to suppress the migration of halide ions.143 the optoelectronic quality of the films.191,192 de Quilettes et al.
Furthermore, regarding the tolerance factor, the incorporation observed the correlation between a high Cl content and a high
of the relatively small Cs cation would compensate for the PL emission at a local site (few micrometers scale).122 They
increased tolerance factor due to Br incorporation (t = 0.99 for suggested that Cl might be present at the surface, interstitial
FAPbI3, and t = 1.01 for FAPbBr3), which will help in site, or crystal lattice after aiding the crystallization of the
maintaining the structural stability of the cubic perovskite perovskite film. It was also suggested that the residual Cl− in
phase. Indeed, the incorporation of the Cs cation (17 mol %) the film (∼2 wt % or less) has beneficial effects on the
enables diverse Br incorporation without the formation of a optoelectronic properties of the film. From first-principles
secondary phase. Consequently, a continuous blue-shift of the calculations, it was suggested that doping lead iodide
absorption onset was observed with the incorporation of Cs in perovskites with the Cl atom inherently reduces the
FAPb(I1−xBrx)3 (Figure 15c).189 Owing to the stronger recombination velocity of the photoexcited charge carriers in
interaction of FA and Cs with the PbX6 octahedra compared the bulk and GB regions by reducing the electron−phonon
to the case with MA, the FA1−xCsxPb(I1−yBry)3 perovskite thin coupling and lifetime of quantum coherence for the charge
films did not exhibit phase segregation; therefore, the shift in recombination processes.103,104
S https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 16. (a) Schematics of the device incorporating a polycrystalline 3D perovskite film with 2D perovskite at the GBs. Reproduced with
permission from ref 201. Copyright, 2018 Springer Nature. (b) Proposed electronic band alignment between 2D and 3D perovskites. (c) Current
density−voltage (J−V) curves of the devices without (control) and with 2D perovskite (target). Stability measurement of (d) perovskite films and
(f) devices. Reproduced with permission from ref 201. Copyright, 2018 Springer Nature.

3.2.4. Low-Dimensional Perovskites. If the size of the Owing to the improved charge-carrier lifetime, Wang et al.
A-site cation becomes larger than the volume of the reported an improved stabilized PCE of 19.5% obtained with a
cuboctahedral site formed by the corner-sharing PbX6 network, butylammonium (BA)-based 2D perovskite compared to that
the PbX6 octahedra would not be able to maintain their corner- achieved with the pure mixed-cation-halide perovskite
sharing geometry of the ABX3 3D perovskite, initiating the (FA0.83Cs0.17PbI2.4Br0.6) (17.5%).203 Lee et al. utilized the
formation of a layered 2D A′2BX4 structure. These crystalline low bandgap FA0.98Cs0.02PbI3 perovskite incorporated with
materials are classified under the 2D perovskite family. 2D 1.67 mol % of 2D PEA2PbI4 perovskite and achieved a
perovskite materials typically incorporate large organic cations stabilized PCE of 20.64% compared to 14.91% achieved with
with long alkyl chains or aromatic rings; hence, they are the device based on a pure 3D perovskite (Figure 16c,d).201
relatively less hygroscopic compared to 3D perovskite With the improvement in the PCE, the added 2D perovskite
materials.193,194 However, the bulky organic cations form was found to also be beneficial to the environmental and
potential barriers to the charge-carrier transport inorganic operational stabilities of the film and device (Figure 16e,f).
network; consequently, their charge-carrier transport proper- The defective GB region is typically more vulnerable to
ties are inferior to those of 3D perovskite materials. Quasi-2D degradation by moisture due to its relatively high surface
perovskite materials have been utilized as light-absorbing layers energy. The relatively hydrophobic 2D perovskite can protect
in an attempt to exploit their excellent stability. Upon mixing the GBs from moisture and, thus, improve the environmental
2D and 3D perovskite materials, the A′2An−1BnX3n+1 quasi-2D stability of the perovskite film.201 Furthermore, the defective
perovskite is formed, where the n layers of the 3D perovskite GBs act as a major pathway for ion migration, which is
are separated by a bilayer of bulky A′ cations.195 Although the regarded as the main factor causing the degradation of the
quasi-2D perovskite materials exhibited outstanding environ- solar cells under operational conditions. While the 2D
mental and operational stability, they compromised the PCE of perovskite itself has a relatively high activation energy for ion
the devices due to the presence of insulating organic layers in migration, it probably passivates the defective GBs of the 3D
their bulk crystals. The maximum PCE barely exceed perovskite, which, in turn, enhances the activation energy for
15%.195−200 ion migration.201 Consequently, the device based on the film
A strategy to overcome the poor charge-transport properties with 2D perovskite showed significantly improved operational
of quasi-2D perovskite materials has been developed. It was stability compared to that based on the pure 3D perovskite
found that if the concentration of the 2D perovskite is below a film. Recently, a similar approach was adopted to stabilize
certain threshold (<5 mol%), the added 2D perovskite is CsPbI3 perovskite, where both the PCE and stability of the
preferentially formed at the GB region due to its low surface CsPbI3 solar cell were improved.150 This approach involving
energy and the high activation energy required for the utilization of a 2D perovskite material might be promising
nucleation.201,202 The formation of the 2D perovskite at the for achieving both high efficiency and long-term stability of the
localized GB region does not degrade the bulk charge- devices.
transporting properties; however, the 2D perovskite forms a
type-I band alignment with the grain interior of the 3D 3.3. Processing Engineering
perovskite to repel photogenerated charge carriers from the The performance of PSCs is determined by film qualities such
defective GB region (Figure 16a,b).201,203 Consequently, the as crystallinity and morphology, which are dependent upon the
PL lifetime of the film incorporating the 2D perovskite was film fabrication process. For example, highly efficient PSCs
significantly enhanced compared to that of bare 3D perovskite. possess a dense microstructure without any voids. In this
T https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 17. (a) LaMer model for nucleation and growth of perovskite thin films. Cs is the supersaturation concentration of a precursor solution. (b)
Free energy diagram of nucleation. ΔGs, surface free energy; ΔGν, bulk free energy; ΔG, total free energy, ΔGc; critical free energy; and rc, critical
radius of nucleus. (c) A correction term for heterogeneous (⌀) nucleation as a function of contact angle of the solution on a substrate. (d) A
schematic illustration of the nucleation and growth of perovskite films at each stage. Reproduced with permission from ref 206. Copyright 2019
John Wiley and Sons.

section, we introduce many proposed fabrication processes for nucleation rate using an Arrhenius type equation, expressed in
obtaining high efficiency when fabricating perovskite layers. In eq 6:204,205
ij ΔG yz
= A expjjj− c zzz
typical PSC fabrication, the perovskite thin films are formed by
j kBT z
dN
k {
low-temperature solution processes. The low formation
enthalpy of metal halide perovskite materials and the high dt (6)
ionic nature of their chemical bonding enable the utilization of
cost-effective solution processes at low temperatures. However, where t is time, N is the number of nuclei, A is the pre-
during the solution processes, the crystallization of the exponential factor, kB is the Boltzmann constant, and T is the
perovskite films is completed within few seconds, limiting temperature. The crystal free energy (ΔGc) can be written as a
function of the surface energy γ, molar volume ν, and
effective control over the crystallinity and orientation of the
supersaturation of solution S, which yields eq 7:204,205
ij 16πγ 3ν 2 yzz
film. In solution processes, intermolecular interactions between

= A expjjj− 3 3 z
j 3k T (ln S)2 zz
the perovskite precursor and solvent molecules, as well as dN
k {
fabrication protocols, were found to have critical influences on
dt (7)
the crystallization kinetics of the perovskite films. B

To understand the crystallization processes of the perovskite When a substrate is used in device fabrication, the interface
films, the La Mer mechanism can be applied. In Figure 17, between the precursor solution and substrate can act as a
three different stages of the crystallization processes of heterogeneous nucleation site. Heterogeneous nucleation can
perovskite films are demonstrated according to the La Mer be described by introducing the correction term ⌀,204,205
mechanism. In stage I, the perovskite precursors exist as ions
and molecules in the precursor solution until the concentration ΔGchetero = ⌀ΔGchomo (8)
of the solution increases to a supersaturation level (Cs) owing
to the evaporation of the solvents during the spin coating and (2 + cos θ )(1 − cos θ )2
⌀=
heat treatment of the film. As the concentration reaches Cs, in 4 (9)
stage II, perovskite crystal nuclei start to form and grow with a
supply of solute owing to a diffusion process. The nucleation where θ is the contact angle of the solution on the substrate. As
continues until the concentration of the solution decreases shown in Figure 17c, ⌀ significantly varies depending on θ;
hence, the nucleation kinetics largely depend on the substrate
below Cs, as the consumption of the solutes becomes faster
used for film fabrication. For example, considering the good
than the evaporation of the solutes in stage III. In stage III, wettability of the perovskite solution on typical metal oxide
nucleation does not occur, and only the growth of the formed transporting layers for an n-i-p structured device (e.g., θ = 15−
nuclei progresses until the precursor solution is depleted. 20° for SnO2), ΔGc is expected to be significantly lower at the
The crystallization of perovskite thin films can be modulated interface between the solution and substrate. Therefore,
by tuning the rate of nucleation and growth. According to the nucleation is likely to dominantly occur at the interface
classical theory, the rate of homogeneous nucleation is between the solution and substrate (the situation can change
described as a function of the critical free energy (ΔGc in with different substrate materials and processes).
Figure 17b), representing the free energy required for the During nucleation, growth occurs because of the diffusion of
formed nuclei to be stable without being redissolved in the the precursor molecules and subsequent surface reaction. If the
precursor solution. This critical free energy is defined as the surface reaction is the rate-determining process, the growth
activation energy for nucleation, which is used to describe the rate can be expressed using the following equation,
U https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 18. Schematic diagrams displaying procedure for one-step spin-coating of perovskite thin films using (a) conventional antisolvents and (b)
anisole. Reproduced with permission from ref 212. Copyright 2018 John Wiley and Sons. (c) Photograph of the 10 cm × 10 cm perovskite solar
module. Reproduced with permission from ref 213. Copyright 2019 American Chemical Society. (d) J−V curves of 0.14 and 1.08 cm2 PSC devices
fabricated by anisole as an antisolvent. Reproduced with permission from ref 212. Copyright 2018 John Wiley and Sons.

dr nucleation of perovskite or preperovskite materials. The as-


= kv(C b − Cr)
dt (10) dried films are then annealed at a low temperature in the range
of 100−150 °C for controlling the nucleation and growth of
where r is the radius of the particle, k is the surface reaction the perovskite phase.
rate constant, and Cb and Cr are the concentrations of the At an early stage, the conventional one-step process
solution at the bulk and at the surface of the particle, presented some critical issues that prevented the attainment
respectively. If the diffusion of precursor molecules is the rate- of high efficiency, such as difficulty in fully covering the
limiting process, the growth rate is given by substrate surface with the perovskite layer and irregular grain
dr Dv growth due to the unsuccessful control of nucleation and
= (C b − Cr) growth. For example, in a well-used DMF solvent system,
dt r (11)
irregular rod or needle-like structures with poor coverage on
where D is the diffusion coefficient of the precursor molecules. the substrate have been frequently observed.208 The high
3.3.1. One-Step vs Two-Step Processes. The fabrication boiling point and low vapor pressure of DMF retard
processes of perovskite films can be classified as one-step and evaporation, leading to slow nucleation and rapid growth,
two-step. The one-step process involves the formation of a thereby preventing the generation of a uniform grain structure.
perovskite layer via nucleation and growth from a solution Therefore, research on the one-step process has focused on
containing all precursors that constitute perovskite halide methods to quickly remove the solvent for control over the
compounds. The two-step process indicates a perovskite film rapid and uniform nucleation of perovskite mediates. Rapid
fabrication procedure that yields a perovskite layer through the solvent removal has been attained through various methods,
step-by-step deposition of each precursor layer, followed by a such as solvent engineering, substrate heating, gas blowing, and
thermal interdiffusion process.207 Thus far, the understanding vacuum treatment.
of perovskite formation behavior has been restricted because of The solvent engineering method, also known as the
limited characterization tools, making it difficult to arrive at an antisolvent extraction method, was first reported by Cheng et
exact theory for perovskite formation. al.209 The antisolvent should be miscible with the polar aprotic
3.3.1.1. One-Step Process. In the one-step process, solvent but should not dissolve perovskite ingredients such as
perovskite films are formed via the evaporation of the solvent PbI2 and MAI. They dripped chlorobenzene as an antisolvent,
from undried films that are deposited from a homogeneous which rapidly extracted solvent from the cast film,
solution containing all components. Various methods have consequently facilitating uniform nucleation. Thus far, many
been proposed to improve photovoltaic performance via the antisolvents, including toluene,20 hexane,210 ethyl acetate,211
fabrication of uniform, well-crystallized, and void-free perov- and diethyl ether,21 have been explored. Many high-efficiency
skite layers. The one-step process was first proposed by Kojima solar cells have been fabricated using this antisolvent extraction
et al. in 2009.3 The raw precursors, including PbI2 and method. Moreover, antisolvent extraction can be applied to
methylammonium iodide (MAI), are dissolved in a polar fabricate mixed halide compositions, such as
aprotic solvent such as DMF, DMSO, and GBL, following [CsPbI3]0.05[(FAPbI3)0.85(MAPbBr3)0.15]0.95.212 However, un-
which the solution is directly cast onto the substrate using like MAPbI3, the diffusivity of other ions such as Br is different
spin-coating. During spinning, the solvent evaporates; from that of I, and relatively high, which induces the formation
concurrently, the concentration of the dissolved precursor of pinholes in the mixed halide perovskite. Furthermore, a
reaches the supersaturation state, thereby inducing the molecular interaction between the perovskite precursor ions
V https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 19. (a) Schematic procedure for the gas-blowing spin-coating method and (b) J−V curve of PSC fabricated by gas-blowing. Reproduced
with permission from ref 225. Copyright 2014 Elsevier. (c) Schematic description of cryocontrolled nucleation for perovskite synthesis and (d) J−V
curve of PSC fabricated by cryogenic process. Reproduced with permission from ref 227. Copyright 2018 John Wiley and Sons. (e) Schematic
illustration of nucleation and crystallization procedures during the formation of a perovskite film via vacuum flash-assisted solution processing
(VASP) and (f) J−V curves for the best-performing devices using perovskite films prepared by the conventional process (black) or VASP (red)
method. Reproduced with permission from ref 228. Copyright 2016 American Association for the Advancement of Science.

and solvent molecules in the solution is dependent on the efficiency. For example, the use of a saturated mixture of
composition of the added precursors. Therefore, a change in methylamine in acetonitrile (ACN) can improve the perovskite
the composition of the perovskite results in variations in the microstructure.214 This modified solution allows the formation
precursor solubility and the solvent evaporation rate, inducing of a liquid “melt” of MAPbI3·xCH3NH2, which results in the
differences in the optimized antisolvent timing. The process direct crystallization of the perovskite during spin coating. The
window for the mixed halide perovskite was found to be fairly perovskite layer prepared by this method possessed fewer
narrow for conventional one-step fabrication using a defects as well as a compact microstructure, yielding over 18%
chlorobenzene antisolvent. Recently, anisole has been intro- PCE. The addition of the cyclic urea compound 1,3-dimethyl-
duced as an antisolvent.212 Owing to a low evaporation rate 2-imidazolidinone (DMI) in the precursor solution of MAPbI3
and intermolecular interaction between the anisole and DMF/ induced weak coordination with PbI2 and formed a MAPbI3
DMSO, anisole has an ultrawide processing window for the film with no intermediate phase.215 The resultant perovskite
one-step fabrication of efficient PSCs, for which the small-cell film showed smooth grain structure with an average crystal size
efficiency (0.14 cm2) is 19.76% (Figure 18c).213 Further, the of 1 μm and reduced the number of defect sites. The solar cells
application of anisole enables the fabrication of uniform large- based on these films exhibited a PCE of 17.6%. The
area perovskite films as large as 100 cm2. The module incorporation of Bpy (2,2′-bipyridine) and Tpy (2,2′:6′,2″-
efficiency of the anisole-derived PSC (100 cm2) was 14%.213 terpyridine) additives increased the grain size and crystallinity
Therefore, antisolvent extraction ensures a reproducible of the perovskite films and reduced the defect density, which
fabrication process for highly efficient PSCs as well as large- improved the PCE and operational stability of the device.216
area devices, if one can identify a suitable antisolvent. Compared to the controlled bare film cell, the Bpy-and Tpy-
In the antisolvent method, modification of the solvent and added cells showed higher PCEs of 19.02% and 18.68% in the
the addition of additives have been attempted to improve reverse and forward scans, respectively. This work highlights
W https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 20. Schematic depicting the stages of the reaction in sequential deposition. Dashed arrows indicate mass transfer. (a) Nucleation and
growth of PbI2. (b) Intercalation of MAI and structural reorganization to form CH3NH3PbI3 perovskite. (c) Ostwald ripening, in which the
perovskite from the mesoporous layer is transported to the capping layer. (d) Further Ostwald ripening at longer dipping times in which the
perovskite from the small crystals in the capping layer is transported to larger crystals. (e) Gibbs free energy as a function of the reaction coordinate.
Reproduced with permission from ref 230. Copyright 2018 American Association for the Advancement of Science.

the importance of additive engineering for morphology and high PCE of 20.26%. Because this gas-blowing method does
defect passivation improvement in highly efficient and stable not use toxic antisolvents such as toluene, chlorobenzene, and
PSCs. The addition of a polymeric Lewis base, such as ethylene diethyl ether, this process is an environmentally friendly
carbonate (EC), propylene carbonate (PC), and poly fabrication method. Recently, a cryogenic process was used to
propylene carbonate (PPC), increased the activation energy control the crystallization of perovskite films. This process
for the nucleation and diffusion of the precursor molecules, enabled the decoupling of nucleation and crystallization phases
leading to the formation of perovskite films with high by inhibiting chemical reactions in as-deposited films that were
crystallinity and significantly enlarged grains.217 Moreover, rapidly cooled by immersion in liquid nitrogen (Figure
this polymer enabled the formation of polymer−perovskite 19c).227 The cooling was followed by conventional gas-
composites between grains, which improved the stability of the blowing. This process initiated the formation of a uniform
PSCs. The highest PCE for these cells was 20.06%. Most precursor seed layer, resulting in increased grain size,
methods for achieving high efficiency over 20% have been crystallinity, and low defect density. The best-performing cell
solvent engineering processes, which include a formation step attained a PCE of 21.4% (Figure 19d).
for organic halide−Pb halide−DMSO intermediates and an Another effective method for evaporating the solvent is to
antisolvent extraction step.160,161,218−222 use a vacuum. Li et al.228 devised vacuum flash-assisted
The simplest way to quickly evaporate solvents is to apply solution processing (VASP), which could yield smooth
heat rapidly. Nie et al.223 developed the so-named hot casting perovskite films of high electronic quality over large areas
method. They dripped a 70 °C mixture of PbI2 and MACl (Figure 19e). The flash evaporation of a solvent produced a
solution onto a 180 °C heated substrate with spinning. This burst of perovskite precursor crystals and prevented dewetting,
process yielded exceptionally large millimeter-scale grains with yielding homogeneous films without pinholes. The resultant
reduced trap sites, leading to a high PCE of 18%. The heat PCE was 20.5% owing to good crystallinity with reduced trap
energy enabled efficient evaporation of the solvent in the as- sites. Further, a combination of the gas blowing and vacuum
deposited film, which induced rapid nucleation. Moreover, the process, termed the gas-flow-induced gas pump (GGPM)
heat energy enhanced the diffusivity of ions, which is one of method, was developed. The maximum efficiency of a GGPM-
the reasons this process yielded large-sized grains. As a manufactured cell was 20.44%.229
derivative of the hot casting method, Deng et al. used a hot- 3.3.1.2. Two-Step Process. The two-step process, or
blade coating process and achieved a high efficiency of sequential deposition method, for the fabrication of PSCs
20.3%.224 A gas-blowing or gas-quenching method can involves (i) the deposition of Pb precursors, such as a Pb
evaporate out the solvent from the as-deposited film by halide or Pb-based adduct, on a substrate, (ii) exposure to a
blowing a nonreactive gas, such as N2 or Ar, onto it. This liquid, vapor, or solid of an organic salt, and (iii) a diffusion-
process also promotes a high degree of supersaturation and driven formation of perovskite materials that is usually enabled
consequent rapid and homogeneous nucleation. Huang et by heat treatment. This process is beneficial for the moderate
al.225 employed an Ar gas blowing method and successfully control of reactions between the Pb halide and organic salt,
produced highly uniform perovskite thin films consisting of which induces the crystallization of perovskite materials.
densely packed single-crystalline grains (Figure 19a). They Recently, a detailed path for the two-step process for
achieved a high PCE of 17%, which was a fairly high efficiency MAPbI3 formation has been proposed by Ummadisingu et
in 2014 (Figure 19b). Recently, a facile and scalable air-knife- al.230 As shown in Figure 20, the formation behavior adopts a
assisted meniscus coating technique has been demonstrated.226 sequential path involving (i) the nucleation and growth of
A N2 gas air knife quickly blows out the solvent, leading to a PbI2, (ii) the intercalation of MAI and structural reorganization
X https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 21. (a) Schematics of FAPbI3 perovskite crystallization involving direct intramolecular exchange of DMSO molecules intercalated in PbI2
with formamidinium iodide (FAI). The DMSO molecules are intercalated between edge-sharing [PbI6] octahedral layers, as shown in the field-
emission scanning electron microscopy surface images of a FAPbI3-based layer formed on mp-TiO2 by the (b) intramolecular exchange process
(IEP) and (c) conventional method, (d) J−V curves of best device measured with a 40 ms scanning delay in reverse (from 1.2 to 0 V) and forward
(from 0 to 1.2 V) modes under standard AM 1.5G illumination, and (d) EQE spectra for best device and integrated Jsc. Reproduced with
permission from ref 233. Copyright 2015 American Association for the Advancement of Science.

to form MAPbI3, (iii) Ostwald ripening, in which the The light-harvesting efficiency and charge carrier extraction
perovskite from the mesoporous TiO2 is transported to the were found to strongly depend on the cuboid size. High
capping layer, and (iv) further Ostwald ripening over dipping average PCEs were obtained using MAI concentrations higher
times in which the perovskite from the small crystals in the than 0.038 M and lower than 0.050 M.
capping layer is transported to large crystals. Severe Ostwald Substrate preheating for PbI2 deposition is another strategy
ripening is detrimental to a high PCE. Each step should be for improving the PCE.232 During spin coating, the preheated
carefully controlled to obtain a highly crystalline and dense substrate enabled full coverage of the mesoporous TiO2 film
morphology for attaining high efficiency. This section reviews with PbI2, induced by better infiltration of PbI2 into the
the efforts invested toward obtaining high-efficiency PSCs mesopores. The optimized preheating temperature was 50 °C,
prepared using the two-step process. exhibiting a high PCE of 15.76%. Beyond 50 °C, the
The two-step method was first proposed by Liang et al.231 crystallinity of deposited PbI2 increased, which retarded the
They fabricated MAPbI3 films through the deposition of PbI2 formation reaction of MAPbI3 with MAI. Without preheating,
on a glass or quartz substrate and subsequent dipping into a the average PCE was 10−11%, which demonstrates that
MAI solution. In PSCs, Burschka et al.19 applied this two-step
preheating during PbI2 coating is critical for producing high-
method using nanoporous TiO2 films. In their report, they
quality MAPbI3.
employed slightly modified conditions for the deposition of
Mixed halide materials, such as a mixture of FA and MA
PbI2 as well as the formation reaction for increasing the loading
cation-based perovskite materials, were successfully synthe-
of the perovskite absorber on the mesoporous TiO2 film. The
spinning time for PbI2 casting was short at 5 s, and sized using the two-step process, which extended the optical
subsequently, the PbI2 films were prewetted by dipping them absorption onset further into the red region for the
in 2-propanol for 1−2 s before dipping in a solution of MAI enhancement of light harvesting. The mixed cation
and 2-propanol. This process led to a PCE of 15.0% with good MA0.6FA0.4PbI3 PSC yielded a PCE of 14.9% by virtue of
reproducibility. The PCE was further increased to 17.01% by superior light harvesting and charge-collection efficiency.154 A
adopting the size-controlled cuboidal MAPbI3 proposed by Im highly efficient PSC was realized by exploiting the intra-
et al.8 The cuboid MAPbI3 was prepared using the two-step molecular exchange process (IEP), wherein the PbI2·DMSO
method, and its size was controlled by changing the film was deposited as a first step, followed by the deposition of
concentration of MAI and the exposure time of PbI2 to the FAI and postannealing.233 During annealing, DMSO was
MAI solution before spin-coating. The average cuboid size exchanged with FAI, consequently forming FAPbI3 with large-
decreased with the concentration of MAI and was determined grained dense microstructures and flat surfaces without
to be ∼720 nm for 0.038 M, ∼360 nm for 0.044 M, ∼190 nm residual PbI2 (Figure 21). In addition, mixed FAPbI3/
for 0.050 M, ∼130 nm for 0.057 M, and ∼90 nm for 0.063 M. MAPbBr3 was synthesized by IEP. The resultant cell efficiency
Y https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 22. Schematic illustration of the stamping method comprising two steps: (a) preannealing and (b) bifacial stamping, in which each
perovskite film was prepared at a low temperature with a short annealing time in the preannealing step. Two perovskite films were bifacially
contacted at 100 °C for 9 min without pressure during the stamping step. (b) Perovskite printing technology based on bifacial stamping, in which
(c) the letters “SKKU” were printed by inserting a mask between the MAPbI3 and yellow d-FAPbI3 via the delta−alpha phase transition of FAPbI3,
and (d) a portrait image using an EAPbI3 film was produced by inserting a mask between the MAPbI3 and the yellow EAPbI3 via the reduction of
the bandgap. Reproduced with permission from ref 237. Copyright 2019 Royal Society of Chemistry.

was 20.1%, certified by the standardized method in a PV 2DMSO adduct enabled the broadening of the antisolvent
calibration laboratory. dropping window for obtaining good-quality perovskite films.
3.3.2. Adduct Intermediated Process. Although exten- Another benefit of this adduct compound is the improved
sive efforts have been invested to improve the efficiency of mobility of the grain-boundary motion, which is favorable for
PSCs, the standard deviation of the average PCE is fairly large, forming a dense grain morphology. During the decomposition
which indicates that the reproducibility of high-efficiency PSCs of (MA)2Pb3I8·2DMSO, the evolution of DMSO vapor
is a persisting challenge. The presence of the intermediate induced the local dissolution of MAPbI3, similar to solvent
phase was initially observed by Jeon et al.,20 which was later annealing. As a result, the p-i-n planar structured PSC
found to be a Lewis acid−base adduct composed of perovskite employing the NiO hole conducting layer exhibited a PCE
precursors and Lewis base solvent molecules.21 The Lewis base of 18.4%. Further, this adduct method was adopted to fabricate
adduct of MAI·PbI2·DMSO was formed by spin-casting an FAPbI3 perovskite materials.235 FAPbI3, with large grains, high
equimolar MAI, PbI2, and DMSO solution dissolved in DMF crystallinity, and long carrier lifetime, was obtained using an
solvent.21 The diethyl ether quickly extracted the DMF adduct of PbI2 with sulfur-donor thiourea as the Lewis base.
solvent, which formed the adduct of MAI·PbI2·DMSO. The For FAPbI3, thiourea was more suitable compared to DMSO
adduct film converted into MAPbI3 by simple postannealing. because of the difference in their chemical interaction natures.
The resultant perovskite film showed more uniform morphol- The FAPbI3 prepared by the adduct method exhibited superior
ogy compared with the perovskite film fabricated using photovoltaic performance because of its faster charge transport
conventional PbI2. Chiefly, the adduct method yielded a high and reduced trap sites. The Lewis base NMP made it possible
reproducibility, i.e., a narrow standard deviation with average to fabricate a PSC with better photovoltaic performance via
PCE of 18.3% from 41 cells and the best PCE of 19.7%. forming a stable intermediate FAI·PbI2·NMP adduct.236 DFT
Another adduct compound, MA2Pb3I8·2(DMSO), was calculations showed a higher interaction energy of the NMP
prepared by changing the DMF/DMSO solvent ratio.234 The with the FA cation, supporting the experimental results on the
perovskite film fabricated from the adduct showed a dense formation of a strong coordinative bond with FAI and NMP.
structure with large grains. When no DMSO was added, the Owing to the excellent uniformity and reproducibility of
resultant perovskite film exhibited a high density of pinholes FAPbI3, the resulting PSC attained a high PCE of 20.19%.
and voids at the interface between the perovskite and NiO film Moreover, in this report, criteria for selecting Lewis bases were
substrate. These results suggested that the crystallization of the suggested: (1) a high hydrogen bond-accepting ability and low
perovskite was initiated from the top surface of the film rather hydrogen bond-donating ability, (2) a sterically accessible
than from the substrate. This top-surface crystallization is the electron-donating atom, and (3) matching the hardness and
origin of the poor contact with the NiO surface. For the softness of the Lewis acid and base.
intermediate adduct compound, the perovskite film crystallized Recently, a new fabrication method for perovskite layers
from the bottom surface during annealing, enabling facile based on an adduct compound has been reported. Simple face-
contact with the NiO film. Moreover, the similar densities of to-face contact without pressure, termed the bifacial stamping
(MA)2Pb3I8·2DMSO and MAPbI3 induced less strain during process, enabled the low-temperature phase transition of
the formation of the perovskite, preventing interfacial FAPbI3 and the effective interface modification of MAPbI3,
delamination. The improved stability of the (MA)2Pb3I8· yielding superior photovoltaic performances for both perov-
Z https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Table 5. Efficiencies of PSCs Fabricated by One-Step and Two-Step Processes


ref materials planar/mesoscopic type method PCE (%) year
207 MAPb(I1−xBrx)3 mesoscopic 1-step solvent engineering 16.2 2014
210 MAPbI3 mesoscopic 1-step solvent engineering 17.08 2017
211 MAPbI3 planar 1-step solvent engineering 15.58 2016
212 [CsPbI3]0.05[(FAPbI3)0.85(MAPbBr3)0.15]0.95 mesoscopic 1-step solvent engineering 19.76 2018
213 [CsPbI3]0.05[(FAPbI3)0.85(MAPbBr3)0.15]0.95 mesoscopic 1-step solvent engineering 19.4 2019
214 MAPbI3 planar 1-step solvent engineering 19.0 2017
215 MAPbI3 mesoscopic 1-step solvent engineering 17.6 2018
216 MAPbI3 mesoscopic 1-step solvent engineering 19.02 2019
217 MAPbI3 planar 1-step solvent engineering 20.06 2019
161 RbCsMAFA mesoscopic 1-step solvent engineering 21.6 2016
218 MAPbI3 mesoscopic 1-step solvent engineering 21.2 2017
219 FA1−xMAxPb(I1−yBry)3 mesoscopic 1-step solvent engineering 20.8 2016
220 FA0.81MA0.15Pb(I0.836Br0.15)3 mesoscopic 1-step solvent engineering 20.6 2017
221 (FAPbI3)0.95(MAPbBr3)0.05 mesoscopic 1-step solvent engineering 22.85 2018
222 (FAPbI3)0.85(MAPbBr3)0.15 mesoscopic 1-step solvent engineering 20.3 2017
160 Csx(MA0.17FA0.83)(1‑x)Pb(I0.83Br0.17)3 mesoscopic 1-step solvent engineering 21.1 2016
223 MAPb(I1−xClx)3 planar 1-step heating substrate 18 2015
224 MAPbI3 planar 1-step heating substrate 20.3 2018
225 MAPbI3 planar 1-step gas blowing 17 2014
226 Cs0.05FA0.81MA0.14PbI2.55Br0.45 planar 1-step gas blowing 20.26 2019
227 Cs0.05(MA0.17FA0.83)0.95 Pb(I0.84Br0.16)3 planar 1-step gas blowing 21.4 2018
228 FA0.81MA0.15PbI2.51Br0.45 mesoscopic 1-step vacuum treatment 20.5 2016
229 FA0.1MA0.9PbI3 planar 1-step vacuum treatment 20.44 2017
21 MAPbI3 mesoscopic 1-step adduct intermediate 19.7 2015
234 MAPbI3 planar 1-step adduct intermediate 18.4 2017
236 FAPbI3 mesoscopic 1-step Adduct intermediate 20.19 2018
237 MAPbI3 mesoscopic 1-step Adduct intermediate 20.18 2019
19 MAPbI3 mesoscopic 2-step 15.0 2013
8 MAPbI3 mesoscopic 2-step 17.01 2014
232 MAPbI3 mesoscopic 2-step 15.76 2015
154 MA0.6FA0.4PbI3 mesoscopic 2-step 14.9 2014
233 (FAPbI3)0.95(MAPbBr3)0.05 mesoscopic 2-step 20.1 2015

skites (Figure 22).237 This result was explained in terms of the a structural similarity to organic photovoltaics or thin-film solar
content of the DMSO reservoir in the MAPbI3 adduct film, cells (e.g., CdTe, CIGS, and CZTS). Mesoscopic devices
which induced the phase transition of FAPbI3 and high-quality generally require higher processing temperatures to sinter the
MAPbI3. During the bifacial stamping, a DMSO-mediated ion nanoparticle layer, whereas planar devices can be prepared at
exchange reaction took place, which facilitated ion transport lower temperatures, especially when the device contains carbon
from the Lewis base reservoir in the MAPbI3 adduct film, or organic charge-transporting layers. Therefore, planar devices
consequently resulting in the fast phase transition of FAPbI3 at are preferred when fabricating flexible solar cells with plastic
a low temperature. The adduct-intermediate process is substrates or tandem devices, although the best single-junction
promising for preparing dense perovskite films with large PSC still has a mesoscopic structure. Second, PSCs can be
grains, which is among the requirements for achieving high classified as either n-i-p type or p-i-n type, depending on the
efficiency. direction of charge collection. For n-i-p type devices, electrons
In Table 5, the efficiencies of PSCs fabricated by one-step are collected to the substrate through which the incident light
and two-step processes are summarized. Thus, far, champion enters, whereas for p-i-n type devices, holes are collected
cell has employed a mixed-halide perovskite layer prepared by through the substrate.
the one-step process. 3.4.1. Selective Contact. 3.4.1.1. Inorganic Electron
3.4. Device Physics Transport Layer (ETL). Titanium dioxide (TiO2) has been
In terms of the device geometry, PSCs can be classified in two mostly used as an ETL in PSCs because the first PSC
different ways. First, they can be classified either as a constructed using the current device structure originates from
mesoscopic structure or as a planar structure depending on solid-state DSSCs.1 TiO2 offers several advantages as an ETL,
the existence of the mesoporous nanoparticle layer. Meso- including a wide bandgap to minimize parasitic absorption, an
scopic devices have a structural similarity to DSSCs, but the appropriate conduction band alignment with the perovskite for
mesoporous charge-transporting layer in PSCs is much thinner efficient electron injection, and good electron-transport
than the sensitized solar cells, and the perovskite layer capability. Therefore, many record efficiencies have been
generally fills the pores in the mesoporous nanoparticle layer obtained using TiO2 ETLs. However, TiO2 does present some
and forms an overlayer on top of it. In contrast, planar devices disadvantages as an ETL in PSCs, such as UV-induced
consist of stacked multiple thin layers, such as selective photocatalytic activity, which causes degradation in the device
contacts, light-absorbing layers, and window layers, which have performance, capacitance issues that create I−V hysteresis and
AA https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 23. (a) I−V hysteresis in n-i-p devices constructed using a TiO2 ETL and their relevant capacitance spectra. Reproduced with permission
from ref 246. Copyright 2015 American Chemical Society. (b) Mechanisms of I−V hysteresis. Reproduced with permission from ref 248.
Copyright 2016 American Chemical Society. (c) J−V curves of the n-i-p device constructed using a SnO2 ETL. Reproduced with permission from
ref 254. Copyright 2017 Royal Society of Chemistry.

high processing temperatures (>450 °C). Because of the photostability. SnO2 ETLs have been prepared via various
limited compatibility of their processing, most TiO2 ETLs are synthetic approaches, such as spin-coating of nanopar-
deposited directly onto a TCO substrate and applied in n-i-p ticles213,253 or precursor solutions,254,255 atomic layer deposi-
type devices. It should also be noted that most high-efficiency tion (ALD),256,257 and chemical bath deposition (CBD);257
devices incorporate a mesoporous ETL consisting of TiO2 among the ETLs prepared by these methods, most high-
nanoparticles due to the severe I−V hysteresis observed in performance SnO2 ETLs reported to date have been prepared
planar devices constructed using a thin-film TiO2 ETL. At the at temperatures below 200 °C. The conversion efficiency of
early stage of perovskite research, most TiO2-related research PSCs constructed using SnO2-based ETLs has continuously
studies conducted toward improving the photovoltaic perform- increased using various optimization processes, such as bilayer
ance involved the modification of TiO2, including the formation,258 interfacial cross-linking,259 and EDTA complex-
crystalline phase (i.e., anatase vs rutile),238,239 morphology of ation,260 leading to the highest certified efficiency of >23%.261
its nanoparticles (i.e., size and shape), 238,240,241 and In contrast to TiO2-based ETLs, most highly efficient PSCs
doping.9,242,243 The lesson learned from these early studies constructed using SnO2-based ETLs reported to date have
was that a thin mesoporous TiO2 nanoparticle layer was the been prepared with a planar n-i-p geometry incorporating a
optimal form of a TiO2 ETL; the conversion efficiencies SnO2 thin film because SnO2-based PSCs usually exhibit
obtained for PSCs constructed using a TiO2 ETL have insignificant I−V hysteresis (Figure 23c). The origin of the
continuously increased since then, to >23%.221 However, reduced I−V hysteresis in SnO2 ETLs when compared to their
issues such as UV-instability, I−V hysteresis, and high TiO2-based counterparts remains largely unclear, but fast
processing temperature still need to be resolved. With regard charge extraction due to the deeper conduction band edge and
to the photocatalytic degradation observed under UV low interfacial charge accumulation due to high conductivity
irradiation, Ito et al. introduced a Sb2S3 layer between the (or electron mobility) have been proposed as possible
TiO2 and perovskite layers as a blocking layer, which resulted causes.252 Like other oxide materials, SnO2 is stable under a
in significantly improved stability under visible-light and UV variety of environments, such as oxygen-rich and humid
irradiation.244 Optical modulations such as UV down- conditions, which helps to improve the long-term stability of
conversion can also be applied to avoid photocatalysis,245 the PSC when compared to its organic counterparts. In
but the only way to completely mitigate this issue is to use addition, PSCs constructed using SnO2-based ETLs are very
another ETL material without any photocatalytic activity. I−V stable under visible-light illumination, especially in the
hysteresis is another fundamental issue for TiO2 ETLs. This presence of UV light262 because SnO2, unlike TiO2, does not
hysteresis is mainly ascribed to the capacitance of the TiO2 exhibit any photocatalytic activity. There have been several
layer (Figure 23a).246 The relatively high dielectric constants approaches using an SnO2 ETL in p-i-n type devices, which are
observed for TiO2 materials (∼40 for anatase and ∼100 for much more beneficial toward the stability of the device because
rutile)247 are primarily responsible for the capacitance. Other the inorganic ETL on top of the perovskite layer effectively
potential causes include charge accumulation due to blocks the permeation of water and/or oxygen molecules by
unbalanced charge extraction/transport between the ETL acting as an encapsulation layer. To date, SnO2 ETLs can only
and HTL, and interfacial defects/traps (Figure 23b).248 be deposited on an additional ETL, such as C60263 and
Successful attempts to circumvent the issues related to PCBM,264 due to technical difficulties, but devices constructed
interfacial charge extraction and traps include the incorpo- using top-layer SnO2 ETLs exhibit significantly improved long-
ration of a PCBM for efficient electron extraction,249,250 term stability. Besides the two most popular inorganic ETL
complete crystallization to reduce the interfacial trap materials (i.e., TiO2 and SnO2), various oxide materials with
density,251 and Cl treatment on the TiO2 surface.243 desirable electrical/optical properties (e.g., ZnO, 265
SnO2 ETLs have also been extensively investigated as an Zn2SnO4,266,267 and BaSnO3218) have been investigated as
alternative to conventional TiO2 ETLs because SnO2 exhibits potential alternatives, but their ETL performances have yet to
good conduction band alignment with perovskites, high be improved to compete with those of TiO2 and SnO2.
electron mobility, a wide bandgap, high transmittance, good 3.4.1.2. Organic ETLs. Fullerene (C60) and its derivatives,
stability, and offers easy processing.252 More importantly, SnO2 such as PCBM, are the most popular organic ETLs, especially
possesses some critical advantages over TiO2, such as lower for p-i-n type and/or flexible PSCs, due to their appropriate
processing temperatures, less I−V hysteresis, and improved electrical properties and physical softness. PSCs constructed
AB https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 24. (a) J−V curves obtained for highly efficient NiO-based p-i-n perovskite solar cells. Reproduced with permission from ref 269. Copyright
2017 American Chemical Society. (b) Operational stability of NiO-based perovskite solar cells. Reproduced with permission from ref 159.
Copyright 2017 John Wiley and Sons. (c) Operational stability of highly efficient CuSCN-based n-i-p perovskite solar cells. Reproduced with
permission from ref 293. Copyright 2017 American Association for the Advancement of Science. (d) SEM image and (e) J−V curves obtained for
the n-i-p device constructed using a CuGaO2 HTL. Reproduced with permission from ref 301. Copyright 2017 John Wiley and Sons.

from C60 or PCBM exhibit high FF values and negligible I−V For p-i-n devices, on the other hand, poly(3,4-ethyl-
hysteresis due to their high electron mobility and excellent enedioxythiophene)−poly(styrenesulfonate) (PEDOT:PSS)
electron-extracting capabilities. In addition, their processing was initially used as the HTL because it has been used in
conditions using either vacuum deposition or solution organic bulk heterojunction solar cells.277 This polymeric
deposition are well-established because they have been widely material presents many advantages such as an appropriate
used in organic solar cells. In particular, they can be dissolved HOMO energy level (−5.5 eV), good conductivity, and low-
in a nonpolar solvent that is orthogonal to the perovskite layer, temperature processing but also presents the catastrophic
and hence, can be deposited directly on top of the perovskite disadvantage of low stability resulting from its hygroscopic and
layer during the solution-phase process. One of the drawbacks acidic nature.278 Despite these drawbacks, PEDOT:PSS HTLs
of fullerene-based ETLs is that stable noble metals such as Ag are still commonly used in small-bandgap PSCs (∼1.2 eV)
cannot be used due to their inappropriate energy alignment. incorporating Sn ions, particularly with the p-i-n geometry
The energy barrier for electron transfer from PCBM (or C60) designed for tandem solar cells.279,280
to the Ag metal electrode can be successfully reduced by Another widely used polymeric HTL based on poly[bis(4-
introducing interfacial layers including bathocuproine phenyl)(2,4,6-trimethylphenyl)amine] (PTAA) shows higher
(BCP),268 polyethyleneimine ethoxylated (PEIE),269,270 and intrinsic hole mobility (10−3−10−2 cm2/(V s)) and enhanced
stability when compared to other small molecules and
other small-271/macro-molecules272 between PCBM (or C60)
PEDOT:PSS. Interestingly, PTAA HTLs have been applied
and the metal electrode, leading to the significantly improved
to both n-i-p183 and p-i-n281 type devices, resulting in high
FF values observed for p-i-n type PSCs.
conversion efficiencies regardless of doping282 and work very
3.4.1.3. Organic Hole Transport Layer (HTL). Similar to
well in p-i-n type wide-bandgap PSCs.283 Recently, the use of
TiO2 ETLs, the HTL used for the first PSC was also adopted poly(3-hexylthiophene) (P3HT) as a HTL in highly efficient
from solid-state DSSCs. Spiro-OMeTAD, which has been used (22.7%) n-i-p type PSCs without doping has also been
as a hole-transport material in solid-state DSSCs,273 was used reported, where successful defect passivation at the P3HT/
as the first HTL in PSCs1 and is still the most commonly used perovskite interface not only increases the Voc (more than 200
HTL, especially in n-i-p type devices. State-of-the-art spiro- mV) and FF (10%) values but also minimizes I−V
OMeTAD or related small molecules require the addition of hysteresis.284
dopants and/or additives such as lithium bis(trifluoro- 3.4.1.4. Inorganic HTLs. To overcome the instability issues
methane)sulfonimide (Li-TFSI) and 4-tert-butylpyridine presented by organic HTLs, various inorganic p-type semi-
(tBP) due to the low conductivity (∼10−5 mS/cm) and low conducting materials have been investigated as HTLs in PSCs.
hole mobility (10−5−10−4 cm2/(V s)) of pristine spiro- Among them, NiO is one of the most extensively investigated
OMeTAD. However, essential additives such as Li-TFSI are p-type materials because of its large bandgap, high trans-
known to cause significant instability due to their hygroscopic mittance, and deep valence band (Figure 24a,b).285 In
nature.274 Therefore, there have been various approaches addition, it can be easily deposited using various fabrication
toward the development of new dopants with good efficiency/ methods, such as spin-coating,278 electrodeposition,269 sputter-
stability275 and new dopant-free HTLs.276 ing,286 e-beam evaporation,287 and ALD.288 However, NiO
AC https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Table 6. Various Materials Used for Electron/Hole-Selective Contact Layers


classification materials device type process
inorganic ETL TiO2 mesoscopic/planar n-i-p nanoparticle1,221
213,253
SnO2 planar n-i-p nanoparticle, solution,254,255 ALD,256,257 CBD257

organic ETL C60 planar n-i-p/p-i-n evaporation302


PCBM planar p-i-n solution303

inorganic HTL NiO planar p-i-n solution,278 electrodeposition,269 sputtering,286 evaporation,287 ALD288
CuSCN mesoscopic/planar n-i-p/p-i-n solution293
CuOx planar n-i-p/p-i-n solution, evaporation,295 sputtering296
294

delafossite mesoscopic/planar n-i-p/p-i-n nanoparticle299,301

organic HTL spiro-OMeTAD mesoscopic/planar n-i-p solution1


PEDOT:PSS planar p-i-n solution278
PTAA mesoscopic/planar n-i-p/p-i-n solution183,281

HTLs also suffer from some drawbacks, such as low hole higher bandgap energies, such as CuCrO2 (3.0 eV)297 and
conductivity and poor electrical/physical contact with the CuGaO2 (3.6 eV).298 Delafossite-based HTLs generally
perovskite layer, which have been successfully improved upon require doping to improve the hole conductivity and have
doping with monovalent acceptor ions such as Li+289 and been mostly prepared using nanoparticle-based spin-coating
Cu+,290 interfacial treatment with PEDOT:PSS,278 and the deposition at low temperature, which can be used for both p-i-
introduction of nanoscale roughness on the surface of the NiO n and n-i-p devices. The p-i-n and n-i-p devices constructed
HTL.269 It should be noted that NiO HTLs have been mostly using CuCrO2-based HTLs show high conversion efficiencies
used in p-i-n type devices due to the relatively harsh processing of 20.54%299 and 16.25%,300 respectively. It is noteworthy that
conditions of oxide materials. the n-i-p device in which the undoped CuCrO2 HTL was
Copper thiocyanate (CuSCN) is another widely used directly deposited onto the perovskite layer exhibits a high
inorganic HTL material with a deep valence band, high carrier conversion efficiency of 16.25% and significantly improved
mobility, good thermal stability, and high transparency.291 It long-term stability, indicating that the device performance can
readily dissolves in a variety of solvents and does not need a be further improved via an appropriate doping strategy during
postannealing process for crystallization; hence, it has been the synthesis of the nanoparticles. CuGaO2-based HTLs have
used in both p-i-n and n-i-p type devices. However, in p-i-n also exhibited good performance (20.15%) in the p-i-n device
devices, the underlying CuSCN HTL was found to react with with a Zn-doping and bilayered structure, and NiO HTL.298
the perovskite layer during the annealing conducted to The corresponding n-i-p device constructed using an undoped
crystallize the perovskite layer, resulting in the degradation of CuGaO2 HTL directly deposited onto the perovskite layer
the device performance.292 Therefore, higher efficiencies and showed excellent ambient long-term stability and a conversion
improved stability have been reported for n-i-p devices. For efficiency of 18.51% (Figure 24d,e).301 The various contact
example, a high conversion efficiency of 20.4% can be achieved materials discussed in this section are summarized in Table 6.
for n-i-p devices by forming a highly compact and conformal 3.4.2. Interface Engineering. A PSC is typically
CuSCN layer using a dynamic drop-casting method, in which constructed from stacked multiple layers consisting of the
the device showed excellent long-term operational stability perovskite active layer, charge-transporting layers for electrons
over 1000 h even at an elevated temperature of 60 °C, with a and holes, and current-collecting electrodes; consequently,
spacer layer consisting of reduced graphene oxide (rGO) there exist four major interfaces between the layers (i.e.,
between CuSCN and Au (Figure 24c).293 The role of the rGO electrode/ETL/perovskite/HTL/electrode). The properties of
layer was described as an encapsulation layer for the CuSCN/ the interfaces are very important for achieving good device
perovskite layers and a protection layer for the potential- performance because they directly influence charge extraction/
induced reaction that occurs between the CuSCN and Au transport/recombination as well as photon transmission. The
layers. interfaces can be further classified into two simplified groups
Copper oxide (CuOx) has also been investigated as a HTL (i.e., one between the charge-transporting layer and the
material in PSCs due to its deep valence band and high carrier current-collecting electrode and another between the perov-
mobility. CuOx HTLs have been prepared using a variety of skite layer and charge-transporting layer). Some of the
synthetic methods such as solution-phase,294 evaporation,295 representative studies used to improve the solar cell perform-
and sputtering processes.296 However, the relatively low ance by modifying each interface are described below.
bandgap energy of CuOx (1.2 eV for cupric oxide, CuO; 2.1 3.4.2.1. Electrode/ETL and Electrode/HTL Interfaces. The
eV for cuprous oxide, Cu2O) causes significant optical loss, work function of the metallic electrode is an important
especially in p-i-n devices where the incident light should enter parameter governing the charge transfer observed at the
via the HTL. n-i-p devices are free from the issue of optical interface with the charge-transporting layer. For example, metal
loss, but only a limited number of synthetic approaches (e.g., electrodes for collecting electrons generally require a relatively
evaporation) are available due to the degradation of the low work function (e.g., Al, 4.1 eV) that matches the CB (or
perovskite layer during the processing. Therefore, research LUMO level) of the ETL (below CBperovskite, −3.9 eV),
efforts have been invested toward increasing the bandgap whereas the electrodes for hole transfer require a relatively high
energy of CuOx by forming new p-type delafossite oxides with work function (e.g., Ag, 4.2−4.7 eV; Au, 5.3 eV) due to the VB
AD https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 25. (a) Illustration of the dipole in PEIE and (b) work function shift observed using PEIE. Reproduced with permission from ref 304.
Copyright 2012 American Association for the Advancement of Science. Effects of the PEIE layer on the J−V curves obtained for the (c) n-i-p
device9 and (d) p-i-n device (unpublished results). Reproduced with permission from ref 9. Copyright 2014 American Association for the
Advancement of Science. (e) PEDOT-treated NiO HTL. Reproduced with permission from ref 278. Copyright 2015 American Chemical Society.
(f) An SnO2-based n-i-p device constructed using an ionic liquid (ImAcHCl). Reproduced with permission from ref 259. Copyright 2019 John
Wiley and Sons.

energy level (or HOMO level) of the HTLs (above VBperovskite, increase the FF value of p-i-n type devices. Although its
−5.4 eV). In terms of the long-term stability, noble metals with working mechanism has not been fully elucidated, Shibayama
better resistance to chemical reactions are considered as better et al. reported that the thin BCP layer (<5 nm) facilitates
candidates for the current-collecting electrode despite their electron collection by changing the Schottky contact at the
high cost. Noble metals function suitably with most HTLs PCBM/Ag interface into an ohmic contact.268
reported to date, but their work functions are generally too 3.4.2.2. Perovskite/ETL and Perovskite/HTL Interfaces. The
high to collect electrons efficiently from the ETL. A universal interface formed between the perovskite and charge-trans-
method to reduce the work function of the electrodes using porting layers has been mostly investigated toward improving
polymer modifiers has been reported by Zhou and Kippelen et the photovoltaic performance of PSCs regardless of their type
al.304 The work functions of various conducting electrodes can because several major phenomena, including the charge
be significantly reduced by the physical adsorption of an extraction/recombination and degradation of materials, occur
ultrathin layer (1−10 nm) of an amine-containing polymer, at the interface. Most of these studies have focused on oxide-
such as polyethylenimine ethoxylated (PEIE) and branched based ETLs and HTLs because the heterogeneous interface
PEI, on the surface of the electrodes, leading to a significant formed between the perovskite and oxide charge-transporting
increase in the FF of their corresponding organic solar cells. layers generally exhibits poorer physical/electrical contact
The intrinsic molecular dipole moments associated with the properties when compared with their organic counterparts.
neutral amine groups contained in the insulating polymer layer NiO HTLs, for example, often exhibit lower FF values than
along with the charge-transfer characteristics of their organic HTLs due to their inefficient hole extraction
interaction with the conductor surface reduce the work capabilities.269,278,288 Park et al. first demonstrated that the
function of a wide range of conductors (Figure 25a,b).304 surface treatment of an NiO HTL using a dilute solution of
Later, Zhou and Yang et al. adopted this approach to n-i-p type PEDOT:PSS drastically improves hole extraction, which was
PSCs, where PEIE treatment of the ITO substrate resulted in a confirmed by the enhanced PL quenching and significantly
reduced shunt conductance, decreased series resistance, and reduced total series resistance obtained in the presence of an
increased FF, which can be ascribed to the facilitated electron extremely thin interlayer (Figure 25e).278 As a consequence, a
extraction at the ETL/electrode interface (Figure 25c).9 This high FF of 0.72 and conversion efficiency of 15.1% without I−
approach cannot be directly applied to p-i-n devices because V hysteresis can be achieved, which was one of the best
the metal electrodes are deposited on top of the ETL layers. performances observed for NiO-based p-i-n type PSCs at that
Park et al. have shown that the PEIE layer deposited on the time. As mentioned earlier in this chapter, some TiO2 ETLs
surface of the PCBM ETL also reduces the work function of also show sluggish electron transfer from the perovskite layer,
the Ag electrode deposited on top of the PEIE layer and FF of resulting in charge accumulation and significant I−V hysteresis.
the p-i-n type PSC significantly increases upon the addition of As with NiO HTLs, it has been reported that electron transfer
the PEIE layer (Figure 25d).159,269 A thin layer of bath- from the perovskite to the TiO2 ETL can be effectively
ocuproine (BCP) between PCBM and Ag is also known to facilitated upon inserting carbon-based materials.250,305,306 Li
AE https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 26. (a,b) Effect of Cl− on the quality of the TiO2/perovskite interface. The (c) charge extraction/transport, (d) charge recombination, (e)
I−V hysteresis, and (f) J−V curves obtained for the optimum device constructed using the TiO2−Cl ETL. Reproduced with permission from ref
243. Copyright 2017 American Association for the Advancement of Science.

et al. reported that the modification of the TiO2 surface using a interaction between the perovskite and NiO formed at the
triblock fullerene derivative (PCBB-2CN-2C8) improves interface, and the excellent long-term stability was discussed in
charge extraction from the perovskite layer, leading to terms of the greatly suppressed ion migration process.303 Chen
drastically increased Voc and FF values along with reduced et al. introduced an ionic-liquid-based multifunctional chemical
I−V hysteresis.305 Passivation of the deep trap states of TiO2 linker between the perovskite and SnO2 ETL (Figure 25f).259
along with facilitated charge extraction have also been 4-Imidazoleacetic acid hydrochloride (ImAcHCl) provided a
proposed to be responsible for the enhanced device perform- chemical bridge between SnO2 and the perovskite using an
ance, in which the long-term stability increased 4-fold in the ester bond with SnO2 formed via an esterification reaction and
presence of the fullerene-based interlayer. Kim et al. also an electrostatic interaction with the perovskite (i.e., between
reported similar phenomena upon forming a hybrid ETL layer the imidazolium cation in ImAcHCl and an iodide anion in the
consisting of PCBM and ALD-deposited amorphous TiO2 for perovskite). The n-i-p device constructed using the ionic liquid
use in flexible PSCs.250 In the hybrid ETL, the PCBM layer exhibited significantly improved charge extraction and reduced
reduces I−V hysteresis by facilitating charge extraction and charge recombination, leading to a high Voc of 1.14 V and high
reducing interfacial charge accumulation, and the amorphous efficiency of 21% with good long-term stability and negligible
TiO2 layer deposited using plasma-enhanced ALD suppresses I−V hysteresis.
the dark current and shunt conductance of PCBM ETLs. Atomic doping or treatment of the oxide charge-transporting
Graphene and related two-dimensional materials are also often layer was also found to improve the interfacial contact
introduced into the perovskite/TiO2 interface.307 Agresti et al. properties with the perovskite layer, especially for TiO2
have reported that the light intensity dependency of Jsc in n-i-p ETLs.9,304,243 Zhou et al. reported the use of Y-doped TiO2
devices becomes more linear in the presence of an interlayer (Y-TiO2) as an ETL in an n-i-p device, where the Y-TiO2 ETL
consisting of graphene flakes and/or lithium-neutralized showed improved charge-extraction capability when compared
graphene flakes, which indicates the absence of energy barriers to its undoped counterpart due to the slight upward shift in the
and improved charge extraction.284 Most approaches employed Fermi level of the Y-TiO2 layer.9 The increased donor
for interface engineering of the TiO2 ETLs such as fullerene,308 concentration of the Y-TiO2 layer and the reduced depletion
PCBM,309 and graphene quantum dots,310 have also been width at the interface with the perovskite layer was also
applied to SnO2 ETLs, and their effects have been discussed in proposed to be responsible for the lower contact resistance and
a similar manner (i.e., facilitated charge extraction and reduced facilitated charge extraction. Tan et al. reported an interfacial
energy barrier). One recent and new approach toward engineering strategy involving the use of Cl-capped TiO2
improving the performance of PSCs constructed using oxide (TiO2−Cl) colloidal nanocrystals in the ETL, which mitigated
ETLs is to incorporate an ionic liquid.311 Bai et al. introduced interfacial recombination and improved interfacial binding in
1-butyl-3-methylimidazolium tetrafluoroborate (BMIMBF4) low-temperature planar solar cells (Figure 26).243 The band
into a variety of NiO-based p-i-n devices, in which the alignment between TiO 2 −Cl and the perovskite was
improved device performance was ascribed to the improved determined using ultraviolet photoelectron spectroscopy
AF https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 27. (a) Effect of adding K+ ions on the I−V hysteresis of a n-i-p device (up/left). Reproduced with permission from ref.313 Copyright 2018
American Chemical Society. Effect of adding Rb+ ions on (b) J−V curves and (c) conductivity and tDOS of a p-i-n device (up/right). Reproduced
with permission from ref 159. Copyright 2017 American Chemical Society. Effect of adding divalent MDACl2 on (d) carrier lifetime, (e) XPS
results, and (f) TOF-SIMS profiles of pure α-FAPbI3. Reproduced by permission from ref 316. Copyright 2019 American Association for the
Advancement of Science.

(UPS), and the absorption measurements showed an excellent ful strategies for passivating point defects and interfacial defects
match in the CB minimum, which allowed efficient electron of the PSC, including various additive-based and non-
transfer to TiO2−Cl, whereas the high offset in the VB stoichiometric approaches will be summarized below.
maximum provided efficient hole blocking. Therefore, the 3.4.3.1. Passivation of Bulk Defects. The bulk defects in
TiO2−Cl device showed a high conversion efficiency of 21.4% perovskites are mostly point defects, such as cation/anion
with negligible I−V hysteresis. DFT calculations also reveal the vacancies, interstitials, and substituted ions, which are known
stronger chemical binding at the TiO2−Cl/perovskite interface to act as nonradiative recombination centers with shallow
as opposed to the relatively weak binding observed between energy levels. Son et al. systematically studied the effect of
the bare TiO2 and perovskite, indicating that the TiO2−Cl alkali metal ions (i.e., Li+, Na+, K+, Rb+, and Cs+) on the I−V
device was likely to undergo stronger electronic coupling at the hysteresis of PSCs (Figure 27a).313 It was found that the I−V
ETL/perovskite interface and exhibit enhanced long-term hysteresis strongly depends on the size of the alkali metal ions,
stability. Inorganic materials have been proposed by some and the addition of potassium ions, which have an intermediate
research groups as interface modifiers for TiO2 ETLs. Han et ionic radius, completely eliminate I−V hysteresis unlike smaller
al. introduced a thin layer of MgO between TiO2 and the (i.e., Li+ and Na+) or larger (i.e., Rb+ and Cs+) metal ions.
perovskite, where the MgO layer was found to increase the Voc Theoretical calculations revealed that the reduced I−V
and FF due to the retarded charge recombination.312 However, hysteresis observed upon the addition of potassium ions was
it should be noted that a thicker MgO layer blocks charge associated with the reduced number of Frenkel defects. The
extraction from the perovskite to TiO2. Therefore, maintaining potassium doping effect is much more pronounced for mixed-
a sufficiently low thickness is essential in this approach. Ito et cation perovskites because the potassium interstitials can be
al. deposited a thin Sb2S3 layer between perovskite and TiO2 in energetically more stabilized. Saidaminov et al. reported that
an n-i-p device constructed using the CuSCN HTL.244 The the formation energy of the atomic vacancies increases, and
device showed an improved Jsc and conversion efficiency in the thus, the formation of vacancies can be successfully suppressed
presence of the Sb2S3 layer and, more importantly, significant upon the addition of isovalent small ions, such as Cd2+ and
improvement in the durability of the solar cell. Cl−.314 Although the device performance was barely affected,
3.4.3. Defect Passivation. As discussed in earlier chapters, the addition of isovalent ions greatly improved the long-term
highly efficient PSCs are based on a polycrystalline film and stability due to the reduced number of vacancies, which are
contain a significant amount of defects with various origins, prone to react with water and oxygen molecules. Rb ions are
including vacancies, interstitials, GBs, interfaces, impurities, often added to stabilize α-FAPbI3,315 but this strategy has
and atomic clustering/segregation. The density of defects and rarely been reported for MAPbI3. Park et al. have reported that
their properties directly influence the device performance of Rb ions can also be incorporated into MAPbI3 at levels up to 5
solar cells, especially the Voc and FF, because these defects are at% (i.e., Rb0.05MA0.95PbI3) for a p-i-n devices (Figure
generally nonradiative recombination centers. Several success- 27b,c).159 The addition of Rb+ effectively reduces the density
AG https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 28. (a) Formation mechanism and TEM images of the (quasi-)2D/3D structure constructed using PEAI and Pb(SCN)2 as additives.
Reproduced with permission from ref 283. Copyright 2019 Cell Press. (b) Cross-linking mechanism using TMTA molecules. Reproduced with
permission from ref 319. Copyright 2018 Springer Nature.

of the deep-level trap states and increases the electrical ethylammonium (EA+), n-butylammonium (BA+), isobutyl
conductivity of the perovskite layer, leading to improved Voc ammonium (iso-BA+), and GA+ have also been used to
and Jsc, respectively. Min et al. introduced the divalent cation, passivate the interface/surface, most of which improve the
methylene diammonium (+H3N−CH2−NH3+; MDA), to pure device performance and carrier lifetime by forming two-
α-FAPbI3 by adding MDACl2 molecules into the perovskite dimensional (2D) perovskite layers as mentioned in section
coating solution (Figure 27d−f).316 The addition of ∼5% of 3.2, with the exception of those with longer alkyl chains, such
MDACl2 successfully stabilized the α-FAPbI3 phase due to the as octyl ammonium (OA+), which directly act as passivation
stronger ionic interactions of the divalent state, leading to a molecules rather than forming the 2D perovskite.317 Recently,
high conversion efficiency of 23.7% and excellent thermal alkylammonium halides bearing benzene moieties have been
stability over 20 h at 150 °C under an ambient atmosphere. It shown to exhibit good passivation properties, among which
should be noted that the existence of the divalent cation allows phenethylammonium iodide (PEAI) is the most promising
Cl− interstitials in the lattice of the perovskite, which enables additive used in highly efficient PSCs. Jiang et al. reported a
increased charge-carrier lifetime. The existence of Cl− ions in highly efficient (23.2%) SnO2-based planar p-i-n device
the perovskite layer was confirmed using TOF-SIMS and XPS constructed using a PEAI passivation layer, where the best
analysis. passivation effect was achieved without annealing due to the
3.4.3.2. Passivation of Grain-Boundary/Surface Defects. transformation of PEAI into the 2D perovskite (i.e., PEA2PbI4)
The most popular additives used to passivate GBs or surface at elevated temperature.261 The proposed passivation mecha-
defects are alkylammonium halides because they can passivate nism of PEAI involves the synergetic effect of the π-conjugated
both cation and anion defects and may enable enhanced structure of the benzene ring, which promotes charge transport
passivation via hydrogen/ionic bonding. Using a nonstoichio- and reduces the neutral defects, the amine groups which form
metric approach, Son et al. first reported that an excess of hydrogen bonds with the iodide ions and coordinate with the
methylammonium iodide (MAI) is precipitated at GBs, where Pb2+ interstitials, and the iodide ions that fill the iodine
the MAI layer suppresses nonradiative recombination and vacancies. However, the amount of PEAI added is limited by
improves hole and electron extraction at the GBs by forming its insulating properties. Kim et al. introduced the PEAI
highly ionic conducting pathways.10 The MAPbI3-based n-i-p additive together with Pb(SCN)2 into a wide-bandgap PSC
device constructed using the MAI passivation layer exhibited a (1.68 eV), where the reduced defect density and energetic
high conversion efficiency of >20% without I−V hysteresis. disorder as well as the improved charge carrier mobility and
Alkylammonium halides bearing long alkyl chains such as lifetime led to a high conversion efficiency of 19.8% (Figure
AH https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

28a).283 The simultaneous use of PEA+ (from PEAI) and terminal monolithic tandem solar cell, two subcells are directly
SCN− (from Pb(SCN)2) was crucial for the formation of connected in series with a tunnel junction326 or recombination
(quasi-)2D/3D structures at the GBs and good device layer,302 in which the overall tandem current is limited by the
performance. subcell exhibiting the lowest photocurrent. Therefore, it is very
Cross-linking grains with functional organic molecules has important that each subcell generates the same photocurrent
been shown to enable passivation of the defects/imperfections by efficiently sharing the incident photons in order to obtain a
located at the GBs of a perovskite layer. Li and Grätzel et al. high matching photocurrent of the tandem solar cell. The
introduced a bifunctional molecule, butylphosphonic acid 4- photocurrent in the subcells can be successfully matched by
ammonium chloride (4-ABPACl), which cross-linked the controlling the optical properties, such as the bandgap/
MAPbI3 grains via the formation of hydrogen bonds (O− transmittance of the layers,176,280 light scattering,327 and up/
H···I and N−H···I).318 Supramolecular hydrogen bonds were down-conversion of incident light.328 Most perovskite-based
found to be formed between 4-ABPACl and MAPbI3 in the tandem solar cells including the champion device are prepared
precursor solution, and the conversion efficiency increased using a crystalline Si bottom cell. The optimum bandgap
from 8.8 to 16.5% in the presence of these cross-linking energy of the perovskite top cell that perfectly matches that of
molecules. Li and Fang et al. employed trimethylolpropane Si (1.12 eV) is 1.68 eV, which is somewhat higher than that for
triacrylate (TMTA) in the MAPbI3 layer, in which the TMTA an optimal PSC and requires a mixed halide system
molecules were chemically anchored to the GBs and then incorporating Br. To improve the photovoltage of tandem
cross-linked in situ to form a robust continuous polymeric solar cells, the photovoltage of each subcell has to be improved
network after thermal annealing (Figure 28b).319 The addition independently. Therefore, numerous approaches that have
of TMTA improved the Voc and conversion efficiency, which been applied toward improving the photovoltage of single-
was attributed to the passivation of the defects. More junction PSCs can be applied to large-bandgap perovskites,
interestingly, the cross-links formed by the TMTA molecules such as the passivation of defects using additives.261,283
greatly improved the stability of the device against UV, However, the relatively large Voc-deficit (i.e., Eg−Voc) and
temperature, and moisture. The operational stability of the low photostability of mixed halides (e.g., phase segregation)
MAPbI3-based device was improved 590-fold upon TMTA have yet to be addressed.329,330 The FF values of tandem solar
addition and the subsequent cross-linking process. Another cells can be increased by either reducing the dark current and/
powerful approach to passivate both positively and negatively or shunt conductance or by reducing the series resistance of
charged ionic defects is to use zwitterions that possess both the multiple layers and the interfacial charge transfer resistance
positively and negatively charged functional groups. Zheng et between them.331 In addition, developing more reliable
al. first reported choline zwitterions, such as L-α-phosphati- characterization methods including EQE/J−V measurements
dylcholine, choline chloride, and choline iodide, which can and electron dynamics for tandem solar cells and their subcells
passivate both types of charged defects in halide perovskites.320 are important areas of research.302,332
Theoretical analysis suggested that these zwitterions can The first successful demonstration of a perovskite/Si tandem
passivate two typical deep surface defects consisting of anionic solar cell with a certified conversion efficiency of 23.6% was
Pb−I antisites and cationic Pb clusters. Upon the addition of provided by Bush et al. A perovskite top cell with an Eg of 1.63
choline chloride, the conversion efficiency of the MAPbI3 and eV (Cs0.17FA0.83Pb(Br0.17I0.83)3) was directly deposited onto a
mixed-cation/anion perovskite devices increased from 17.1 to polished Si heterojunction bottom cell with a textured
20.0% and 19.2 to 21.0%, respectively. The improvement in backside.323 A number of additional interfacial layers were
the efficiency was mainly attributed to the increased Voc, and deposited to fabricate the functioning tandem solar cell, such
the Voc deficit of the optimal device constructed using the as the SnO2/ZTO layer used to protect the perovskite cell
mixed cations/anions was as low as 0.39 V. from the plasma damage that occurs during the deposition of
the ITO window layer and the silicon nanoparticle layer on the
3.5. Over Shockley−Queisser Limit
textured backside used to improve the near IR photoresponse
3.5.1. Tandem Solar Cells. The formation of a tandem of the Si bottom cell. More important to note than the high
configuration, in which two or more photosystems with conversion efficiency is that the tandem solar cell passed a
different bandgap energies are combined in series or parallel, is damp heat test designed for crystalline silicon modules (IEC
the only feasible approach to achieve conversion efficiencies 61215), suggesting that the stability of perovskite-based
above the Shockley−Queisser limit. This has been proven by tandem solar cells can be significantly improved with the aid
the high efficiencies observed in multijunction solar cells.321 In of an appropriate encapsulation agent, such as an inorganic
a double-junction solar cell (or tandem solar cell with two TCO layer, EVA film, and the glass cover sheet. One can
photosystems), for example, the top cell with a higher bandgap expect significantly higher photoresponses with textures on
energy absorbs high-energy photons, whereas the bottom cell both sides of the Si bottom cell, especially toward the near-IR
with a lower bandgap energy absorbs low-energy photons region; however, directly fabricating a perovskite top cell on
which have transmitted through the top cell.322 In the tandem top of a textured Si bottom cell is quite challenging. Sahli et al.
configuration, a wide range of solar radiation can be effectively reported a vacuum-solution hybrid approach used to fabricate
shared by the two photosystems while minimizing the a monolithic perovskite/Si tandem solar cell with a fully
thermalization loss of the photons’ excess energy. Therefore, textured Si bottom cell.333 Because of the effective light
one can expect much higher conversion efficiencies for tandem scattering of the textured top surface, a very high overall
solar cells when compared with their single-junction counter- tandem current density close to 20 mA/cm2 and high
parts. conversion efficiency of >25% could be achieved. As a result
In terms of the device structure, tandem solar cells can be of the continued research efforts of several research groups, the
classified as monolithic two-terminal,323,302 mechanical four- highest efficiency of perovskite/Si tandem solar cells has
terminal,324 and optical four-terminal devices.325 In a two- recently reached 29.1%.321 Very recently, several notable
AI https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 29. (a) Triple-halide perovskites with enhanced stability. Reproduced with permission from ref 335. Copyright 2020 American Association
for the Advancement of Science. (b) Microscopy/performance of a perovskite top cell with anion-engineered 2D additives and a corresponding
tandem device. Reproduced with permission from ref 336. Copyright 2020 American Association for the Advancement of Science.

Figure 30. (left) Structure of the monolithic two-terminal perovskite/CIGS tandem solar cell and its photovoltaic properties. Reproduced with
permission from ref 337. Copyright 2018 American Association for the Advancement of Science. (right) TEM image of the GuaSCN passivated
perovskite grains and photovoltaic properties of the GuaSCN-passivated monolithic two-terminal perovskite/perovskite tandem solar cell.
Reproduced with permission from ref 280. Copyright 2019 American Association for the Advancement of Science.

papers on highly efficient and stable perovskite−Si tandem improve the solar cell performance and stability of the
solar cells have been published. Hou et al. demonstrated that perovskite top cell, and reported a highly efficient and stable
solution processes can be used to prepare perovskite top cells perovskite-Si tandem solar cell with an operational efficiency of
on textured Si bottom cells and reported a certified efficiency 26.7% and certified efficiency of 26.2% (Figure 29b).336
of 25.7% with good device stability.334 Xu et al. reported that Perovskite-based tandem solar cells constructed using thin-
the phase segregation observed in perovskite top cells using film-based bottom cells have also been actively investigated
mixed-halide (i.e., Br/I) compositions can be successfully because of the various materials available with low bandgap
suppressed by the incorporation of Cl− ions and, thus, by the energies of 1.0−1.2 eV as well as their potential applications,
formation of triple-halide compositions (Figure 29a).335 A which are not possible with Si bottom cells, such as flexible
perovskite-Si tandem solar cell constructed using the triple- solar cells. A mechanically stacked four-terminal tandem solar
halide perovskite showed a conversion efficiency of 27.04% (its cell constructed from perovskite and CIGS subcells with an
maximum power point tracking was not certified). Kim et al. optimal efficiency of 22.1% has been reported by Fu et al.,324 in
proposed the concept of anion-engineered additives to which the perovskite top cell exhibited a low temperature
AJ https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 31. (a) Efficiency contour map obtained for an ideal double-junction tandem solar cell. Reproduced with permission from ref 322.
Copyright 2017 Elsevier. (b) Efficiency contour map obtained for a triple-junction tandem solar cell, where the first and second cells refer to the
top and middle cells, respectively and the bandgap of the bottom cell was fixed at 1.1 eV. Reproduced with permission from ref 340. Copyright
2015 Elsevier. (c) Schematic representation of the three-terminal measurement platform and (d) EQE curves and (e) Nyquist plots obtained for
the subcells using the three-terminal platform. Reproduced with permission from ref 302. Copyright 2018 Elsevier. (f) FF as a function of the
current mismatching between subcells. Reproduced with permission from ref 332. Copyright 2019 Royal Society of Chemistry.

coefficient of −0.18%/°C (25−65 °C) and a reversible light- recombination.175 Tong et al. reported that the addition of
soaking effect due to the photoconductivity of ZnO nano- GASCN to the Sn−Pb perovskite effectively reduces the defect
particles. Later, Kim et al. reported a higher efficiency of 25.9% density, leading to a carrier lifetime of >1 μs and diffusion
for a four-terminal device by using bimolecular additives283 length of >2.5 μm for the bottom cell (Figure 30b).280 As a
and Han et al. reported a certified conversion efficiency of result, the resulting all-perovskite-based tandem solar cells
22.43% for a monolithic two-terminal perovskite/CIGS exhibit high conversion efficiencies of 23.1 and 25% in the two-
tandem solar cell constructed using a PTAA HTL, PCBM terminal and four-terminal devices, respectively. More recently,
ETL, and ZnO nanoparticle protection layer (Figure 30a).337 Lin et al. reported a simple but effective approach to suppress
The top ITO layer of the CIGS bottom cell was polished using the oxidation of Sn2+ to Sn4+ in the precursor solution by
a chemical mechanical polishing method, which was found to incorporating metallic Sn in the precursor solution.178 The
be critical for the fabrication of highly efficient perovskite/ suppression of Sn oxidation was found to successfully passivate
CIGS tandem solar cells. Low-bandgap (∼1.2 eV) PSCs are the defects, reduce the charge recombination losses, and
also good candidates for the bottom cell of perovskite tandem improve the near IR photoresponse, leading to a high
solar cells, where the bandgap of the perovskite top cell must conversion efficiency of 24.8%.
be increased up to 1.8 eV. The first perovskite/perovskite Given the rapidly increasing conversion efficiencies of
tandem solar cell was reported by Eperon et al., in which a perovskite-based tandem solar cells, it is reasonable to expect
conversion efficiency of 17.0% was reported using a ultrahigh efficiencies of >30% in the near future. Then, what
FA 0.83 Cs 0.17 Pb(I 0.5 Br 0.5 ) 3 bottom cell (1.8 eV) and are the key scientific issues or the technical challenges?
FA0.75Cs0.25Pb0.5Sn0.5I3 bottom cell (1.2 eV).279 A combination Bandgap engineering is probably one of the most important
of high bandgap energies was also reported by Forgács et al., issues, which not only involves tuning the bandgap to the
where the wide bandgaps of the MAPbI3 bottom cell (1.55 eV) desired value but also optimization of the single-junction
and the Cs0.15FA0.85Pb(I0.3Br0.7)3 top cell (2 eV) led to a high device performance with a specific bandgap energy. The typical
tandem photovoltage of 2.1 V.338 It is notable that all the layers bandgap energies of perovskite required for tandem solar cells
were deposited using thermal evaporation except for the are 1.2−1.25 eV for the bottom cell of perovskite/perovskite
solution-processed wide-bandgap layer. Further improvements tandem devices, 1.4−1.5 eV for the middle cell of triple
to these perovskite/perovskite tandem solar cells were realized junction devices, 1.65−1.7 eV for the top cell of perovskite/
by improving the Sn-incorporating bottom cell with an Eg of Si(CIGS) tandem devices, 1.75−1.8 eV for the top cell of
1.25 eV.175,280,339 Zhao et al. reported that the thickness of the perovskite−perovskite tandem devices, and 1.9−2.0 eV for the
Sn−Pb perovskite layer should be greater than that of the other top cell of triple junction devices (Figure 31a,b). The biggest
perovskite layer (>1000 nm)174 and also showed that the challenge for small-bandgap perovskites (i.e., 1.2−1.25 eV) is
addition of Cl− ions further improved the device performance to prevent the oxidation of Sn2+ to Sn4+ so that the device can
by increasing the grain size and suppressing charge exhibit high conversion efficiency for a longer time, whereas
AK https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

that for wide-bandgap perovskites (i.e., 1.65−2.0 eV) is the contacts, the extracted charge carriers cannot relax to their
large Voc deficit and light-induced halide segregation. As lower energy states. Because this increases the overall chemical
described in the earlier sections of this review, various additives potential (quasi Fermi-level splitting) from Δμ for conven-
have been investigated toward improving the Voc and FF via tional solar cells to μceh for hot carrier solar cells, the open
defect passivation as well as enhancing the long-term stability circuit voltage of the device can be enhanced, as shown by the
by suppressing Sn2+ oxidation or halide segregation. Increasing following equation,342
ij T yz
eVOC = μceh = ΔEehjjj1 − L zzz + Δμ L
the bandgap without mixing halide ions or minimizing them

j z
T
k Teh {
may be another good approach toward enhancing the
photostability of perovskites, especially for top-cell applications Teh (12)
with a Si bottom cell. In addition, preserving the top surface
texture of the Si bottom cell or developing a fabrication where μceh and Δμ are the quasi Fermi-level splitting in the hot
process for perovskite top cells on textured Si substrates is also carrier solar cells and conventional solar cells, respectively,
important to further increase the matching current of tandem ΔEeh is the energy difference between the ESCs of the
solar cells by improving the NIR photoresponse of the Si electrons and holes, and Teh and TL are the temperatures of the
bottom cell. Another important issue related to the efficiency hot carriers and electrode lattice, respectively. For hot carrier
of perovskite-based tandem solar cells is the accurate solar cells, a maximum efficiency of 66% under 1 sun
measurement of the device performance. It is known that illumination (up to 85% under concentrated light) was
several commercial solar simulators use xenon lamps as the estimated with a bandgap of 0.7−0.9 eV.341,342
light source, and the strong NIR spikes observed in the Although the theoretical modeling predicted the strong
simulated solar spectrum often cause spectral mismatch issues. potential for hot carrier solar cells to overcome the
Spectral mismatch issues can easily be avoided for single- thermodynamic efficiency limit of single-junction solar cells,
junction solar cells by adjusting the light intensity using a it has been challenging to experimentally demonstrate hot
certified reference cell with or without auxiliary cutoff filters, so carrier devices ideally working with efficiencies higher than
that the total number of electrons coming out of the solar cell those of conventional single-junction solar cells. To fabricate
can be calibrated. However, these calibration methods are not hot carrier solar cells, a suitable absorber as well as ESC
effective for tandem solar cells because the NIR photons are materials are necessary. For the absorber material, the hot
only absorbed by the bottom cells. Theoretically, the Jsc values carrier cooling time should be sufficiently long to allow for the
of tandem solar cells and their subcells can be calculated using extraction of the hot carriers prior to relaxation, while the ESC
EQE curves without being disrupted by spectral mismatch materials should have a narrow DOS at an appropriate energy
issues, but measuring accurate subcell EQE curves is also quite level to extract the hot carriers from the absorber layer.
challenging. Park et al. reported a three-terminal measurement Previously, group III−V semiconductor compounds have been
platform for monolithic perovskite/Si tandem solar cells, in actively investigated for application as an absorbing layer in hot
which the EQE curves of each subcell and the tandem device carrier solar cells. Well-established epitaxial growth techniques
can be accurately measured without any external bias light or enable the growth of high-quality crystals with low defect
bias potential (Figure 31c−e).302 With this measurement density and desirable nanostructures, minimizing the inter-
platform, one can easily measure the actual Jsc that the tandem actions between the charge carriers and phonons. For example,
solar cell can produce under 1 sun/AM1.5 irradiation a carrier cooling lifetime of 5.8 ± 0.1 ns has been reported for a
conditions, which can also be used to investigate the electron GaAsP/InGaAs quantum well structure with a steady-state
dynamics of each subcell without being disturbed by the other carrier population temperature of >100 K above the lattice
subcell. However, it should be noted that the device is likely to temperature under 1000 sun.343 In addition, a hot carrier
show some current mismatching between the subcells during lifetime close to 1 ns with a carrier temperature exceeding 1000
the J−V measurements, even if an accurately matched current K has been reported for InAs/GaAs quantum dots.344
density can be obtained from the EQE measurements. It has Encouragingly, an increase in the current density and voltage
been reported that the current mismatching observed between in a InGaAsP-based quantum well heterostructure due to the
the subcells causes an over/underestimation of the FF contribution from the hot carriers has been reported.345
obtained for tandem solar cells (Figure 31f).303,312 However, it is still challenging to demonstrate the efficiency
3.5.2. Hot Carrier Device. In addition to tandem devices, gain over a conventional single-junction device because of the
another feasible approach to reducing the energy loss from the difficulty in achieving a long hot carrier lifetime with all the
thermalization of photoexcited charge carriers is the utilization desirable properties to be used as an absorber material in solar
of hot carriers. The concept of hot carrier solar cells was cells (e.g., bandgap, high absorption coefficient, long carrier
proposed in the 1980s.341 In single-junction solar cells, diffusion lengths, high mobility, etc.).
photons with energies higher than the bandgap generate Recently, halide perovskite materials have attracted increas-
excited charge carriers populated above the band edge states ing attention in this field after a relatively long carrier cooling
(electrons at states above the CB minimum and holes below lifetime (>100 ps) was reported by several groups when
the VB maximum). These excited charge carriers are referred compared to classical semiconductors (typically <1 ps).346,347
to as “hot carriers” and rapidly relax to the band edge state From transient absorption and PL measurements, two
(typically in picoseconds) upon interaction with the lattice pronounced cooling stages were observed. The primary
vibrations. The energy loss during the relaxation of charge temperature cooling step occurred during the first stage with
carriers is significant, constituting the major portion of the a time scale in the range of subpicoseconds to 2 ps, while the
energy losses that set the thermodynamic limit of solar cells. In second step extended to several hundreds of picoseconds to
hot carrier solar cells, the hot carriers with excess energy are nanoseconds.132,346,348,349 Upon the development of strategies
extracted to energy-selective contacts (ESCs) with a narrow used to enhance the crystal quality and tailor nanostructures,
DOS prior to relaxation.322 Because of the narrow DOSs in the further improvements in the hot carrier lifetime may be
AL https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 32. Spatial mapping of the emission and photon spectra. (a) Graphical representation of the microscope setup and measurement geometry.
(b) Experimentally measured light emission map obtained for different separation distances between the excitation and collection points (a.u.,
arbitrary units). (c) Comparison of the normalized PL with PL excitation (PLE) and photothermal deflection spectra. (d) Predicted spatial light
emission spectra obtained from the cylindrically decaying Beer−Lambert law. (e) Comparison between the experimental (solid lines) and
theoretical decay (dashed lines) data obtained from the Beer−Lambert law at 765 and 800 nm. The experimental data are not in agreement with
simple linear absorption, which suggests that additional processes, such as photon recycling, maintain a substantial photon intensity at large
distances. Reproduced with permission from ref 353. Copyright 2016 American Association for the Advancement of Science. (f) Theoretical open-
circuit voltage (Voc) obtained for various Schottky−Read−Hall (SRH) lifetimes (τSRH) for the case assuming Lambert−Beer-like absorption
(planar structure, top panel) and assuming a textured structure with Lambertian light trapping (mid panel). In each case, we show the situation
with and without photon recycling, whereby the case without photon recycling is unphysical but serves as a reference point. The bottom panel
shows the difference (ΔVocPR) between the cases with and without photon recycling. Note that the calculation of the internal generation rate due to
photon recycling for the Lambert−Beer case was limited to 104 reflections. Reproduced with permission from ref 354. Copyright 2016 American
Chemical Society.

possible. However, the exact mechanism for the extended have attempted to explore the possibility of extracting hot
carrier cooling lifetime still appears to be lacking general carriers with existing contact materials.351 Therefore, intensive
consent. Several different mechanisms such as the hot-phonon research effort is required to demonstrate efficient hot carrier
bottleneck, acoustical-optical phonon up-conversion, Auger PSCs.
heating, and large polarons have been suggested, but some of 3.5.3. Photon Recycling. Inspired from contradictory
the results contradict one another.132,346,349 Therefore, further observations such as the coexistence of long charge carrier
investigations are required to unravel the mechanism of hot diffusion paths and the high density of crystalline disorder, the
carrier cooling in halide perovskite materials. Understanding possibility of an active photon recycling process in perovskite
the exact origin of the long hot carrier lifetime will allow the thin films has been investigated. Effective recycling of the
design of customized perovskite materials and nanostructures photons can prolong the charge carrier lifetime beyond their
for use in hot carrier solar cells. Although still at the initial diffusion lengths where energy transport occurs via multiple
stage of investigations, an encouraging observation is that hot cycles of charge recombination and reabsorption. This photon
carrier transport in halide perovskite thin films is superior to recycling process can enhance the charge carrier density in the
that in conventional semiconductors.350 Quasi-ballistic trans- active layer of solar cells and, thus, the quasi-Fermi level
port of the hot carriers driven by excess kinetic energy in a few splitting, increasing the Voc of the devices toward the
hundreds of femtoseconds has been measured to be as long as Shockley−Queisser limit. In GaAs solar cells, photon recycling
∼230 nm, and their subsequent nonequilibrium transport over was found to play a central role in enhancing the PCE of these
tens of picoseconds persists over ∼600 nm. The quasi-ballistic devices.352
transport lengths are much greater than those observed for Pazos-Outón et al. designed an experimental setup to probe
GaAs (∼85 nm), Si (∼20 nm), and GaN (∼14 nm), indicating the photon recycling processes in perovskite thin films (Figure
that the possibility of harvesting the hot carriers from realistic 32a).353 Using photophysical studies, the charge carrier
devices is much higher than that from conventional solar diffusion lengths in perovskite thin films were measured to
cells.350 However, despite the promising hot carrier lifetime of be on a scale of ∼1 μm, implying that the PL generated by
perovskite materials, the development of ESC materials with bimolecular recombination should be mostly observed near the
suitable energy states is necessary to extract the hot carriers excitation spot. However, the research group observed that the
without losing energy. Only a few studies published to date PL emission intensity away from the excitation spot was
AM https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

significantly higher than that calculated theoretically (Figure where n is an ideality factor, k is the Boltzmann constant, T is
32b−e). As shown in Figure 32b,d, upon increasing the temperature, e is the elementary charge, and J0 is the reverse
distance from the excitation spot to the detection spot, the saturation current. With an X-fold enhancement in the
decay of the measured PL emission became significantly slower irradiation intensity, the Voc is expected to increase because

ÄÅ ÉÑ
the Jsc is enhanced X-fold (XJsc), as shown in eq 14.

nkT ÅÅÅÅ XJSC ÑÑÑ


than that calculated theoretically based on the Beer−Lambert

+ 1ÑÑÑ ≈ Voc(1sun) +
law. The enhanced PL emission was observed up to 50 μm,
lnÅÅ
ÅÅ J ÑÑ
ÅÇ 0 ÑÖ
which is far greater than the measured charge carrier diffusion nkT
Voc = ln X
lengths (Figure 32b,e). They attributed the retarded PL decay e e (14)

ji zy
to the recycling of photons generated from the recombination

PCE ∝ jjj1 + ln X zzz


of excited charge carriers. Using a bottom contact lateral
j z
nkT
k {
device, the recycled photons were found to contribute to the
photocurrent of the devices. On the basis of the measured eVoc(1) (15)
parameters, they suggested that the internal photon density
under 1 sun illumination can be doubled using nonquenching Therefore, the PCE enhancement is dependent on an ideality
electrodes and can be further enhanced by minimizing the factor, the Voc under 1 sun (Voc(1)) and light intensity (X), as
nonradiative decay channels. Kirchartz et al. estimated that the shown in eq 15. Because the dominant charge recombination
voltage gain by photon recycling is in the range of 50−100 mV mechanism is dependent on the carrier density in the active
for planar devices and 10−50 mV for textured devices (Figure layer, n varies as the irradiation intensity increases. At a low
carrier concentration, monomolecular recombination is dom-
32f).354
inant (n = 2), while as the carrier concentration increases, the
While Pazos-Outón et al. suggested the significant
dominant recombination process changes to bimolecular
contribution from the photon recycling process to the
radiative (n = 1) and Auger recombination (n < 1). While
photovoltaic performance of these devices under 1 sun
eq 15 does not consider the resistance effect with the
illumination, contradictory results have still been reported. assumption of low irradiation intensity, the effect of the
Fang et al. have developed an experimental setup to measure resistance becomes pronounced under a high irradiation
the photon recycling efficiency of perovskite single crystals.355 intensity and, thus, photocurrent density. As the current
By utilizing a polarized excitation light source, they were able density increases, the potential drop at the series resistance
to differentiate the recycled PL emission from the transmitted originating from the selective contacts and electrode becomes
PL. They estimated the photon recycling efficiency in a significant, which causes a deviation in the VOC, FF, and JSC
MAPbBr3 single crystal to be <0.5% under 1 sun illumination values from those in the ideal case.
regardless of its strong reabsorption probability exceeding 70%. The performance of PSCs under concentrated light has been
They also suggested that the observed PL emission in the recently investigated.357,358 Wang et al. studied the perform-
longer-wavelength region originates from the PL filtered by the ance evolution of a planar PSC based on FA0.83Cs0.17Pb2.7Br0.3
crystal itself after self-absorption and multiple reflections. The upon increasing the light irradiation intensity from 0.1 sun to
measured low photon recycling efficiency was ascribed to the 128 sun.337 Under 1 sun illumination, the device showed a
poor PL quantum yield of the perovskite under 1 sun PCE of 21.1% (Jsc = 23.2 mA/cm2, Voc = 1.14 V, and FF =
illumination. As a result, only a 0.5% increase in the PL 0.80). Notably, with a linear increase in the Jsc, the Voc of the
lifetime, corresponding to a 0.26 mV enhancement in the Voc, device semilogarithmically increases to enhance the PCE up to
was estimated, whereas the 4−6-fold enhancement in the PL 23.6% under 14 sun and 22.9% under 31 sun. The Voc reaches
lifetime was attributed to photon recycling in GaAs.355,356 1.26 V at 53 sun, while the FF starts to deteriorate when the
Because the experiment was conducted based on single irradiation intensity exceeded 10 sun. Upon consideration of
crystals, further studies may be required to characterize and the charge carrier recombination constant and consequent
realize photon recycling in thin film-based PSC devices. estimation of the charge-carrier diffusion lengths in the
3.5.4. Concentrator Solar Cells. Typically, the concen- perovskite absorber under concentrated light, they suggested
trator solar cell architecture is incorporated in a device when that the decrease in the FF originates from the resistance in the
the fabrication cost of the active material is greater than that of charge extraction layers, contact resistance between the
a concentrator mirror. For example, the fabrication processes internal layers, or parasitic resistance in the fluorine-doped
used for single- or multijunction GaAs solar cells based on tin oxide (FTO) electrode. According to their calculations, the
epitaxial growth are very complex and thus, expensive; device performance should not be limited by charge-carrier
consequently, a photovoltaic concentrator array is often used. recombination up to 1000 sun, where the estimated charge
In concentrator solar cells, solar irradiation is concentrated on carrier diffusion lengths are still >1 μm. However, with the
a small active area in the device. development of appropriate contact layers and device
The photocurrent density is almost linearly proportional to architectures, the enhancement in the photostability of
the irradiation intensity at low concentrations (<100 sun) with perovskite materials may be a prerequisite for their application
a reasonably approximated charge extraction rate (∼108 in concentrator solar cell applications.
s−1),357 while an additional enhancement in the Voc via
enhanced quasi-Fermi level splitting contributes to the PCE of 4. SUMMARY
the device. The Voc enhancement can be estimated from the Because the development of solid-state PSCs in 2012, the PCE

ÄÅ ÉÑ
following diode equation, of PSCs has presented an unprecedentedly rapid enhancement

nkT ÅÅÅÅ JSC ÑÑ


Ñ
+ 1ÑÑÑ
from 9.7% to 25.2% in less than 10 years. The high
lnÅÅ
ÅÅ J ÑÑ
performance of these devices was found to be due to their

ÇÅ 0 ÖÑ
Voc = superior optoelectronic properties, such as high absorption
e (13) coefficient, long charge carrier lifetimes and diffusion lengths,
AN https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

and high defect tolerance, which originate from the character- materials based on a Sn/Pb mixed composition has made it
istic chemical bonding nature, crystal structure, and resulting possible to fabricate perovskite/perovskite tandem solar cells.
electronic band structures of metal halide perovskite materials. Adjusting the bandgap of these materials, while minimizing the
Many unique features of perovskites have also been studied potential loss, will facilitate further enhancements in the PCE
to explain the optoelectronic properties of perovskite materials to >30%. However, accurate measurement techniques used to
but still require further study to understand properties such as characterize tandem devices also need to be developed. To
Rashba splitting, large polaron formation, and GB effects. In reduce the thermalization loss, the possibility of utilizing hot
particular, understanding the nature of the defects in perovskite carriers has also been explored. While promising hot carrier
materials appears to be key to further enhancing the lifetimes have been observed, the design of the device used to
performance of PSCs. In terms of materials chemistry, the extract the hot carriers is still challenging. The photon
composition of perovskite materials has evolved from simple recycling effect is another approach used to reduce the
MAPbI3 to more complex mixed compositions or FA-based potential loss due to charge recombination; however, there still
perovskite materials. Recently, various organic cations have persists an active debate on the occurrence of the photon
also been developed to broaden the compositional window of recycling effect in PSCs. Recently, PSCs have also been tested
perovskite materials. However, most of these fundamental under concentrated sunlight, where an enhancement in the
studies have still focused on MAPbI3 and, therefore, further PCE was observed. However, an improvement in the
studies on their extended systems will be required to design photostability of perovskite materials seems to be a
more ideal perovskite materials. The development of various prerequisite for application in concentrator solar cell
process techniques has enabled the formation of high-quality applications.
perovskite films with compact morphologies and low defect
densities. Control over the nucleation and growth has been AUTHOR INFORMATION
found to be critical toward the formation of uniform and highly Corresponding Authors
crystalline perovskite films, which has been enabled using
antisolvent-assisted one-step, two-step, and adduct methods. Nam-Gyu Park − School of Chemical Engineering,
However, the density of the defects in solution-processed films Sungkyunkwan University, Suwon 16419, Republic of Korea;
is still high, and thus, the development of effective methods to orcid.org/0000-0003-2368-6300; Phone: 821095841386;
passivate the defects in the bulk and surface/GBs has been Email: npark@skku.edu
essential to achieve high efficiencies of >20%. Understanding Hyun Suk Jung − School of Advanced Materials Science and
the nature of the defects in state-of-the-art perovskite materials Engineering, Sungkyunkwan University, Suwon 16419, Republic
will facilitate the development of effective strategies to of Korea; orcid.org/0000-0002-7803-6930;
passivate detrimental defects. In full devices, the ETL and Email: hsjung1@skku.edu
HTL, and their interfaces with the perovskite layer, are equally Hyunjung Shin − Department of Energy Science, Sungkyunkwan
important to achieve high efficiency. Although various charge University, Suwon 16419, Republic of Korea; orcid.org/
transporting materials have been adopted in the device, a 0000-0003-1284-9098; Email: hshin@skku.edu
mesoporous TiO2 or planar SnO2 layer for the ETL and spiro- Authors
MeOTAD for the HTL are still widely used in high-efficiency
devices. Recently, interface engineering to improve the energy Jin Young Kim − Department of Materials Science and
level alignment and electrical coupling has also been actively Engineering, Seoul National University, Seoul 08826, Republic
studied. Not only achieving high efficiency but also resolving of Korea; orcid.org/0000-0001-7746-9972
the stability issue is of critical importance to enable the Jin-Wook Lee − SKKU Advanced Institute of Nanotechnology
practical use of high-efficiency PSCs. All the issues covered in (SAINT) and Department of Nanoengineering, Sungkyunkwan
this review, including the structural design, materials chemistry, University, Suwon 16419, Republic of Korea
process engineering, and device physics, are closely related to Complete contact information is available at:
the operational stability of the device. While working toward https://pubs.acs.org/10.1021/acs.chemrev.0c00107
the enhancement of the efficiency of these devices, it is
necessary to simultaneously consider the durability of the PCE. Author Contributions
For example, the durability of defect passivation agents and J. Y. K. and J.-W.L. contributed equally to this work.
interface engineering layers under prolonged operational
conditions should be considered. Such an approach will Notes
ultimately contribute to the development of commercially The authors declare no competing financial interest.
viable perovskite photovoltaic devices.
Biographies
The unique features of perovskite materials enable the use of
other approaches to further enhance the PCE toward and Jin Young Kim is an Associate Professor in the Department of
beyond the Shockley−Queisser limit. Owing to the tunable Materials Science and Engineering at Seoul National University
bandgap and low-temperature fabrication processes of PSCs, (SNU), Korea. He received his B.S. (2000), M.S. (2002), and Ph.D.
their application as the top subcell in tandem devices is easily (2006) in Materials Science at Seoul National University. Before
achievable. At an initial stage, however, tandem devices were joining SNU, he worked at the National Renewable Energy
not able to surpass the PCE of their corresponding single- Laboratory (NREL), USA (2007−2011), and Korea Institute of
junction devices, but intense research efforts on the develop- Science and Technology (KIST), Korea (2011−2015), on next-
ment of interconnecting layers and high-bandgap perovskite generation thin film solar cells. His current research interests include
materials have allowed a rapid enhancement in the PCE to the fabrication of nanostructured electrodes via various approaches
reach 29.1% based on a perovskite/silicon tandem structure. including electrochemical approaches and their applications in solar
Furthermore, the development of low-bandgap perovskite energy utilization including thin-film solar cells (e.g., CIGS, CZTS,

AO https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

perovskite, and tandem solar cells) and solar fuel generation (e.g., H2 Institute of Advanced Materials (RIAM) under contract no.
production and CO2 reduction). 20193091010310. H.S.J. and H.S. are grateful to the MISP for
Jin-Wook Lee is an Assistant Professor at Sungkyunkwan University financial support under contracts NRF-2019M3D1A2104108
(SKKU) Advanced Institute of Nanotechnology (SAINT) and and NRF-2019R1A2C3009157.
Department of Nanoengineering. He received his B.S. in Electronic
and Electrical Engineering and Ph.D. in Energy Science from SKKU REFERENCES
in 2011 and 2016, respectively. From 2016 to 2019, he worked as a (1) Kim, H.-S.; Lee, C.-R.; Im, J.-H.; Lee, K.-B.; Moehl, T.;
Postdoctoral Researcher at the University of California at Los Angeles Marchioro, A.; Moon, S.-J.; Humphry-Baker, R.; Yum, J.-H.; Moser, J.
(UCLA) before joining SKKU as an Assistant Professor in 2019. His E.; et al. Lead Iodide Perovskite Sensitized All-Solid-State Submicron
research has been focused on the development of highly efficient and Thin Film Mesoscopic Solar Cell with Efficiency Exceeding 9%. Sci.
stable perovskite optoelectronics (solar cells, light emitting diodes, Rep. 2012, 2, 591.
and photodetectors) based on material design, crystal growth, and (2) Lee, M. M.; Teuscher, J.; Miyasaka, T.; Murakami, T. N.; Snaith,
defect engineering. H. J. Efficient Hybrid Solar Cells Based on Meso-Superstructured
Organometal Halide Perovskites. Science 2012, 338, 643−647.
Hyun Suk Jung is a SKKU Young Fellow Professor at the School of (3) Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal
Advanced Materials Science and Engineering, Sungkyunkwan Halide Perovskites as Visible-Light Sensitizers for Photovoltaic Cells.
University (SKKU). He received his B.S., M.S., and Ph.D. in J. Am. Chem. Soc. 2009, 131, 6050−6051.
Materials Science and Engineering from Seoul National University (4) Im, J.-H.; Lee, C.-R.; Lee, J.-W.; Park, S.-W.; Park, N.-G. 6.5%
(SNU) in 1997, 1999, and 2004, respectively. He joined Los Alamos Efficient Perovskite Quantum-Dot-Sensitized Solar Cell. Nanoscale
National Laboratory (LANL) as a Director’s Postdoctoral Fellow in 2011, 3, 4088−4093.
2005. He has published over 200 peer-reviewed papers and holds (5) Wenk, H.-R.; Bulakh, A. Minerals: Their Constitution and Origin;
Cambridge University Press, 2016.
more than 50 patents on the synthesis of inorganic nanomaterials and
(6) Heo, J. H.; Im, S. H.; Noh, J. H.; Mandal, T. N.; Lim, C.-S.;
solar energy conversion devices such as photovoltaic and photo- Chang, J. A.; Lee, Y. H.; Kim, H.-j.; Sarkar, A.; Nazeeruddin, M. K.
electrochemical devices. His current research interests include Efficient Inorganic−Organic Hybrid Heterojunction Solar Cells
perovskite solar cells and flexible solar cells. Containing Perovskite Compound and Polymeric Hole Conductors.
Hyunjung Shin is currently a Professor in the Department of Energy Nat. Photonics 2013, 7, 486−491.
Science at Sungkyunkwan University (SKKU), Korea. He received his (7) Lee, J. W.; Seol, D. J.; Cho, A. N.; Park, N. G. High-Efficiency
B.S. degree at YonSei University, Korea in 1991. He then received his Perovskite Solar Cells Based on the Black Polymorph of HC-
(NH2)2PbI3. Adv. Mater. 2014, 26, 4991−4998.
M.S. and Ph.D. at Case Western Reserve University, Cleveland, Ohio,
(8) Im, J.-H.; Jang, I.-H.; Pellet, N.; Grätzel, M.; Park, N.-G. Growth
in 1994 and 1996, respectively. He was an Alexander von Humboldt of CH3NH3PbI3 Cuboids with Controlled Size for High-Efficiency
Research Fellow at the Max-Planck-Institut für Metallforschung in Perovskite Solar Cells. Nat. Nanotechnol. 2014, 9, 927−932.
Stuttgart, Germany (1996−1997). He started his professional career (9) Zhou, H.; Chen, Q.; Li, G.; Luo, S.; Song, T.-b.; Duan, H.-S.;
as a Member of Research Staff in Samsung Advanced Institute of Hong, Z.; You, J.; Liu, Y.; Yang, Y. Interface Engineering of Highly
Technology, Suwon, Korea (1997−2002), and as an Assistant, Efficient Perovskite Solar Cells. Science 2014, 345, 542−546.
Associated, and Full Professor at Kookmin University, Seoul, Korea (10) Son, D.-Y.; Lee, J.-W.; Choi, Y. J.; Jang, I.-H.; Lee, S.; Yoo, P. J.;
(2002−2012). He moved to SKKU in 2012 as a Tenured Full Shin, H.; Ahn, N.; Choi, M.; Kim, D.; et al. Self-Formed Grain
Professor. His research interests include renewable energy conversion Boundary Healing Layer for Highly Efficient CH3NH3PbI3 Perovskite
and storage devices. Solar Cells. Nat. Energy 2016, 1, 16081.
(11) Green, M. A.; Emery, K.; Hishikawa, Y.; Warta, W.; Dunlop, E.
Nam-Gyu Park is a Professor and SKKU-Fellow at the School of D. Solar Cell Efficiency Tables (Version 42). Prog. Photovoltaics 2013,
Chemical Engineering, Sungkyunkwan University (SKKU). He 21, 827−837.
received his B.S., M.S., and Ph.D. degrees from Seoul National (12) Green, M. A.; Emery, K.; Hishikawa, Y.; Warta, W.; Dunlop, E.
University (SNU) in 1988, 1992, and 1995, respectively. He worked D. Solar Cell Efficiency Tables (Version 44). Prog. Photovoltaics 2014,
at ICMCB-CNRS, France, and National Renewable Energy 22, 701−710.
Laboratory (NREL), USA, from 1996 to 1999 as a Postdoctoral (13) Green, M. A.; Emery, K.; Hishikawa, Y.; Warta, W.; Dunlop, E.
D. Solar Cell Efficiency Tables (Version 45). Prog. Photovoltaics 2015,
Researcher. He worked as the Director of Solar Cell Research Center
23, 1−9.
at the Korea Institute of Science and Technology (KIST) in Korea (14) Green, M. A.; Emery, K.; Hishikawa, Y.; Warta, W.; Dunlop, E.
before joining SKKU as a Full Professor in 2009. He is the pioneer of D. Solar Cell Efficiency Tables (Version 48). Prog. Photovoltaics 2016,
the solid-state perovskite solar cell, discovered in 2012. He received 24, 905−913.
the Clarivate Citation Laureates in Chemistry in 2017 and was (15) Green, M. A.; Hishikawa, Y.; Dunlop, E. D.; Levi, D. H.; Hohl-
selected as a Highly Cited Researcher in 2017, 2018, and 2019. Ebinger, J.; Ho-Baillie, A. W. Y. Solar Cell Efficiency Tables (Version
51). Prog. Photovoltaics 2018, 26, 3−12.
(16) Green, M. A.; Hishikawa, Y.; Dunlop, E. D.; Levi, D. H.; Hohl-
ACKNOWLEDGMENTS Ebinger, J.; Yoshita, M.; Ho-Baillie, A. W. Y. Solar Cell Efficiency
This work was supported by National Research Foundation of Tables (Version 53). Prog. Photovoltaics 2019, 27, 3−12.
Korea (NRF) grants funded by the Ministry of Science, ICT (17) Green, M. A.; Dunlop, E. D.; Levi, D. H.; Hohl-Ebinger, J.;
Future Planning (MSIP) of Korea under contracts NRF- Yoshita, M.; Ho-Baillie, A. W. Solar Cell Efficiency Tables (Version
2012M3A6A7054861 (Global Frontier R&D Program on 54). Prog. Photovoltaics 2019, 27, 565−575.
(18) Green, M. A.; Dunlop, E. D.; Hohl-Ebinger, J.; Yoshita, M.;
Center for Multiscale Energy System), NRF- Kopidakis, N.; Ho-Baillie, A. W. Y. Solar Cell Efficiency Tables
2016M3D1A1027663 and NRF-2016M3D1A1027664 (Future (Version 55). Prog. Photovoltaics 2020, 28, 3−15.
Materials Discovery Program), and NRF-2015M1A2A2053004 (19) Burschka, J.; Pellet, N.; Moon, S.-J.; Humphry-Baker, R.; Gao,
(Climate Change Management Program). J.Y.K. is grateful for P.; Nazeeruddin, M. K.; Grätzel, M. Sequential Deposition as a Route
the financial support from the Korea Institute of Energy to High-Performance Perovskite-Sensitized Solar Cells. Nature 2013,
Technology Evaluation and Planning (KETEP) and Research 499, 316.

AP https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

(20) Jeon, N. J.; Noh, J. H.; Kim, Y. C.; Yang, W. S.; Ryu, S.; Seok, S. (41) Richter, J. M.; Abdi-Jalebi, M.; Sadhanala, A.; Tabachnyk, M.;
I. Solvent Engineering for High-Performance Inorganic−Organic Rivett, J. P.; Pazos-Outón, L. M.; Gödel, K. C.; Price, M.; Deschler, F.;
Hybrid Perovskite Solar Cells. Nat. Mater. 2014, 13, 897−903. Friend, R. H. Enhancing Photoluminescence Yields in Lead Halide
(21) Ahn, N.; Son, D.-Y.; Jang, I.-H.; Kang, S. M.; Choi, M.; Park, Perovskites by Photon Recycling and Light out-Coupling. Nat.
N.-G. Highly Reproducible Perovskite Solar Cells with Average Commun. 2016, 7, 13941.
Efficiency of 18.3% and Best Efficiency of 19.7% Fabricated Via Lewis (42) Sarritzu, V.; Sestu, N.; Marongiu, D.; Chang, X.; Masi, S.;
Base Adduct of Lead (II) Iodide. J. Am. Chem. Soc. 2015, 137, 8696− Rizzo, A.; Colella, S.; Quochi, F.; Saba, M.; Mura, A.; et al. Optical
8699. Determination of Shockley-Read-Hall and Interface Recombination
(22) D’innocenzo, V.; Grancini, G.; Alcocer, M. J.; Kandada, A. R. Currents in Hybrid Perovskites. Sci. Rep. 2017, 7, 44629.
S.; Stranks, S. D.; Lee, M. M.; Lanzani, G.; Snaith, H. J.; Petrozza, A. (43) Chen, J.; Park, N. G. Causes and Solutions of Recombination in
Excitons Versus Free Charges in Organo-Lead Tri-Halide Perovskites. Perovskite Solar Cells. Adv. Mater. 2019, 31, 1803019.
Nat. Commun. 2014, 5, 3586. (44) Wolff, C. M.; Caprioglio, P.; Stolterfoht, M.; Neher, D.
(23) Miyata, A.; Mitioglu, A.; Plochocka, P.; Portugall, O.; Wang, J. Nonradiative Recombination in Perovskite Solar Cells: The Role of
T.-W.; Stranks, S. D.; Snaith, H. J.; Nicholas, R. J. Direct Interfaces. Adv. Mater. 2019, 31, 1902762.
Measurement of the Exciton Binding Energy and Effective Masses (45) Jones, T. W.; Osherov, A.; Alsari, M.; Sponseller, M.; Duck, B.
for Charge Carriers in Organic−Inorganic Tri-Halide Perovskites. C.; Jung, Y.-K.; Settens, C.; Niroui, F.; Brenes, R.; Stan, C. V.; et al.
Nat. Nat. Phys. 2015, 11, 582−587. Lattice Strain Causes Non-Radiative Losses in Halide Perovskites.
(24) Meissler, G. L.; Fischer, P. J.; Tarr, D. A. Inorganic Chemistry; Energy Environ. Sci. 2019, 12, 596−606.
5th ed.; Pearson, 2013. (46) Baikie, T.; Fang, Y.; Kadro, J. M.; Schreyer, M.; Wei, F.;
(25) Clementi, E.; Raimondi, D.-L. Atomic Screening Constants Mhaisalkar, S. G.; Graetzel, M.; White, T. J. Synthesis and Crystal
from Scf Functions. J. Chem. Phys. 1963, 38, 2686−2689. Chemistry of the Hybrid Perovskite (CH3NH3)PbI3 for Solid-State
(26) Clementi, E.; Raimondi, D.; Reinhardt, W. P. Atomic Screening Sensitised Solar Cell Applications. J. Mater. Chem. A 2013, 1, 5628−
Constants from Scf Functions. Ii. Atoms with 37 to 86 Electrons. J. 5641.
Chem. Phys. 1967, 47, 1300−1307. (47) Li, C.; Lu, X.; Ding, W.; Feng, L.; Gao, Y.; Guo, Z. Formability
(27) Howard, S. Ab Initio Effective Core Potential Calculations on of ABX3 (X= F, Cl, Br, I) Halide Perovskites. Acta Crystallogr., Sect. B:
HgI2, PtI2, and PbI2. J. Phys. Chem. 1994, 98, 6110−6113. Struct. Sci. 2008, 64, 702−707.
(28) Targhi, F. F.; Jalili, Y. S.; Kanjouri, F. MAPbI3 and FAPbI3 (48) Park, N.-G. Perovskite Solar Cells: An Emerging Photovoltaic
Perovskites as Solar Cells: Case Study on Structural, Electrical and Technology. Mater. Mater. Today 2015, 18, 65−72.
Optical Properties. Results Phys. 2018, 10, 616−627. (49) She, L.; Liu, M.; Zhong, D. Atomic Structures of CH3NH3PbI3
(29) Fox, M. Quantum Optics: An Introduction; OUP: Oxford, 2006; (001). ACS Nano 2016, 10, 1126−1131.
Vol. 29, pp 247−267. (50) Egger, D. A.; Kronik, L. Role of Dispersive Interactions in
(30) Yin, W.-J.; Shi, T.; Yan, Y. Superior Photovoltaic Properties of Determining Structural Properties of Organic−Inorganic Halide
Lead Halide Perovskites: Insights from First-Principles Theory. J. Perovskites: Insights from First-Principles Calculations. J. Phys.
Phys. Chem. C 2015, 119, 5253−5264. Chem. Lett. 2014, 5, 2728−2733.
(31) Jiao, Y.; Zhang, S.; Yang, Z.; Lu, G. Indirect-to-Direct Band (51) Hirotsu, S.; Harada, J.; Iizumi, M.; Gesi, K. Structural Phase
Gap Transition and Optical Properties of Metal Alloying of Transitions in CsPbBr3. J. Phys. Soc. Jpn. 1974, 37, 1393−1398.
Cs2AgMxBr6 (M= Bi, in, Sb): Insights from the First Principles. (52) Wasylishen, R. E.; Knop, O.; Macdonald, J. B. Cation Rotation
Comput. Theor. Chem. 2019, 1148, 55−59. in Methylammonium Lead Halides. Solid State Commun. 1985, 56,
(32) Wang, T.; Daiber, B.; Frost, J. M.; Mann, S. A.; Garnett, E. C.; 581−582.
Walsh, A.; Ehrler, B. Indirect to Direct Bandgap Transition in (53) Nandi, P.; Giri, C.; Swain, D.; Manju, U.; Topwal, D. Room
Methylammonium Lead Halide Perovskite. Energy Environ. Sci. 2017, Temperature Growth of CH3NH3PbCl3 Single Crystals by Solvent
10, 509−515. Evaporation Method. CrystEngComm 2019, 21, 656−661.
(33) Gunawan, O.; Pae, S. R.; Bishop, D. M.; Virgus, Y.; Noh, J. H.; (54) Postorino, P.; Malavasi, L. Pressure-Induced Effects in
Jeon, N. J.; Lee, Y. S.; Shao, X.; Todorov, T.; Mitzi, D. B. Carrier- Organic−Inorganic Hybrid Perovskites. J. Phys. Chem. Lett. 2017, 8,
Resolved Photo-Hall Effect. Nature 2019, 575, 151−155. 2613−2622.
(34) Liu, M.; Endo, M.; Shimazaki, A.; Wakamiya, A.; Tachibana, Y. (55) Jaffe, A.; Lin, Y.; Beavers, C. M.; Voss, J.; Mao, W. L.;
Identifying an Optimum Perovskite Solar Cell Structure by Kinetic Karunadasa, H. I. High-Pressure Single-Crystal Structures of 3D Lead-
Analysis: Planar, Mesoporous Based, or Extremely Thin Absorber Halide Hybrid Perovskites and Pressure Effects on Their Electronic
Structure. ACS Appl. Energy Mater. 2018, 1, 3722−3732. and Optical Properties. ACS Cent. Sci. 2016, 2, 201−209.
(35) Herz, L. M. Charge-Carrier Mobilities in Metal Halide (56) Mosconi, E.; Amat, A.; Nazeeruddin, M. K.; Grätzel, M.; De
Perovskites: Fundamental Mechanisms and Limits. ACS Energy Lett. Angelis, F. First-Principles Modeling of Mixed Halide Organometal
2017, 2, 1539−1548. Perovskites for Photovoltaic Applications. J. Phys. Chem. C 2013, 117,
(36) Jiang, Y.; Wang, X.; Pan, A. Properties of Excitons and 13902−13913.
Photogenerated Charge Carriers in Metal Halide Perovskites. Adv. (57) Yun, S.; Zhou, X.; Even, J.; Hagfeldt, A. Theoretical Treatment
Mater. 2019, 31, 1806671. of CH3NH3PbI3 Perovskite Solar Cells. Angew. Chem., Int. Ed. 2017,
(37) Adinolfi, V.; Peng, W.; Walters, G.; Bakr, O. M.; Sargent, E. H. 56, 15806−15817.
The Electrical and Optical Properties of Organometal Halide (58) Hoye, R. L.; Schulz, P.; Schelhas, L. T.; Holder, A. M.; Stone,
Perovskites Relevant to Optoelectronic Performance. Adv. Mater. K. H.; Perkins, J. D.; Vigil-Fowler, D.; Siol, S.; Scanlon, D. O.;
2018, 30, 1700764. Zakutayev, A.; et al. Perovskite-Inspired Photovoltaic Materials:
(38) Reid, O. G.; Yang, M.; Kopidakis, N.; Zhu, K.; Rumbles, G. Toward Best Practices in Materials Characterization and Calculations.
Grain-Size-Limited Mobility in Methylammonium Lead Iodide Chem. Mater. 2017, 29, 1964−1988.
Perovskite Thin Films. ACS Energy Lett. 2016, 1, 561−565. (59) Lindblad, R.; Bi, D.; Park, B.-w.; Oscarsson, J.; Gorgoi, M.;
(39) Rehman, W.; Milot, R. L.; Eperon, G. E.; Wehrenfennig, C.; Siegbahn, H.; Odelius, M.; Johansson, E. M.; Rensmo, H. k.
Boland, J. L.; Snaith, H. J.; Johnston, M. B.; Herz, L. M. Charge- Electronic Structure of TiO2/CH3NH3PbI3 Perovskite Solar Cell
Carrier Dynamics and Mobilities in Formamidinium Lead Mixed- Interfaces. J. Phys. Chem. Lett. 2014, 5, 648−653.
Halide Perovskites. Adv. Mater. 2015, 27, 7938−7944. (60) Frost, J. M.; Walsh, A. What Is Moving in Hybrid Halide
(40) Johnston, M. B.; Herz, L. M. Hybrid Perovskites for Perovskite Solar Cells? Acc. Chem. Res. 2016, 49, 528−535.
Photovoltaics: Charge-Carrier Recombination, Diffusion, and Radia- (61) Drużbicki, K.; Pinna, R. S.; Rudić, S.; Jura, M.; Gorini, G.;
tive Efficiencies. Acc. Chem. Res. 2016, 49, 146−154. Fernandez-Alonso, F. Unexpected Cation Dynamics in the Low-

AQ https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Temperature Phase of Methylammonium Lead Iodide: The Need for (83) Walsh, A.; Scanlon, D. O.; Chen, S.; Gong, X.; Wei, S. H. Self-
Improved Models. J. Phys. Chem. Lett. 2016, 7, 4701−4709. Regulation Mechanism for Charged Point Defects in Hybrid Halide
(62) Even, J.; Pedesseau, L.; Jancu, J.-M.; Katan, C. Importance of Perovskites. Angew. Chem., Int. Ed. 2015, 54, 1791−1794.
Spin−Orbit Coupling in Hybrid Organic/Inorganic Perovskites for (84) Adinolfi, V.; Yuan, M.; Comin, R.; Thibau, E. S.; Shi, D.;
Photovoltaic Applications. J. Phys. Chem. Lett. 2013, 4, 2999−3005. Saidaminov, M. I.; Kanjanaboos, P.; Kopilovic, D.; Hoogland, S.; Lu,
(63) Kim, M.; Im, J.; Freeman, A. J.; Ihm, J.; Jin, H. Switchable S= Z. H.; et al. The in-Gap Electronic State Spectrum of Methyl-
1/2 and J= 1/2 Rashba Bands in Ferroelectric Halide Perovskites. ammonium Lead Iodide Single-Crystal Perovskites. Adv. Mater. 2016,
Proc. Natl. Acad. Sci. U. S. A. 2014, 111, 6900−6904. 28, 3406−3410.
(64) Kepenekian, M.; Even, J. Rashba and Dresselhaus Couplings in (85) Stranks, S. D.; Eperon, G. E.; Grancini, G.; Menelaou, C.;
Halide Perovskites: Accomplishments and Opportunities for Spin- Alcocer, M. J.; Leijtens, T.; Herz, L. M.; Petrozza, A.; Snaith, H. J.
tronics and Spin−Orbitronics. J. Phys. Chem. Lett. 2017, 8, 3362− Electron-Hole Diffusion Lengths Exceeding 1 Micrometer in an
3370. Organometal Trihalide Perovskite Absorber. Science 2013, 342, 341−
(65) Kirchartz, T.; Rau, U. Decreasing Radiative Recombination 344.
Coefficients Via an Indirect Band Gap in Lead Halide Perovskites. J. (86) Löper, P.; Stuckelberger, M.; Niesen, B.; Werner, J.; Filipič, M.;
Phys. Chem. Lett. 2017, 8, 1265−1271. Moon, S.-J.; Yum, J.-H.; Topič, M.; De Wolf, S.; Ballif, C. Complex
(66) Zheng, C.; Rubel, O. Ionization Energy as a Stability Criterion Refractive Index Spectra of CH3NH3PbI3 Perovskite Thin Films
for Halide Perovskites. J. Phys. Chem. C 2017, 121, 11977−11984. Determined by Spectroscopic Ellipsometry and Spectrophotometry. J.
(67) Uratani, H.; Yamashita, K. Charge Carrier Trapping at Surface Phys. Chem. Lett. 2015, 6, 66−71.
Defects of Perovskite Solar Cell Absorbers: A First-Principles Study. J. (87) Almond, D. P.; Bowen, C. R. An Explanation of the
Phys. Chem. Lett. 2017, 8, 742−746. Photoinduced Giant Dielectric Constant of Lead Halide Perovskite
(68) Sherkar, T. S.; Momblona, C.; Gil-Escrig, L. n.; Á vila, J.; Solar Cells. J. Phys. Chem. Lett. 2015, 6, 1736−1740.
Sessolo, M.; Bolink, H. J.; Koster, L. J. A. Recombination in (88) Miyata, K.; Meggiolaro, D.; Trinh, M. T.; Joshi, P. P.; Mosconi,
Perovskite Solar Cells: Significance of Grain Boundaries, Interface E.; Jones, S. C.; De Angelis, F.; Zhu, X.-Y. Large Polarons in Lead
Traps, and Defect Ions. ACS Energy Lett. 2017, 2, 1214−1222. Halide Perovskites. Sci. Adv. 2017, 3, No. e1701217.
(69) Ball, J. M.; Petrozza, A. Defects in Perovskite-Halides and Their (89) Hill, A. H.; Smyser, K. E.; Kennedy, C. L.; Massaro, E. S.;
Effects in Solar Cells. Nat. Energy 2016, 1, 16149. Grumstrup, E. M. Screened Charge Carrier Transport in Methyl-
(70) Brenes, R.; Guo, D.; Osherov, A.; Noel, N. K.; Eames, C.; ammonium Lead Iodide Perovskite Thin Films. J. Phys. Chem. Lett.
Hutter, E. M.; Pathak, S. K.; Niroui, F.; Friend, R. H.; Islam, M. S.;
2017, 8, 948−953.
et al. Metal Halide Perovskite Polycrystalline Films Exhibiting (90) Kang, J.; Wang, L.-W. High Defect Tolerance in Lead Halide
Properties of Single Crystals. Joule 2017, 1, 155−167.
Perovskite Cspbbr3. J. Phys. Chem. Lett. 2017, 8, 489−493.
(71) Li, W.; Long, R.; Tang, J.; Prezhdo, O. V. Influence of Defects
(91) Li, W.; Liu, J.; Bai, F.-Q.; Zhang, H.-X.; Prezhdo, O. V. Hole
on Excited State Dynamics in Lead Halide Perovskites: Time-Domain
Trapping by Iodine Interstitial Defects Decreases Free Carrier Losses
Ab Initio Studies. J. Phys. Chem. Lett. 2019, 10, 3788−3804.
in Perovskite Solar Cells: A Time-Domain Ab Initio Study. ACS
(72) Yin, W.-J.; Shi, T.; Yan, Y. Unusual Defect Physics in
Energy Lett. 2017, 2, 1270−1278.
Ch3nh3pbi3 Perovskite Solar Cell Absorber. Appl. Phys. Lett. 2014,
(92) West, B. M.; Stuckelberger, M.; Guthrey, H.; Chen, L.; Lai, B.;
104, 063903.
Maser, J.; Rose, V.; Shafarman, W.; Al-Jassim, M.; Bertoni, M. I. Grain
(73) Meggiolaro, D.; De Angelis, F. First-Principles Modeling of
Defects in Lead Halide Perovskites: Best Practices and Open Issues. Engineering: How Nanoscale Inhomogeneities Can Control Charge
ACS Energy Lett. 2018, 3, 2206−2222. Collection in Solar Cells. Nano Energy 2017, 32, 488−493.
(74) Han, D.; Dai, C.; Chen, S. Calculation Studies on Point Defects (93) Terry, M. L.; Straub, A.; Inns, D.; Song, D.; Aberle, A. G. Large
in Perovskite Solar Cells. J. Semicond. 2017, 38, 011006. Open-Circuit Voltage Improvement by Rapid Thermal Annealing of
(75) Yin, W.-J.; Yang, J.-H.; Kang, J.; Yan, Y.; Wei, S.-H. Halide Evaporated Solid-Phase-Crystallized Thin-Film Silicon Solar Cells on
Perovskite Materials for Solar Cells: A Theoretical Review. J. Mater. Glass. Appl. Phys. Lett. 2005, 86, 172108.
Chem. A 2015, 3, 8926−8942. (94) Yan, Y.; Jiang, C.-S.; Noufi, R.; Wei, S.-H.; Moutinho, H.; Al-
(76) Frolova, L. A.; Dremova, N. N.; Troshin, P. A. The Chemical Jassim, M. Electrically Benign Behavior of Grain Boundaries in
Origin of the P-Type and N-Type Doping Effects in the Hybrid Polycrystalline CuInSe2 Films. Phys. Rev. Lett. 2007, 99, 235504.
Methylammonium−Lead Iodide (MAPbI3) Perovskite Solar Cells. (95) Li, J. B.; Chawla, V.; Clemens, B. M. Investigating the Role of
Chem. Commun. 2015, 51, 14917−14920. Grain Boundaries in Czts and Cztsse Thin Film Solar Cells with
(77) Zohar, A.; Levine, I.; Gupta, S.; Davidson, O.; Azulay, D.; Scanning Probe Microscopy. Adv. Mater. 2012, 24, 720−723.
Millo, O.; Balberg, I.; Hodes, G.; Cahen, D. What Is the Mechanism (96) Li, C.; Wu, Y.; Poplawsky, J.; Pennycook, T. J.; Paudel, N.; Yin,
of MAPbI3 P-doping by I2? Insights from Optoelectronic Properties. W.; Haigh, S. J.; Oxley, M. P.; Lupini, A. R.; Al-Jassim, M.; et al.
ACS Energy Lett. 2017, 2, 2408−2414. Grain-Boundary-Enhanced Carrier Collection in Cdte Solar Cells.
(78) Motti, S. G.; Meggiolaro, D.; Martani, S.; Sorrentino, R.; Phys. Rev. Lett. 2014, 112, 156103.
Barker, A. J.; De Angelis, F.; Petrozza, A. Defect Activity in Lead (97) Lee, J.-W.; Bae, S.-H.; De Marco, N.; Hsieh, Y.-T.; Dai, Z.;
Halide Perovskites. Adv. Mater. 2019, 31, 1901183. Yang, Y. The Role of Grain Boundaries in Perovskite Solar Cells.
(79) Vogel, D. J.; Inerbaev, T. M.; Kilin, D. S. Role of Lead Mater. Today Energy 2018, 7, 149−160.
Vacancies for Optoelectronic Properties of Lead-Halide Perovskites. J. (98) Adhyaksa, G. W.; Brittman, S.; A̅ boliņs,̌ H.; Lof, A.; Li, X.;
Phys. Chem. C 2018, 122, 5216−5226. Keelor, J. D.; Luo, Y.; Duevski, T.; Heeren, R. M.; Ellis, S. R.; et al.
(80) Yang, W. S.; Park, B.-W.; Jung, E. H.; Jeon, N. J.; Kim, Y. C.; Understanding Detrimental and Beneficial Grain Boundary Effects in
Lee, D. U.; Shin, S. S.; Seo, J.; Kim, E. K.; Noh, J. H.; et al. Iodide Halide Perovskites. Adv. Mater. 2018, 30, 1804792.
Management in Formamidinium-Lead-Halide−Based Perovskite (99) Haruyama, J.; Sodeyama, K.; Han, L.; Tateyama, Y. Surface
Layers for Efficient Solar Cells. Science 2017, 356, 1376−1379. Properties of CH3NH3PbI3 for Perovskite Solar Cells. Acc. Chem. Res.
(81) He, J.; Long, R. Lead Vacancy Can Explain the Suppressed 2016, 49, 554−561.
Nonradiative Electron−Hole Recombination in Fapbi3 Perovskite (100) Haruyama, J.; Sodeyama, K.; Han, L.; Tateyama, Y.
under Iodine-Rich Conditions: A Time-Domain Ab Initio Study. J. Termination Dependence of Tetragonal CH3NH3PbI3 Surfaces for
Phys. Chem. Lett. 2018, 9, 6489−6495. Perovskite Solar Cells. J. Phys. Chem. Lett. 2014, 5, 2903−2909.
(82) Liu, N.; Yam, C. First-Principles Study of Intrinsic Defects in (101) Yin, W. J.; Shi, T.; Yan, Y. Unique Properties of Halide
Formamidinium Lead Triiodide Perovskite Solar Cell Absorbers. Phys. Perovskites as Possible Origins of the Superior Solar Cell Perform-
Chem. Chem. Phys. 2018, 20, 6800−6804. ance. Adv. Mater. 2014, 26, 4653−4658.

AR https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

(102) Yin, W. J.; Chen, H.; Shi, T.; Wei, S. H.; Yan, Y. Origin of (120) Bae, S.; Park, J.-S.; Han, I. K.; Shin, T. J.; Jo, W. H.
High Electronic Quality in Structurally Disordered CH3NH3PbI3 and CH3NH3PbI3 Crystal Orientation and Photovoltaic Performance of
the Passivation Effect of Cl and O at Grain Boundaries. Adv. Electron. Planar Heterojunction Perovskite Solar Cells. Sol. Energy Mater. Sol.
Mater. 2015, 1, 1500044. Cells 2017, 160, 77−84.
(103) Long, R.; Liu, J.; Prezhdo, O. V. Unravelling the Effects of (121) Kim, W.; Jung, M. S.; Lee, S.; Choi, Y. J.; Kim, J. K.; Chai, S.
Grain Boundary and Chemical Doping on Electron−Hole Recombi- U.; Kim, W.; Choi, D. G.; Ahn, H.; Cho, J. H.; et al. Oriented Grains
nation in CH3NH3PbI3 Perovskite by Time-Domain Atomistic with Preferred Low-Angle Grain Boundaries in Halide Perovskite
Simulation. J. Am. Chem. Soc. 2016, 138, 3884−3890. Films by Pressure-Induced Crystallization. Adv. Energy Mater. 2018, 8,
(104) Liu, J.; Prezhdo, O. V. Chlorine Doping Reduces Electron− 1702369.
Hole Recombination in Lead Iodide Perovskites: Time-Domain Ab (122) de Quilettes, D. W.; Vorpahl, S. M.; Stranks, S. D.; Nagaoka,
Initio Analysis. J. Phys. Chem. Lett. 2015, 6, 4463−4469. H.; Eperon, G. E.; Ziffer, M. E.; Snaith, H. J.; Ginger, D. S. Impact of
(105) Jankowska, J.; Long, R.; Prezhdo, O. V. Quantum Dynamics of Microstructure on Local Carrier Lifetime in Perovskite Solar Cells.
Photogenerated Charge Carriers in Hybrid Perovskites: Dopants, Science 2015, 348, 683−686.
Grain Boundaries, Electric Order, and Other Realistic Aspects. ACS (123) deQuilettes, D. W.; Jariwala, S.; Burke, S.; Ziffer, M. E.; Wang,
Energy Lett. 2017, 2, 1588−1597. J. T.-W.; Snaith, H. J.; Ginger, D. S. Tracking Photoexcited Carriers in
(106) Wang, Y.; Fang, W.-H.; Long, R.; Prezhdo, O. V. Symmetry Hybrid Perovskite Semiconductors: Trap-Dominated Spatial Hetero-
Breaking at Mapbi3 Perovskite Grain Boundaries Suppresses Charge geneity and Diffusion. ACS Nano 2017, 11, 11488−11496.
Recombination: Time-Domain Ab Initio Analysis. J. Phys. Chem. Lett. (124) Ham, S.; Choi, Y. J.; Lee, J.-W.; Park, N.-G.; Kim, D. Impact
2019, 10, 1617−1623. of Excess CH3NH3I on Free Carrier Dynamics in High-Performance
(107) Yun, J. S.; Ho-Baillie, A.; Huang, S.; Woo, S. H.; Heo, Y.; Nonstoichiometric Perovskites. J. Phys. Chem. C 2017, 121, 3143−
Seidel, J.; Huang, F.; Cheng, Y.-B.; Green, M. A. Benefit of Grain 3148.
Boundaries in Organic−Inorganic Halide Planar Perovskite Solar (125) Kieslich, G.; Sun, S.; Cheetham, A. K. Solid-State Principles
Cells. J. Phys. Chem. Lett. 2015, 6, 875−880. Applied to Organic−Inorganic Perovskites: New Tricks for an Old
(108) Chen, Q.; Zhou, H.; Song, T.-B.; Luo, S.; Hong, Z.; Duan, H.- Dog. Chem. Sci. 2014, 5, 4712−4715.
S.; Dou, L.; Liu, Y.; Yang, Y. Controllable Self-Induced Passivation of (126) Travis, W.; Glover, E.; Bronstein, H.; Scanlon, D.; Palgrave, R.
Hybrid Lead Iodide Perovskites toward High Performance Solar On the Application of the Tolerance Factor to Inorganic and Hybrid
Cells. Nano Lett. 2014, 14, 4158−4163. Halide Perovskites: A Revised System. Chem. Sci. 2016, 7, 4548−
(109) Shih, M.-C.; Li, S.-S.; Hsieh, C.-H.; Wang, Y.-C.; Yang, H.-D.; 4556.
Chiu, Y.-P.; Chang, C.-S.; Chen, C.-W. Spatially Resolved Imaging on (127) Becker, M.; Klüner, T.; Wark, M. Formation of Hybrid ABX3
Photocarrier Generations and Band Alignments at Perovskite/PbI2 Perovskite Compounds for Solar Cell Application: First-Principles
Heterointerfaces of Perovskite Solar Cells by Light-Modulated Calculations of Effective Ionic Radii and Determination of Tolerance
Scanning Tunneling Microscopy. Nano Lett. 2017, 17, 1154−1160. Factors. Dalton Trans. 2017, 46, 3500−3509.
(110) Chu, Z.; Yang, M.; Schulz, P.; Wu, D.; Ma, X.; Seifert, E.; Sun, (128) Shannon, R. D. Revised Effective Ionic Radii and Systematic
L.; Li, X.; Zhu, K.; Lai, K. Impact of Grain Boundaries on Efficiency Studies of Interatomic Distances in Halides and Chalcogenides. Acta
and Stability of Organic-Inorganic Trihalide Perovskites. Nat. Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr. 1976,
Commun. 2017, 8, 2230. 32, 751−767.
(111) Seol, D.; Jeong, A.; Han, M. H.; Seo, S.; Yoo, T. S.; Choi, W. (129) Amat, A.; Mosconi, E.; Ronca, E.; Quarti, C.; Umari, P.;
S.; Jung, H. S.; Shin, H.; Kim, Y. Origin of Hysteresis in CH3NH3PbI3 Nazeeruddin, M. K.; Grätzel, M.; De Angelis, F. Grätzel, M.; De
Perovskite Thin Films. Adv. Funct. Mater. 2017, 27, 1701924. Angelis, F. Cation-Induced Band-Gap Tuning in Organohalide
(112) Nah, S.; Spokoyny, B.; Jiang, X.; Stoumpos, C.; Soe, C.; Perovskites: Interplay of Spin−Orbit Coupling and Octahedra Tilting.
Kanatzidis, M. G.; Harel, E. Transient Sub-Bandgap States in Halide Nano Lett. 2014, 14, 3608−3616.
Perovskite Thin Films. Nano Lett. 2018, 18, 827−831. (130) Giorgi, G.; Fujisawa, J.-I.; Segawa, H.; Yamashita, K. Small
(113) Simpson, M. J.; Doughty, B.; Yang, B.; Xiao, K.; Ma, Y.-Z. Photocarrier Effective Masses Featuring Ambipolar Transport in
Imaging Electronic Trap States in Perovskite Thin Films with Methylammonium Lead Iodide Perovskite: A Density Functional
Combined Fluorescence and Femtosecond Transient Absorption Analysis. J. Phys. Chem. Lett. 2013, 4, 4213−4216.
Microscopy. J. Phys. Chem. Lett. 2016, 7, 1725−1731. (131) Leguy, A. M.; Frost, J. M.; McMahon, A. P.; Sakai, V. G.;
(114) Draguta, S.; Thakur, S.; Morozov, Y. V.; Wang, Y.; Manser, J. Kockelmann, W.; Law, C.; Li, X.; Foglia, F.; Walsh, A.; O’regan, B. C.;
S.; Kamat, P. V.; Kuno, M. Spatially Non-Uniform Trap State et al. The Dynamics of Methylammonium Ions in Hybrid Organic−
Densities in Solution-Processed Hybrid Perovskite Thin Films. J. Phys. Inorganic Perovskite Solar Cells. Nat. Commun. 2015, 6, 7124.
Chem. Lett. 2016, 7, 715−721. (132) Zhu, H.; Miyata, K.; Fu, Y.; Wang, J.; Joshi, P. P.; Niesner, D.;
(115) Tian, W.; Cui, R.; Leng, J.; Liu, J.; Li, Y.; Zhao, C.; Zhang, J.; Williams, K. W.; Jin, S.; Zhu, X.-Y. Screening in Crystalline Liquids
Deng, W.; Lian, T.; Jin, S. Limiting Perovskite Solar Cell Performance Protects Energetic Carriers in Hybrid Perovskites. Science 2016, 353,
by Heterogeneous Carrier Extraction. Angew. Chem., Int. Ed. 2016, 55, 1409−1413.
13067−13071. (133) Chen, T.; Chen, W.-L.; Foley, B. J.; Lee, J.; Ruff, J. P.; Ko, J.
(116) Yang, M.; Zeng, Y.; Li, Z.; Kim, D. H.; Jiang, C.-S.; van de P.; Brown, C. M.; Harriger, L. W.; Zhang, D.; Park, C.; et al. Origin of
Lagemaat, J.; Zhu, K. Do Grain Boundaries Dominate Non-Radiative Long Lifetime of Band-Edge Charge Carriers in Organic−Inorganic
Recombination in CH3NH3PbI3 Perovskite Thin Films? Phys. Chem. Lead Iodide Perovskites. Proc. Natl. Acad. Sci. U. S. A. 2017, 114,
Chem. Phys. 2017, 19, 5043−5050. 7519−7524.
(117) Yang, Y.; Yang, M.; Moore, D. T.; Yan, Y.; Miller, E. M.; Zhu, (134) Ma, F.; Li, J.; Li, W.; Lin, N.; Wang, L.; Qiao, J. Stable α/δ
K.; Beard, M. C. Top and Bottom Surfaces Limit Carrier Lifetime in Phase Junction of Formamidinium Lead Iodide Perovskites for
Lead Iodide Perovskite Films. Nat. Energy 2017, 2, 16207. Enhanced near-Infrared Emission. Chem. Sci. 2017, 8, 800−805.
(118) Snaider, J. M.; Guo, Z.; Wang, T.; Yang, M.; Yuan, L.; Zhu, K.; (135) Koh, T. M.; Fu, K.; Fang, Y.; Chen, S.; Sum, T.; Mathews, N.;
Huang, L. Ultrafast Imaging of Carrier Transport across Grain Mhaisalkar, S. G.; Boix, P. P.; Baikie, T. Formamidinium-Containing
Boundaries in Hybrid Perovskite Thin Films. ACS Energy Lett. 2018, Metal-Halide: An Alternative Material for near-Ir Absorption
3, 1402−1408. Perovskite Solar Cells. J. Phys. Chem. C 2014, 118, 16458−16462.
(119) Yin, J.; Cortecchia, D.; Krishna, A.; Chen, S.; Mathews, N.; (136) Eperon, G. E.; Stranks, S. D.; Menelaou, C.; Johnston, M. B.;
Grimsdale, A. C.; Soci, C. Interfacial Charge Transfer Anisotropy in Herz, L. M.; Snaith, H. J. Formamidinium Lead Trihalide: A Broadly
Polycrystalline Lead Iodide Perovskite Films. J. Phys. Chem. Lett. Tunable Perovskite for Efficient Planar Heterojunction Solar Cells.
2015, 6, 1396−1402. Energy Environ. Sci. 2014, 7, 982−988.

AS https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

(137) Weller, M. T.; Weber, O. J.; Frost, J. M.; Walsh, A. Cubic voltaics for Enhanced Solar-Light Harvesting. Angew. Chem., Int. Ed.
Perovskite Structure of Black Formamidinium Lead Iodide, Α-[Hc 2014, 53, 3151−3157.
(Nh2) 2] Pbi3, at 298 K. J. Phys. Chem. Lett. 2015, 6, 3209−3212. (155) Zheng, X.; Wu, C.; Jha, S. K.; Li, Z.; Zhu, K.; Priya, S.
(138) Chen, T.; Foley, B. J.; Park, C.; Brown, C. M.; Harriger, L. W.; Improved Phase Stability of Formamidinium Lead Triiodide Perov-
Lee, J.; Ruff, J.; Yoon, M.; Choi, J. J.; Lee, S.-H. Entropy-Driven skite by Strain Relaxation. ACS Energy Lett. 2016, 1, 1014−1020.
Structural Transition and Kinetic Trapping in Formamidinium Lead (156) Zhao, Y.; Tan, H.; Yuan, H.; Yang, Z.; Fan, J. Z.; Kim, J.;
Iodide Perovskite. Sci. Adv. 2016, 2, No. e1601650. Voznyy, O.; Gong, X.; Quan, L. N.; Tan, C. S.; et al. Perovskite
(139) Lee, J. W.; Kim, D. H.; Kim, H. S.; Seo, S. W.; Cho, S. M.; Seeding Growth of Formamidinium-Lead-Iodide-Based Perovskites
Park, N. G. Formamidinium and Cesium Hybridization for Photo-and for Efficient and Stable Solar Cells. Nat. Commun. 2018, 9, 1607.
Moisture-Stable Perovskite Solar Cell. Adv. Energy Mater. 2015, 5, (157) Li, Z.; Yang, M.; Park, J.-S.; Wei, S.-H.; Berry, J. J.; Zhu, K.
1501310. Stabilizing Perovskite Structures by Tuning Tolerance Factor:
(140) Lin, J.; Lai, M.; Dou, L.; Kley, C. S.; Chen, H.; Peng, F.; Sun, Formation of Formamidinium and Cesium Lead Iodide Solid-State
J.; Lu, D.; Hawks, S. A.; Xie, C.; et al. Thermochromic Halide Alloys. Chem. Mater. 2016, 28, 284−292.
Perovskite Solar Cells. Nat. Mater. 2018, 17, 261. (158) Yi, C.; Luo, J.; Meloni, S.; Boziki, A.; Ashari-Astani, N.;
(141) Kubicki, D. J.; Prochowicz, D.; Hofstetter, A.; Saski, M.; Grätzel, C.; Zakeeruddin, S. M.; Röthlisberger, U.; Grätzel, M.
Yadav, P.; Bi, D.; Pellet, N.; Lewiński, J.; Zakeeruddin, S. M.; Grätzel, Entropic Stabilization of Mixed A-Cation ABX3 Metal Halide
M. Formation of Stable Mixed Guanidinium−Methylammonium Perovskites for High Performance Perovskite Solar Cells. Energy
Phases with Exceptionally Long Carrier Lifetimes for High-Efficiency Environ. Sci. 2016, 9, 656−662.
Lead Iodide-Based Perovskite Photovoltaics. J. Am. Chem. Soc. 2018, (159) Park, I. J.; Seo, S.; Park, M. A.; Lee, S.; Kim, D. H.; Zhu, K.;
140, 3345−3351. Shin, H.; Kim, J. Y. Effect of Rubidium Incorporation on the
(142) Kubicki, D. J.; Prochowicz, D.; Hofstetter, A.; Péchy, P. t.; Structural, Electrical, and Photovoltaic Properties of Methylammo-
Zakeeruddin, S. M.; Grätzel, M.; Emsley, L. Cation Dynamics in nium Lead Iodide-Based Perovskite Solar Cells. ACS Appl. Mater.
Mixed-Cation (MA)X(FA)1−XPbI3 Hybrid Perovskites from Solid- Interfaces 2017, 9, 41898−41905.
State Nmr. J. Am. Chem. Soc. 2017, 139, 10055−10061. (160) Saliba, M.; Matsui, T.; Seo, J.-Y.; Domanski, K.; Correa-Baena,
(143) Juarez-Perez, E. J.; Ono, L. K.; Qi, Y. Thermal Degradation of J.-P.; Nazeeruddin, M. K.; Zakeeruddin, S. M.; Tress, W.; Abate, A.;
Formamidinium Based Lead Halide Perovskites into Sym-Triazine Hagfeldt, A.; et al. Cesium-Containing Triple Cation Perovskite Solar
and Hydrogen Cyanide Observed by Coupled Thermogravimetry- Cells: Improved Stability, Reproducibility and High Efficiency. Energy
Mass Spectrometry Analysis. J. Mater. Chem. A 2019, 7, 16912− Environ. Sci. 2016, 9, 1989−1997.
16919. (161) Saliba, M.; Matsui, T.; Domanski, K.; Seo, J.-Y.;
(144) Eperon, G. E.; Paterno, G. M.; Sutton, R. J.; Zampetti, A.; Ummadisingu, A.; Zakeeruddin, S. M.; Correa-Baena, J.-P.; Tress,
Haghighirad, A. A.; Cacialli, F.; Snaith, H. J. Inorganic Caesium Lead W. R.; Abate, A.; Hagfeldt, A.; et al. Incorporation of Rubidium
Iodide Perovskite Solar Cells. J. Mater. Chem. A 2015, 3, 19688− Cations into Perovskite Solar Cells Improves Photovoltaic Perform-
19695. ance. Science 2016, 354, 206−209.
(145) Stoumpos, C. C.; Malliakas, C. D.; Kanatzidis, M. G. (162) De Marco, N.; Zhou, H.; Chen, Q.; Sun, P.; Liu, Z.; Meng, L.;
Semiconducting Tin and Lead Iodide Perovskites with Organic Yao, E.-P.; Liu, Y.; Schiffer, A.; Yang, Y. Guanidinium: A Route to
Cations: Phase Transitions, High Mobilities, and near-Infrared Enhanced Carrier Lifetime and Open-Circuit Voltage in Hybrid
Photoluminescent Properties. Inorg. Chem. 2013, 52, 9019−9038. Perovskite Solar Cells. Nano Lett. 2016, 16, 1009−1016.
(146) Swarnkar, A.; Marshall, A. R.; Sanehira, E. M.; Chernomordik, (163) Jodlowski, A. D.; Roldán-Carmona, C.; Grancini, G.; Salado,
B. D.; Moore, D. T.; Christians, J. A.; Chakrabarti, T.; Luther, J. M. M.; Ralaiarisoa, M.; Ahmad, S.; Koch, N.; Camacho, L.; De Miguel,
Quantum Dot−Induced Phase Stabilization of Α-Cspbi3 Perovskite G.; Nazeeruddin, M. K. Large Guanidinium Cation Mixed with
for High-Efficiency Photovoltaics. Science 2016, 354, 92−95. Methylammonium in Lead Iodide Perovskites for 19% Efficient Solar
(147) Sanehira, E. M.; Marshall, A. R.; Christians, J. A.; Harvey, S. Cells. Nat. Energy 2017, 2, 972.
P.; Ciesielski, P. N.; Wheeler, L. M.; Schulz, P.; Lin, L. Y.; Beard, M. (164) Chen, H.; Wei, Q.; Saidaminov, M. I.; Wang, F.; Johnston, A.;
C.; Luther, J. M. Enhanced Mobility CsPbI3 Quantum Dot Arrays for Hou, Y.; Peng, Z.; Xu, K.; Zhou, W.; Liu, Z.; et al. Efficient and Stable
Record-Efficiency, High-Voltage Photovoltaic Cells. Sci. Adv. 2017, 3, Inverted Perovskite Solar Cells Incorporating Secondary Amines. Adv.
No. eaao4204. Mater. 2019, 31, 1903559.
(148) Xue, J.; Lee, J.-W.; Dai, Z.; Wang, R.; Nuryyeva, S.; Liao, M. (165) Ferdani, D.; Pering, S.; Ghosh, D.; Kubiak, P.; Walker, A.;
E.; Chang, S.-Y.; Meng, L.; Meng, D.; Sun, P.; et al. Surface Ligand Lewis, S. E.; Johnson, A. L.; Baker, P. J.; Islam, S.; Cameron, P. J.
Management for Stable FAPbI3 Perovskite Quantum Dot Solar Cells. Partial Cation Substitution Reduces Iodide Ion Transport in Lead
Joule 2018, 2, 1866−1878. Iodide Perovskite Solar Cells. Energy Environ. Sci. 2019, 12, 2264−
(149) Wang, Q.; Zheng, X.; Deng, Y.; Zhao, J.; Chen, Z.; Huang, J. 2272.
Stabilizing the Α-Phase of Cspbi3 Perovskite by Sulfobetaine (166) Lee, J.-W.; Kim, S.-G.; Yang, J.-M.; Yang, Y.; Park, N.-G.
Zwitterions in One-Step Spin-Coating Films. Joule 2017, 1, 371−382. Verification and Mitigation of Ion Migration in Perovskite Solar Cells.
(150) Zhang, T.; Dar, M. I.; Li, G.; Xu, F.; Guo, N.; Grätzel, M.; APL Mater. 2019, 7, 041111.
Zhao, Y. Bication Lead Iodide 2d Perovskite Component to Stabilize (167) Zheng, X.; Hou, Y.; Bao, C.; Yin, J.; Yuan, F.; Huang, Z.;
Inorganic Α-Cspbi3 Perovskite Phase for High-Efficiency Solar Cells. Song, K.; Liu, J.; Troughton, J.; Gasparini, N.; et al. Managing Grains
Sci. Adv. 2017, 3, No. e1700841. and Interfaces Via Ligand Anchoring Enables 22.3%-Efficiency
(151) Wang, Y.; Dar, M. I.; Ono, L. K.; Zhang, T.; Kan, M.; Li, Y.; Inverted Perovskite Solar Cells. Nat. Energy 2020, 5, 131−140.
Zhang, L.; Wang, X.; Yang, Y.; Gao, X.; et al. Thermodynamically (168) Saidaminov, M. I.; Williams, K.; Wei, M.; Johnston, A.;
Stabilized β-CsPbI3−Based Perovskite Solar Cells with Efficiencies> Quintero-Bermudez, R.; Vafaie, M.; Pina, J. M.; Proppe, A. H.; Hou,
18%. Science 2019, 365, 591−595. Y.; Walters, G.; et al. Multi-Cation Perovskites Prevent Carrier
(152) Wang, Y.; Zhang, T.; Kan, M.; Zhao, Y. Bifunctional Reflection from Grain Surfaces. Nat. Mater. 2020, 19, 412−418.
Stabilization of All-Inorganic Α-CsPbI3 Perovskite for 17% Efficiency (169) Tan, S.; Yavuz, I.; De Marco, N.; Huang, T.; Lee, S. J.; Choi,
Photovoltaics. J. Am. Chem. Soc. 2018, 140, 12345−12348. C. S.; Wang, M.; Nuryyeva, S.; Wang, R.; Zhao, Y.; et al. Steric
(153) Binek, A.; Hanusch, F. C.; Docampo, P.; Bein, T. Stabilization Impediment of Ion Migration Contributes to Improved Operational
of the Trigonal High-Temperature Phase of Formamidinium Lead Stability of Perovskite Solar Cells. Adv. Mater. 2020, 32, 1906995.
Iodide. J. Phys. Chem. Lett. 2015, 6, 1249−1253. (170) Kumar, M. H.; Dharani, S.; Leong, W. L.; Boix, P. P.;
(154) Pellet, N.; Gao, P.; Gregori, G.; Yang, T. Y.; Nazeeruddin, M. Prabhakar, R. R.; Baikie, T.; Shi, C.; Ding, H.; Ramesh, R.; Asta, M.;
K.; Maier, J.; Grätzel, M. Mixed-Organic-Cation Perovskite Photo- et al. Lead-Free Halide Perovskite Solar Cells with High Photo-

AT https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

currents Realized through Vacancy Modulation. Adv. Mater. 2014, 26, (187) Kim, M.-c.; Kim, B. J.; Son, D.-Y.; Park, N.-G.; Jung, H. S.;
7122−7127. Choi, M. Observation of Enhanced Hole Extraction in Br
(171) Hao, F.; Stoumpos, C. C.; Cao, D. H.; Chang, R. P.; Concentration Gradient Perovskite Materials. Nano Lett. 2016, 16,
Kanatzidis, M. G. Lead-Free Solid-State Organic−Inorganic Halide 5756−5763.
Perovskite Solar Cells. Nat. Nat. Photonics 2014, 8, 489−494. (188) Slotcavage, D. J.; Karunadasa, H. I.; McGehee, M. D. Light-
(172) Liao, W.; Zhao, D.; Yu, Y.; Grice, C. R.; Wang, C.; Cimaroli, Induced Phase Segregation in Halide-Perovskite Absorbers. ACS
A. J.; Schulz, P.; Meng, W.; Zhu, K.; Xiong, R. G.; et al. Lead-Free Energy Lett. 2016, 1, 1199−1205.
Inverted Planar Formamidinium Tin Triiodide Perovskite Solar Cells (189) McMeekin, D. P.; Sadoughi, G.; Rehman, W.; Eperon, G. E.;
Achieving Power Conversion Efficiencies up to 6.22%. Adv. Mater. Saliba, M.; Hörantner, M. T.; Haghighirad, A.; Sakai, N.; Korte, L.;
2016, 28, 9333−9340. Rech, B.; et al. A Mixed-Cation Lead Mixed-Halide Perovskite
(173) Jiang, X.; Wang, F.; Wei, Q.; Li, H.; Shang, Y.; Zhou, W.; Absorber for Tandem Solar Cells. Science 2016, 351, 151−155.
Wang, C.; Cheng, P.; Chen, Q.; Chen, L.; et al. Ultra-High Open- (190) Rehman, W.; McMeekin, D. P.; Patel, J. B.; Milot, R. L.;
Circuit Voltage of Tin Perovskite Solar Cells Via an Electron Johnston, M. B.; Snaith, H. J.; Herz, L. M. Photovoltaic Mixed-Cation
Transporting Layer Design. Nat. Commun. 2020, 11, 1245. Lead Mixed-Halide Perovskites: Links between Crystallinity, Photo-
(174) Zhao, D.; Yu, Y.; Wang, C.; Liao, W.; Shrestha, N.; Grice, C. Stability and Electronic Properties. Energy Environ. Sci. 2017, 10,
R.; Cimaroli, A. J.; Guan, L.; Ellingson, R. J.; Zhu, K.; et al. Low- 361−369.
Bandgap Mixed Tin−Lead Iodide Perovskite Absorbers with Long (191) Yu, H.; Wang, F.; Xie, F.; Li, W.; Chen, J.; Zhao, N. The Role
Carrier Lifetimes for All-Perovskite Tandem Solar Cells. Nat. Energy of Chlorine in the Formation Process of “Ch3nh3pbi3-Xclx”
2017, 2, 17018. Perovskite. Adv. Funct. Mater. 2014, 24, 7102−7108.
(175) Zhao, D.; Chen, C.; Wang, C.; Junda, M. M.; Song, Z.; Grice, (192) Williams, S. T.; Zuo, F.; Chueh, C.-C.; Liao, C.-Y.; Liang, P.-
C. R.; Yu, Y.; Li, C.; Subedi, B.; Podraza, N. J.; et al. Efficient Two- W.; Jen, A. K.-Y. Role of Chloride in the Morphological Evolution of
Terminal All-Perovskite Tandem Solar Cells Enabled by High-Quality Organo-Lead Halide Perovskite Thin Films. ACS Nano 2014, 8,
Low-Bandgap Absorber Layers. Nat. Energy 2018, 3, 1093. 10640−10654.
(176) Tong, J.; Song, Z.; Kim, D. H.; Chen, X.; Chen, C.; (193) Smith, I. C.; Hoke, E. T.; Solis-Ibarra, D.; McGehee, M. D.;
Palmstrom, A. F.; Ndione, P. F.; Reese, M. O.; Dunfield, S. P.; Reid, Karunadasa, H. I. A Layered Hybrid Perovskite Solar-Cell Absorber
O. G.; et al. Carrier Lifetimes of> 1 Μs in Sn-Pb Perovskites Enable with Enhanced Moisture Stability. Angew. Chem., Int. Ed. 2014, 53,
Efficient All-Perovskite Tandem Solar Cells. Science 2019, 364, 475− 11232−11235.
479. (194) Stoumpos, C. C.; Soe, C. M. M.; Tsai, H.; Nie, W.; Blancon,
(177) Yang, Z.; Yu, Z.; Wei, H.; Xiao, X.; Ni, Z.; Chen, B.; Deng, Y.; J.-C.; Cao, D. H.; Liu, F.; Traoré, B.; Katan, C.; Even, J.; et al. High
Habisreutinger, S. N.; Chen, X.; Wang, K. Enhancing Electron Members of the 2d Ruddlesden-Popper Halide Perovskites: Synthesis,
Optical Properties, and Solar Cells of
Diffusion Length in Narrow-Bandgap Perovskites for Efficient
(CH3(CH2)3NH3)2(CH3NH3)4Pb5I16. Chem. 2017, 2, 427−440.
Monolithic Perovskite Tandem Solar Cells. Nat. Commun. 2019, 10,
(195) Cao, D. H.; Stoumpos, C. C.; Farha, O. K.; Hupp, J. T.;
4498.
Kanatzidis, M. G. 2D Homologous Perovskites as Light-Absorbing
(178) Lin, R.; Xiao, K.; Qin, Z.; Han, Q.; Zhang, C.; Wei, M.;
Materials for Solar Cell Applications. J. Am. Chem. Soc. 2015, 137,
Saidaminov, M. I.; Gao, Y.; Xu, J.; Xiao, M.; et al. Monolithic All-
7843−7850.
Perovskite Tandem Solar Cells with 24.8% Efficiency Exploiting
(196) Tsai, H.; Nie, W.; Blancon, J.-C.; Stoumpos, C. C.; Asadpour,
Comproportionation to Suppress Sn(II) Oxidation in Precursor Ink. R.; Harutyunyan, B.; Neukirch, A. J.; Verduzco, R.; Crochet, J. J.;
Nat. Energy 2019, 4, 864−873. Tretiak, S.; et al. High-Efficiency Two-Dimensional Ruddlesden−
(179) Buin, A.; Comin, R.; Xu, J.; Ip, A. H.; Sargent, E. H. Halide- Popper Perovskite Solar Cells. Nature 2016, 536, 312−316.
Dependent Electronic Structure of Organolead Perovskite Materials. (197) Quan, L. N.; Yuan, M.; Comin, R.; Voznyy, O.; Beauregard, E.
Chem. Mater. 2015, 27, 4405−4412. M.; Hoogland, S.; Buin, A.; Kirmani, A. R.; Zhao, K.; Amassian, A.;
(180) Maculan, G.; Sheikh, A. D.; Abdelhady, A. L.; Saidaminov, M. et al. Ligand-Stabilized Reduced-Dimensionality Perovskites. J. Am.
I.; Haque, M. A.; Murali, B.; Alarousu, E.; Mohammed, O. F.; Wu, T.; Chem. Soc. 2016, 138, 2649−2655.
Bakr, O. M. CH3NH3PbCl3 Single Crystals: Inverse Temperature (198) Grancini, G.; Roldán-Carmona, C.; Zimmermann, I.;
Crystallization and Visible-Blind Uv-Photodetector. J. Phys. Chem. Mosconi, E.; Lee, X.; Martineau, D.; Narbey, S.; Oswald, F.; De
Lett. 2015, 6, 3781−3786. Angelis, F.; Graetzel, M.; et al. One-Year Stable Perovskite Solar Cells
(181) Pauling, L. The Nature of the Chemical Bond. Iv. The Energy by 2d/3d Interface Engineering. Nat. Commun. 2017, 8, 15684.
of Single Bonds and the Relative Electronegativity of Atoms. J. Am. (199) Blancon, J.-C.; Tsai, H.; Nie, W.; Stoumpos, C. C.; Pedesseau,
Chem. Soc. 1932, 54, 3570−3582. L.; Katan, C.; Kepenekian, M.; Soe, C. M. M.; Appavoo, K.; Sfeir, M.
(182) Nagabhushana, G.; Shivaramaiah, R.; Navrotsky, A. Direct Y.; et al. Extremely Efficient Internal Exciton Dissociation through
Calorimetric Verification of Thermodynamic Instability of Lead Edge States in Layered 2d Perovskites. Science 2017, 355, 1288−1292.
Halide Hybrid Perovskites. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, (200) Hu, J.; Oswald, I. W.; Stuard, S. J.; Nahid, M. M.; Zhou, N.;
7717−7721. Williams, O. F.; Guo, Z.; Yan, L.; Hu, H.; Chen, Z. Synthetic Control
(183) Noh, J. H.; Im, S. H.; Heo, J. H.; Mandal, T. N.; Seok, S. I. over Orientational Degeneracy of Spacer Cations Enhances Solar Cell
Chemical Management for Colorful, Efficient, and Stable Inorganic− Efficiency in Two-Dimensional Perovskites. Nat. Commun. 2019, 10,
Organic Hybrid Nanostructured Solar Cells. Nano Lett. 2013, 13, 1276.
1764−1769. (201) Lee, J.-W.; Dai, Z.; Han, T.-H.; Choi, C.; Chang, S.-Y.; Lee, S.-
(184) Eames, C.; Frost, J. M.; Barnes, P. R.; O’regan, B. C.; Walsh, J.; De Marco, N.; Zhao, H.; Sun, P.; Huang, Y. 2d Perovskite
A.; Islam, M. S. Ionic Transport in Hybrid Lead Iodide Perovskite Stabilized Phase-Pure Formamidinium Perovskite Solar Cells. Nat.
Solar Cells. Nat. Commun. 2015, 6, 7497. Commun. 2018, 9, 3021.
(185) Xing, J.; Wang, Q.; Dong, Q.; Yuan, Y.; Fang, Y.; Huang, J. (202) Han, T. H.; Lee, J. W.; Choi, Y. J.; Choi, C.; Tan, S.; Lee, S. J.;
Ultrafast Ion Migration in Hybrid Perovskite Polycrystalline Thin Zhao, Y.; Huang, Y.; Kim, D.; Yang, Y. Surface-2d/Bulk-3d
Films under Light and Suppression in Single Crystals. Phys. Chem. Heterophased Perovskite Nanograins for Long-Term-Stable Light-
Chem. Phys. 2016, 18, 30484−30490. Emitting Diodes. Adv. Mater. 2020, 32, 1905674.
(186) Hoke, E. T.; Slotcavage, D. J.; Dohner, E. R.; Bowring, A. R.; (203) Wang, Z.; Lin, Q.; Chmiel, F. P.; Sakai, N.; Herz, L. M.;
Karunadasa, H. I.; McGehee, M. D. Reversible Photo-Induced Trap Snaith, H. J. Efficient Ambient-Air-Stable Solar Cells with 2d−3d
Formation in Mixed-Halide Hybrid Perovskites for Photovoltaics. Heterostructured Butylammonium-Caesium-Formamidinium Lead
Chem. Sci. 2015, 6, 613−617. Halide Perovskites. Nat. Energy 2017, 2, 17135.

AU https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

(204) Thanh, N. T.; Maclean, N.; Mahiddine, S. Mechanisms of Hole-Transporting Material for Highly Efficient and Stable Perovskite
Nucleation and Growth of Nanoparticles in Solution. Chem. Rev. Solar Cells. Nat. Energy 2018, 3, 682−689.
2014, 114, 7610−7630. (222) Paek, S.; Schouwink, P.; Athanasopoulou, E. N.; Cho, K.;
(205) Kwon, S. G.; Hyeon, T. Formation Mechanisms of Uniform Grancini, G.; Lee, Y.; Zhang, Y.; Stellacci, F.; Nazeeruddin, M. K.;
Nanocrystals Via Hot-Injection and Heat-up Methods. Small 2011, 7, Gao, P. From Nano-to Micrometer Scale: The Role of Antisolvent
2685−2702. Treatment on High Performance Perovskite Solar Cells. Chem. Mater.
(206) Lee, J. W.; Lee, D. K.; Jeong, D. N.; Park, N. G. Control of 2017, 29, 3490−3498.
Crystal Growth toward Scalable Fabrication of Perovskite Solar Cells. (223) Nie, W.; Tsai, H.; Asadpour, R.; Blancon, J.-C.; Neukirch, A.
Adv. Funct. Mater. 2019, 29, 1807047. J.; Gupta, G.; Crochet, J. J.; Chhowalla, M.; Tretiak, S.; Alam, M. A.;
(207) Im, J.-H.; Kim, H.-S.; Park, N.-G. Morphology-Photovoltaic et al. High-Efficiency Solution-Processed Perovskite Solar Cells with
Property Correlation in Perovskite Solar Cells: One-Step Versus Two- Millimeter-Scale Grains. Science 2015, 347, 522−525.
Step Deposition of CH3NH3PbI3. APL Mater. 2014, 2, 081510. (224) Deng, Y.; Zheng, X.; Bai, Y.; Wang, Q.; Zhao, J.; Huang, J.
(208) Eperon, G. E.; Burlakov, V. M.; Docampo, P.; Goriely, A.; Surfactant-Controlled Ink Drying Enables High-Speed Deposition of
Snaith, H. J. Morphological Control for High Performance, Solution- Perovskite Films for Efficient Photovoltaic Modules. Nat. Energy
Processed Planar Heterojunction Perovskite Solar Cells. Adv. Funct. 2018, 3, 560.
Mater. 2014, 24, 151−157. (225) Huang, F.; Dkhissi, Y.; Huang, W.; Xiao, M.; Benesperi, I.;
(209) Xiao, M.; Huang, F.; Huang, W.; Dkhissi, Y.; Zhu, Y.; Rubanov, S.; Zhu, Y.; Lin, X.; Jiang, L.; Zhou, Y.; et al. Gas-Assisted
Etheridge, J.; Gray-Weale, A.; Bach, U.; Cheng, Y. B.; Spiccia, L. A Preparation of Lead Iodide Perovskite Films Consisting of a
Fast Deposition-Crystallization Procedure for Highly Efficient Lead Monolayer of Single Crystalline Grains for High Efficiency Planar
Iodide Perovskite Thin-Film Solar Cells. Angew. Chem., Int. Ed. 2014, Solar Cells. Nano Energy 2014, 10, 10−18.
53, 9898−9903. (226) Hu, H.; Ren, Z.; Fong, P. W. K.; Qin, M.; Liu, D.; Lei, D.; Lu,
(210) Yu, Y.; Yang, S.; Lei, L.; Cao, Q.; Shao, J.; Zhang, S.; Liu, Y. X.; Li, G. Room-Temperature Meniscus Coating of > 20% Perovskite
Ultrasmooth Perovskite Film Via Mixed Anti-Solvent Strategy with Solar Cells: A Film Formation Mechanism Investigation. Adv. Funct.
Improved Efficiency. ACS Appl. Mater. Interfaces 2017, 9, 3667−3676. Mater. 2019, 29, 1900092.
(211) Yin, M.; Xie, F.; Chen, H.; Yang, X.; Ye, F.; Bi, E.; Wu, Y.; Cai, (227) Ng, A.; Ren, Z.; Hu, H.; Fong, P. W.; Shen, Q.; Cheung, S. H.;
M.; Han, L. Annealing-Free Perovskite Films by Instant Crystal- Qin, P.; Lee, J. W.; Djurišić, A. B.; So, S. K. A Cryogenic Process for
lization for Efficient Solar Cells. J. Mater. Chem. A 2016, 4, 8548− Antisolvent-Free High-Performance Perovskite Solar Cells. Adv.
8553. Mater. 2018, 30, 1804402.
(212) Zhao, P.; Kim, B. J.; Ren, X.; Lee, D. G.; Bang, G. J.; Jeon, J. (228) Li, X.; Bi, D.; Yi, C.; Décoppet, J.-D.; Luo, J.; Zakeeruddin, S.
B.; Kim, W. B.; Jung, H. S. Antisolvent with an Ultrawide Processing M.; Hagfeldt, A.; Grätzel, M. A Vacuum Flash−Assisted Solution
Window for the One-Step Fabrication of Efficient and Large-Area Process for High-Efficiency Large-Area Perovskite Solar Cells. Science
2016, 353, 58−62.
Perovskite Solar Cells. Adv. Mater. 2018, 30, 1802763.
(229) Ding, B.; Li, Y.; Huang, S.-Y.; Chu, Q.-Q.; Li, C.-X.; Li, C.-J.;
(213) Han, G. S.; Kim, J.; Bae, S.; Han, S.-H.; Kim, Y. J.; Gong, O.
Yang, G.-J. Material Nucleation/Growth Competition Tuning
Y.; Lee, P.; Ko, M. J.; Jung, H. S. Spin Coating Process for 10 cm × 10
Towards Highly Reproducible Planar Perovskite Solar Cells with
cm Perovskite Solar Modules Enabled by Self-Assembly of SnO2
Efficiency Exceeding 20%. J. Mater. Chem. A 2017, 5, 6840−6848.
Nanocolloids. ACS Energy Lett. 2019, 4, 1845−1851.
(230) Ummadisingu, A.; Grätzel, M. Revealing the Detailed Path of
(214) Noel, N. K.; Habisreutinger, S. N.; Wenger, B.; Klug, M. T.;
Sequential Deposition for Metal Halide Perovskite Formation. Sci.
Hörantner, M. T.; Johnston, M. B.; Nicholas, R. J.; Moore, D. T.;
Adv. 2018, 4, No. e1701402.
Snaith, H. J. A Low Viscosity, Low Boiling Point, Clean Solvent (231) Liang, K.; Mitzi, D. B.; Prikas, M. T. Synthesis and
System for the Rapid Crystallisation of Highly Specular Perovskite Characterization of Organic− Inorganic Perovskite Thin Films
Films. Energy Environ. Sci. 2017, 10, 145−152. Prepared Using a Versatile Two-Step Dipping Technique. Chem.
(215) Xie, L.; Cho, A.-N.; Park, N.-G.; Kim, K. Efficient and Mater. 1998, 10, 403−411.
Reproducible CH3NH3PbI3 Perovskite Layer Prepared Using a Binary (232) Ko, H.-S.; Lee, J.-W.; Park, N.-G. 15.76% Efficiency Perovskite
Solvent Containing a Cyclic Urea Additive. ACS Appl. Mater. Solar Cells Prepared under High Relative Humidity: Importance of
Interfaces 2018, 10, 9390−9397. PbI2 Morphology in Two-Step Deposition of CH3NH3PbI3. J. Mater.
(216) Chen, J.; Kim, S.-G.; Ren, X.; Jung, H. S.; Park, N.-G. Effect of Chem. A 2015, 3, 8808−8815.
Bidentate and Tridentate Additives on the Photovoltaic Performance (233) Yang, W. S.; Noh, J. H.; Jeon, N. J.; Kim, Y. C.; Ryu, S.; Seo,
and Stability of Perovskite Solar Cells. J. Mater. Chem. A 2019, 7, J.; Seok, S. I. High-Performance Photovoltaic Perovskite Layers
4977−4987. Fabricated through Intramolecular Exchange. Science 2015, 348,
(217) Han, T.-H.; Lee, J.-W.; Choi, C.; Tan, S.; Lee, C.; Zhao, Y.; 1234−1237.
Dai, Z.; De Marco, N.; Lee, S.-J.; Bae, S.-H.; et al. Perovskite-Polymer (234) Bai, Y.; Xiao, S.; Hu, C.; Zhang, T.; Meng, X.; Li, Q.; Yang, Y.;
Composite Cross-Linker Approach for Highly-Stable and Efficient Wong, K. S.; Chen, H.; Yang, S. A Pure and Stable Intermediate Phase
Perovskite Solar Cells. Nat. Commun. 2019, 10, 520. Is Key to Growing Aligned and Vertically Monolithic Perovskite
(218) Shin, S. S.; Yeom, E. J.; Yang, W. S.; Hur, S.; Kim, M. G.; Im, Crystals for Efficient Pin Planar Perovskite Solar Cells with High
J.; Seo, J.; Noh, J. H.; Seok, S. I. Colloidally Prepared La-Doped Processibility and Stability. Nano Energy 2017, 34, 58−68.
BaSnO3 Electrodes for Efficient, Photostable Perovskite Solar Cells. (235) Lee, J.-W.; Kim, H.-S.; Park, N.-G. Lewis Acid−Base Adduct
Science 2017, 356, 167−171. Approach for High Efficiency Perovskite Solar Cells. Acc. Chem. Res.
(219) Bi, D.; Tress, W.; Dar, M. I.; Gao, P.; Luo, J.; Renevier, C.; 2016, 49, 311−319.
Schenk, K.; Abate, A.; Giordano, F.; Correa Baena, J.-P.; et al. (236) Lee, J.-W.; Dai, Z.; Lee, C.; Lee, H. M.; Han, T.-H.; De
Efficient Luminescent Solar Cells Based on Tailored Mixed-Cation Marco, N.; Lin, O.; Choi, C. S.; Dunn, B.; Koh, J.; et al. Tuning
Perovskites. Sci. Adv. 2016, 2, No. e1501170. Molecular Interactions for Highly Reproducible and Efficient
(220) Ye, T.; Petrović, M.; Peng, S.; Yoong, J. L. K.; Vijila, C.; Formamidinium Perovskite Solar Cells Via Adduct Approach. J. Am.
Ramakrishna, S. Enhanced Charge Carrier Transport and Device Chem. Soc. 2018, 140, 6317−6324.
Performance through Dual-Cesium Doping in Mixed-Cation Perov- (237) Zhang, Y.; Kim, S.-G.; Lee, D.; Shin, H.; Park, N.-G. Bifacial
skite Solar Cells with near Unity Free Carrier Ratios. ACS Appl. Mater. Stamping for High Efficiency Perovskite Solar Cells. Energy Environ.
Interfaces 2017, 9, 2358−2368. Sci. 2019, 12, 308−321.
(221) Jeon, N. J.; Na, H.; Jung, E. H.; Yang, T.-Y.; Lee, Y. G.; Kim, (238) Kim, H.-S.; Lee, J.-W.; Yantara, N.; Boix, P. P.; Kulkarni, S. A.;
G.; Shin, H.-W.; Seok, S. I.; Lee, J.; Seo, J. A Fluorene-Terminated Mhaisalkar, S.; Grätzel, M.; Park, N.-G. Grätzel, M.; Park, N.-G. High

AV https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Efficiency Solid-State Sensitized Solar Cell-Based on Submicrometer Cells with a Power Conversion Efficiency of 18.3%. Chem. Commun.
Rutile TiO2 Nanorod and CH3NH3PbI3 Perovskite Sensitizer. Nano 2019, 55, 2433−2436.
Lett. 2013, 13, 2412−2417. (257) Anaraki, E. H.; Kermanpur, A.; Steier, L.; Domanski, K.;
(239) Lee, J.-W.; Lee, T.-Y.; Yoo, P. J.; Grätzel, M.; Mhaisalkar, S.; Matsui, T.; Tress, W.; Saliba, M.; Abate, A.; Grätzel, M.; Hagfeldt, A.;
Park, N.-G. Rutile Tio 2-Based Perovskite Solar Cells. J. Mater. Chem. et al. Highly Efficient and Stable Planar Perovskite Solar Cells by
A 2014, 2, 9251−9259. Solution-Processed Tin Oxide. Energy Environ. Sci. 2016, 9, 3128−
(240) Lee, J.-W.; Lee, S. H.; Ko, H.-S.; Kwon, J.; Park, J. H.; Kang, S. 3134.
M.; Ahn, N.; Choi, M.; Kim, J. K.; Park, N.-G. Opto-Electronic (258) Song, S.; Kang, G.; Pyeon, L.; Lim, C.; Lee, G.-Y.; Park, T.;
Properties of TiO2 Nanohelices with Embedded Hc (Nh 2) 2 Pbi 3 Choi, J. Systematically Optimized Bilayered Electron Transport Layer
Perovskite Solar Cells. J. Mater. Chem. A 2015, 3, 9179−9186. for Highly Efficient Planar Perovskite Solar Cells (η= 21.1%). ACS
(241) Lee, D. G.; Kim, M.-c.; Kim, B. J.; Kim, D. H.; Lee, S. M.; Energy Lett. 2017, 2, 2667−2673.
Choi, M.; Lee, S.; Jung, H. S. Effect of TiO2 Particle Size and Layer (259) Chen, J.; Zhao, X.; Kim, S. G.; Park, N. G. Multifunctional
Thickness on Mesoscopic Perovskite Solar Cells. Appl. Surf. Sci. 2019, Chemical Linker Imidazoleacetic Acid Hydrochloride for 21%
477, 131−136. Efficient and Stable Planar Perovskite Solar Cells. Adv. Mater. 2019,
(242) Kim, D. H.; Han, G. S.; Seong, W. M.; Lee, J. W.; Kim, B. J.; 31, 1902902.
Park, N. G.; Hong, K. S.; Lee, S.; Jung, H. S. Niobium Doping Effects (260) Yang, D.; Yang, R.; Wang, K.; Wu, C.; Zhu, X.; Feng, J.; Ren,
on TiO2 Mesoscopic Electron Transport Layer-Based Perovskite Solar X.; Fang, G.; Priya, S.; Liu, S.; et al. High Efficiency Planar-Type
Cells. ChemSusChem 2015, 8, 2392−2398. Perovskite Solar Cells with Negligible Hysteresis Using Edta-
(243) Tan, H.; Jain, A.; Voznyy, O.; Lan, X.; Garcia De Arquer, F. P. Complexed Sno2. Nat. Commun. 2018, 9, 3239.
G.; Fan, J. Z.; Quintero-Bermudez, R.; Yuan, M.; Zhang, B.; Zhao, Y.; (261) Jiang, Q.; Zhao, Y.; Zhang, X.; Yang, X.; Chen, Y.; Chu, Z.; Ye,
et al. Efficient and Stable Solution-Processed Planar Perovskite Solar Q.; Li, X.; Yin, Z.; You, J. Surface Passivation of Perovskite Film for
Cells Via Contact Passivation. Science 2017, 355, 722−726. Efficient Solar Cells. Nat. Nat. Photonics 2019, 13, 460−466.
(244) Ito, S.; Tanaka, S.; Manabe, K.; Nishino, H. Effects of Surface (262) Christians, J. A.; Schulz, P.; Tinkham, J. S.; Schloemer, T. H.;
Blocking Layer of Sb2S3 on Nanocrystalline TiO2 for CH3NH3PbI3 Harvey, S. P.; Tremolet de Villers, B. J.; Sellinger, A.; Berry, J. J.;
Perovskite Solar Cells. J. Phys. Chem. C 2014, 118, 16995−17000. Luther, J. M. Tailored Interfaces of Unencapsulated Perovskite Solar
(245) Rahman, N. U.; Khan, W. U.; Khan, S.; Chen, X.; Khan, J.; Cells for > 1,000 h Operational Stability. Nat. Energy 2018, 3, 68−74.
Zhao, J.; Yang, Z.; Wu, M.; Chi, Z. A Promising Europium-Based (263) Zhu, Z.; Bai, Y.; Liu, X.; Chueh, C.-C.; Yang, S.; Jen, A. K. Y.
Down Conversion Material: Organic−Inorganic Perovskite Solar Enhanced Efficiency and Stability of Inverted Perovskite Solar Cells
Cells with High Photovoltaic Performance and Uv-Light Stability. J. Using Highly Crystalline SnO2 Nanocrystals as the Robust Electron-
Transporting Layer. Adv. Mater. 2016, 28, 6478−6484.
Mater. Chem. A 2019, 7, 6467−6474.
(264) Seo, S.; Jeong, S.; Bae, C.; Park, N. G.; Shin, H. Perovskite
(246) Kim, H.-S.; Jang, I.-H.; Ahn, N.; Choi, M.; Guerrero, A.;
Solar Cells with Inorganic Electron-and Hole-Transport Layers
Bisquert, J.; Park, N.-G. Control of I−V Hysteresis in CH3NH3PbI3
Exhibiting Long-Term (≈ 500 h) Stability at 85° C under Continuous
Perovskite Solar Cell. J. Phys. Chem. Lett. 2015, 6, 4633−4639.
1 Sun Illumination in Ambient Air. Adv. Mater. 2018, 30, 1801010.
(247) Kim, J. Y.; Jung, H. S.; No, J. H.; Kim, J.-R.; Hong, K. S.
(265) Son, D.-Y.; Bae, K.-H.; Kim, H.-S.; Park, N.-G. Effects of Seed
Influence of Anatase-Rutile Phase Transformation on Dielectric
Layer on Growth of Zno Nanorod and Performance of Perovskite
Properties of Sol-Gel Derived TiO2 Thin Films. J. Electroceram. 2006,
Solar Cell. J. Phys. Chem. C 2015, 119, 10321−10328.
16, 447−451. (266) Oh, L. S.; Kim, D. H.; Lee, J. A.; Shin, S. S.; Lee, J.-W.; Park, I.
(248) Chen, B.; Yang, M.; Priya, S.; Zhu, K. Origin of J−V J.; Ko, M. J.; Park, N.-G.; Pyo, S. G.; Hong, K. S.; et al. Zn2SnO4-
Hysteresis in Perovskite Solar Cells. J. Phys. Chem. Lett. 2016, 7, 905− Based Photoelectrodes for Organolead Halide Perovskite Solar Cells.
917. J. Phys. Chem. C 2014, 118, 22991−22994.
(249) Xu, J.; Buin, A.; Ip, A. H.; Li, W.; Voznyy, O.; Comin, R.; (267) Dou, J.; Zhang, Y.; Wang, Q.; Abate, A.; Li, Y.; Wei, M. Highly
Yuan, M.; Jeon, S.; Ning, Z.; McDowell, J. J.; et al. Perovskite− Efficient Zn2SnO4 Perovskite Solar Cells through Band Alignment
Fullerene Hybrid Materials Suppress Hysteresis in Planar Diodes. Nat. Engineering. Chem. Commun. 2019, 55, 14673−14676.
Commun. 2015, 6, 7081. (268) Shibayama, N.; Kanda, H.; Kim, T. W.; Segawa, H.; Ito, S.
(250) Kim, B. J.; Kim, M. c.; Lee, D. G.; Lee, G.; Bang, G. J.; Jeon, J. Design of BCP Buffer Layer for Inverted Perovskite Solar Cells Using
B.; Choi, M.; Jung, H. S. Interface Design of Hybrid Electron Ideal Factor. APL Mater. 2019, 7, 031117.
Extraction Layer for Relieving Hysteresis and Retarding Charge (269) Park, I. J.; Kang, G.; Park, M. A.; Kim, J. S.; Seo, S. W.; Kim,
Recombination in Perovskite Solar Cells. Adv. Mater. Interfaces 2018, D. H.; Zhu, K.; Park, T.; Kim, J. Y. Highly Efficient and Uniform 1
5, 1800993. cm2 Perovskite Solar Cells with an Electrochemically Deposited Niox
(251) Lee, J.-W.; Kim, S.-G.; Bae, S.-H.; Lee, D.-K.; Lin, O.; Yang, Hole-Extraction Layer. ChemSusChem 2017, 10, 2660−2667.
Y.; Park, N.-G. The Interplay between Trap Density and Hysteresis in (270) Song, S.; Moon, B. J.; Hörantner, M. T.; Lim, J.; Kang, G.;
Planar Heterojunction Perovskite Solar Cells. Nano Lett. 2017, 17, Park, M.; Kim, J. Y.; Snaith, H. J.; Park, T. Interfacial Electron
4270−4276. Accumulation for Efficient Homo-Junction Perovskite Solar Cells.
(252) Jiang, Q.; Zhang, X.; You, J. SnO2: A Wonderful Electron Nano Energy 2016, 28, 269−276.
Transport Layer for Perovskite Solar Cells. Small 2018, 14, 1801154. (271) Ciro, J.; Mesa, S.; Uribe, J. I.; Mejía-Escobar, M. A.; Ramirez,
(253) Jiang, Q.; Zhang, L.; Wang, H.; Yang, X.; Meng, J.; Liu, H.; D.; Montoya, J. F.; Betancur, R.; Yoo, H.-S.; Park, N.-G.; Jaramillo, F.
Yin, Z.; Wu, J.; Zhang, X.; You, J. Enhanced Electron Extraction Using Optimization of the Ag/PCBM Interface by a Rhodamine Interlayer
SnO2 for High-Efficiency Planar-Structure HC(NH2)2PbI3-Based to Enhance the Efficiency and Stability of Perovskite Solar Cells.
Perovskite Solar Cells. Nat. Energy 2017, 2, 16177. Nanoscale 2017, 9, 9440−9446.
(254) Jung, K.-H.; Seo, J.-Y.; Lee, S.; Shin, H.; Park, N.-G. Solution- (272) Meng, X.; Ho, C. H. Y.; Xiao, S.; Bai, Y.; Zhang, T.; Hu, C.;
Processed SnO2 Thin Film for a Hysteresis-Free Planar Perovskite Lin, H.; Yang, Y.; So, S. K.; Yang, S. Molecular Design Enabled
Solar Cell with a Power Conversion Efficiency of 19.2%. J. Mater. Reduction of Interface Trap Density Affords Highly Efficient and
Chem. A 2017, 5, 24790−24803. Stable Perovskite Solar Cells with over 83% Fill Factor. Nano Energy
(255) Jiang, Q.; Chu, Z.; Wang, P.; Yang, X.; Liu, H.; Wang, Y.; Yin, 2018, 52, 300−306.
Z.; Wu, J.; Zhang, X.; You, J. Planar-Structure Perovskite Solar Cells (273) Bach, U.; Lupo, D.; Comte, P.; Moser, J. E.; Weissörtel, F.;
with Efficiency Beyond 21%. Adv. Mater. 2017, 29, 1703852. Salbeck, J.; Spreitzer, H.; Grätzel, M. Solid-State Dye-Sensitized
(256) Jeong, S.; Seo, S.; Park, H.; Shin, H. Atomic Layer Deposition Mesoporous Tio2 Solar Cells with High Photon-to-Electron
of a SnO2 Electron-Transporting Layer for Planar Perovskite Solar Conversion Efficiencies. Nature 1998, 395, 583−585.

AW https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

(274) Tan, B.; Raga, S. R.; Chesman, A. S. R.; Fürer, S. O.; Zheng, Layer Via the Combustion Method for High-Performance Thin-Film
F.; McMeekin, D. P.; Jiang, L.; Mao, W.; Lin, X.; Wen, X.; et al. Litfsi- Perovskite Solar Cells. Adv. Mater. 2015, 27, 7874−7880.
Free Spiro-Ometad-Based Perovskite Solar Cells with Power (291) Lyu, M.; Chen, J.; Park, N.-G. Improvement of Efficiency and
Conversion Efficiencies Exceeding 19%. Adv. Energy Mater. 2019, 9, Stability of Cuscn-Based Inverted Perovskite Solar Cells by Post-
1901519. Treatment with Potassium Thiocyanate. J. Solid State Chem. 2019,
(275) Schloemer, T. H.; Christians, J. A.; Luther, J. M.; Sellinger, A. 269, 367−374.
Doping Strategies for Small Molecule Organic Hole-Transport (292) Liu, J.; Pathak, S. K.; Sakai, N.; Sheng, R.; Bai, S.; Wang, Z.;
Materials: Impacts on Perovskite Solar Cell Performance and Stability. Snaith, H. J. Identification and Mitigation of a Critical Interfacial
Chem. Sci. 2019, 10, 1904−1935. Instability in Perovskite Solar Cells Employing Copper Thiocyanate
(276) Rakstys, K.; Igci, C.; Nazeeruddin, M. K. Efficiency Vs. Hole-Transporter. Adv. Mater. Interfaces 2016, 3, 1600571.
Stability: Dopant-Free Hole Transporting Materials Towards (293) Arora, N.; Dar, M. I.; Hinderhofer, A.; Pellet, N.; Schreiber,
Stabilized Perovskite Solar Cells. Chem. Sci. 2019, 10, 6748−6769. F.; Zakeeruddin, S. M.; Grätzel, M. Perovskite Solar Cells with Cuscn
(277) Xiao, Z.; Dong, Q.; Bi, C.; Shao, Y.; Yuan, Y.; Huang, J. Hole Extraction Layers Yield Stabilized Efficiencies Greater Than
Solvent Annealing of Perovskite-Induced Crystal Growth for Photo- 20%. Science 2017, 358, 768.
voltaic-Device Efficiency Enhancement. Adv. Mater. 2014, 26, 6503− (294) Rao, H.; Ye, S.; Sun, W.; Yan, W.; Li, Y.; Peng, H.; Liu, Z.;
6509. Bian, Z.; Li, Y.; Huang, C. A 19.0% Efficiency Achieved in Cuox-
(278) Park, I. J.; Park, M. A.; Kim, D. H.; Park, G. D.; Kim, B. J.; Based Inverted CH3NH3PbI3−xClx Solar Cells by an Effective Cl
Son, H. J.; Ko, M. J.; Lee, D.-K.; Park, T.; Shin, H.; et al. New Hybrid Doping Method. Nano Energy 2016, 27, 51−57.
Hole Extraction Layer of Perovskite Solar Cells with a Planar P−I−N (295) Guo, Y.; Lei, H.; Xiong, L.; Li, B.; Fang, G. An Integrated
Geometry. J. Phys. Chem. C 2015, 119, 27285−27290. Organic−Inorganic Hole Transport Layer for Efficient and Stable
(279) Eperon, G. E.; Leijtens, T.; Bush, K. A.; Prasanna, R.; Green, Perovskite Solar Cells. J. Mater. Chem. A 2018, 6, 2157−2165.
T.; Wang, J. T.-W.; McMeekin, D. P.; Volonakis, G.; Milot, R. L.; (296) Nejand, B. A.; Ahmadi, V.; Gharibzadeh, S.; Shahverdi, H. R.
May, R.; et al. Perovskite-Perovskite Tandem Photovoltaics with Cuprous Oxide as a Potential Low-Cost Hole-Transport Material for
Optimized Band Gaps. Science 2016, 354, 861−865. Stable Perovskite Solar Cells. ChemSusChem 2016, 9, 302−313.
(280) Tong, J.; Song, Z.; Kim, D. H.; Chen, X.; Chen, C.; (297) Jeong, S.; Seo, S.; Shin, H. P-Type CuCrO2 Particulate Films
Palmstrom, A. F.; Ndione, P. F.; Reese, M. O.; Dunfield, S. P.; Reid, as the Hole Transporting Layer for CH3NH3PbI3 Perovskite Solar
O. G.; et al. Carrier Lifetimes of > 1 ms in Sn-Pb Perovskites Enable Cells. RSC Adv. 2018, 8, 27956−27962.
Efficient All-Perovskite Tandem Solar Cells. Science 2019, 364, 475− (298) Chen, Y.; Yang, Z.; Jia, X.; Wu, Y.; Yuan, N.; Ding, J.; Zhang,
479. W.-H.; Liu, S. Thermally Stable Methylammonium-Free Inverted
(281) Bi, C.; Wang, Q.; Shao, Y.; Yuan, Y.; Xiao, Z.; Huang, J. Non- Perovskite Solar Cells with Zn2+ Doped CuGaO2 as Efficient
Wetting Surface-Driven High-Aspect-Ratio Crystalline Grain Growth Mesoporous Hole-Transporting Layer. Nano Energy 2019, 61, 148−
for Efficient Hybrid Perovskite Solar Cells. Nat. Commun. 2015, 6,
157.
7747. (299) Yang, B.; Ouyang, D.; Huang, Z.; Ren, X.; Zhang, H.; Choy,
(282) Liu, X.; Cheng, Y.; Liu, C.; Zhang, T.; Zhang, N.; Zhang, S.;
W. C. H. Multifunctional Synthesis Approach of In:CuCrO2
Chen, J.; Xu, Q.; Ouyang, J.; Gong, H. 20.7% Highly Reproducible
Nanoparticles for Hole Transport Layer in High-Performance
Inverted Planar Perovskite Solar Cells with Enhanced Fill Factor and
Perovskite Solar Cells. Adv. Funct. Mater. 2019, 29, 1902600.
Eliminated Hysteresis. Energy Environ. Sci. 2019, 12, 1622−1633.
(300) Akin, S.; Liu, Y.; Dar, M. I.; Zakeeruddin, S. M.; Grätzel, M.;
(283) Kim, D. H.; Muzzillo, C. P.; Tong, J.; Palmstrom, A. F.;
Turan, S.; Sonmezoglu, S. Hydrothermally Processed Cucro2
Larson, B. W.; Choi, C.; Harvey, S. P.; Glynn, S.; Whitaker, J. B.;
Nanoparticles as an Inorganic Hole Transporting Material for Low-
Zhang, F.; et al. Bimolecular Additives Improve Wide-Band-Gap
Perovskites for Efficient Tandem Solar Cells with Cigs. Joule 2019, 3, Cost Perovskite Solar Cells with Superior Stability. J. Mater. Chem. A
1734−1745. 2018, 6, 20327−20337.
(284) Jung, E. H.; Jeon, N. J.; Park, E. Y.; Moon, C. S.; Shin, T. J.; (301) Zhang, H.; Wang, H.; Chen, W.; Jen, A. K. Y. CuGaO2: A
Yang, T.-Y.; Noh, J. H.; Seo, J. Efficient, Stable and Scalable Promising Inorganic Hole-Transporting Material for Highly Efficient
Perovskite Solar Cells Using Poly(3-Hexylthiophene). Nature 2019, and Stable Perovskite Solar Cells. Adv. Mater. 2017, 29, 1604984.
567, 511−515. (302) Park, I. J.; Park, J. H.; Ji, S. G.; Park, M.-A.; Jang, J. H.; Kim, J.
(285) Jeng, J.-Y.; Chen, K.-C.; Chiang, T.-Y.; Lin, P.-Y.; Tsai, T.-D.; Y. A Three-Terminal Monolithic Perovskite/Si Tandem Solar Cell
Chang, Y.-C.; Guo, T.-F.; Chen, P.; Wen, T.-C.; Hsu, Y.-J. Nickel Characterization Platform. Joule 2019, 3, 807−818.
Oxide Electrode Interlayer in CH3NH3PbI3 Perovskite/Pcbm Planar- (303) Bai, S.; Da, P.; Li, C.; Wang, Z.; Yuan, Z.; Fu, F.; Kawecki, M.;
Heterojunction Hybrid Solar Cells. Adv. Mater. 2014, 26, 4107−4113. Liu, X.; Sakai, N.; Wang, J. T.-W.; et al. Planar Perovskite Solar Cells
(286) Yan, X.; Zheng, J. H.; Zheng, L. L.; Lin, G. H.; Lin, H. D.; with Long-Term Stability Using Ionic Liquid Additives. Nature 2019,
Chen, G.; Du, B. B.; Zhang, F. Y. Optimization of Sputtering Niox 571, 245−250.
Films for Perovskite Solar Cell Applications. Mater. Res. Bull. 2018, (304) Zhou, Y.; Fuentes-Hernandez, C.; Shim, J.; Meyer, J.;
103, 150−157. Giordano, A. J.; Li, H.; Winget, P.; Papadopoulos, T.; Cheun, H.;
(287) Abzieher, T.; Moghadamzadeh, S.; Schackmar, F.; Eggers, H.; Kim, J.; et al. A Universal Method to Produce Low−Work Function
Sutterlüti, F.; Farooq, A.; Kojda, D.; Habicht, K.; Schmager, R.; Electrodes for Organic Electronics. Science 2012, 336, 327.
Mertens, A.; et al. Electron-Beam-Evaporated Nickel Oxide Hole (305) Li, Y.; Zhao, Y.; Chen, Q.; Yang, Y.; Liu, Y.; Hong, Z.; Liu, Z.;
Transport Layers for Perovskite-Based Photovoltaics. Adv. Energy Hsieh, Y.-T.; Meng, L.; Li, Y.; et al. Multifunctional Fullerene
Mater. 2019, 9, 1802995. Derivative for Interface Engineering in Perovskite Solar Cells. J. Am.
(288) Seo, S.; Park, I. J.; Kim, M.; Lee, S.; Bae, C.; Jung, H. S.; Park, Chem. Soc. 2015, 137, 15540−15547.
N.-G.; Kim, J. Y.; Shin, H. An Ultra-Thin, Un-Doped Nio Hole (306) Agresti, A.; Pescetelli, S.; Palma, A. L.; Del Rio Castillo, A. E.;
Transporting Layer of Highly Efficient (16.4%) Organic−Inorganic Konios, D.; Kakavelakis, G.; Razza, S.; Cinà, L.; Kymakis, E.;
Hybrid Perovskite Solar Cells. Nanoscale 2016, 8, 11403−11412. Bonaccorso, F.; et al. Graphene Interface Engineering for Perovskite
(289) Park, M.-A.; Park, I. J.; Park, S.; Kim, J.; Jo, W.; Son, H. J.; Solar Modules: 12.6% Power Conversion Efficiency over 50 cm2
Kim, J. Y. Enhanced Electrical Properties of Li−Doped Niox Hole Active Area. ACS Energy Lett. 2017, 2, 279−287.
Extraction Layer in p−i−n Type Perovskite Solar Cells. Curr. Appl. (307) O’Keeffe, P.; Catone, D.; Paladini, A.; Toschi, F.; Turchini, S.;
Phys. 2018, 18, S55−S59. Avaldi, L.; Martelli, F.; Agresti, A.; Pescetelli, S.; Del Rio Castillo, A.
(290) Jung, J. W.; Chueh, C.-C.; Jen, A. K. Y. A Low-Temperature, E.; et al. Graphene-Induced Improvements of Perovskite Solar Cell
Solution-Processable, Cu-Doped Nickel Oxide Hole-Transporting Stability: Effects on Hot-Carriers. Nano Lett. 2019, 19, 684−691.

AX https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

(308) Wu, W.-Q.; Chen, D.; Cheng, Y.-B.; Caruso, R. A. Thin Films (325) Uzu, H.; Ichikawa, M.; Hino, M.; Nakano, K.; Meguro, T.;
of Tin Oxide Nanosheets Used as the Electron Transporting Layer for Hernández, J. L.; Kim, H.-S.; Park, N.-G.; Yamamoto, K. High
Improved Performance and Ambient Stability of Perovskite Photo- Efficiency Solar Cells Combining a Perovskite and a Silicon
voltaics. Solar RRL 2017, 1, 1700117. Heterojunction Solar Cells Via an Optical Splitting System. Appl.
(309) Ke, W.; Zhao, D.; Xiao, C.; Wang, C.; Cimaroli, A. J.; Grice, Phys. Lett. 2015, 106, 013506.
C. R.; Yang, M.; Li, Z.; Jiang, C.-S.; Al-Jassim, M.; et al. Cooperative (326) Zheng, J.; Lau, C. F. J.; Mehrvarz, H.; Ma, F.-J.; Jiang, Y.;
Tin Oxide Fullerene Electron Selective Layers for High-Performance Deng, X.; Soeriyadi, A.; Kim, J.; Zhang, M.; Hu, L.; et al. Large Area
Planar Perovskite Solar Cells. J. Mater. Chem. A 2016, 4, 14276− Efficient Interface Layer Free Monolithic Perovskite/Homo-Junction-
14283. Silicon Tandem Solar Cell with over 20% Efficiency. Energy Environ.
(310) Xie, J.; Huang, K.; Yu, X.; Yang, Z.; Xiao, K.; Qiang, Y.; Zhu, Sci. 2018, 11, 2432−2443.
X.; Xu, L.; Wang, P.; Cui, C.; et al. Enhanced Electronic Properties of (327) Jošt, M.; Köhnen, E.; Morales-Vilches, A. B.; Lipovšek, B.;
SnO2 Via Electron Transfer from Graphene Quantum Dots for Jäger, K.; Macco, B.; Al-Ashouri, A.; Krč, J.; Korte, L.; Rech, B.; et al.
Efficient Perovskite Solar Cells. ACS Nano 2017, 11, 9176−9182. Textured Interfaces in Monolithic Perovskite/Silicon Tandem Solar
(311) Ghosh, S.; Singh, T. Role of Ionic Liquids in Organic- Cells: Advanced Light Management for Improved Efficiency and
Inorganic Metal Halide Perovskite Solar Cells Efficiency and Stability. Energy Yield. Energy Environ. Sci. 2018, 11, 3511−3523.
Nano Energy 2019, 63, 103828. (328) Zheng, J.; Mehrvarz, H.; Liao, C.; Bing, J.; Cui, X.; Li, Y.;
(312) Han, G. S.; Chung, H. S.; Kim, B. J.; Kim, D. H.; Lee, J. W.; Gonçales, V. R.; Lau, C. F. J.; Lee, D. S.; Li, Y.; et al. Large-Area 23%-
Swain, B. S.; Mahmood, K.; Yoo, J. S.; Park, N.-G.; Lee, J. H.; et al. Efficient Monolithic Perovskite/Homojunction-Silicon Tandem Solar
Retarding Charge Recombination in Perovskite Solar Cells Using Cell with Enhanced Uv Stability Using Down-Shifting Material. ACS
Ultrathin Mgo-Coated TiO2 Nanoparticulate Films. J. Mater. Chem. A Energy Lett. 2019, 4, 2623−2631.
2015, 3, 9160−9164. (329) Bush, K. A.; Frohna, K.; Prasanna, R.; Beal, R. E.; Leijtens, T.;
(313) Son, D.-Y.; Kim, S.-G.; Seo, J.-Y.; Lee, S.-H.; Shin, H.; Lee, D.; Swifter, S. A.; McGehee, M. D. Compositional Engineering for
Park, N.-G. Universal Approach toward Hysteresis-Free Perovskite Efficient Wide Band Gap Perovskites with Improved Stability to
Solar Cell Via Defect Engineering. J. Am. Chem. Soc. 2018, 140, Photoinduced Phase Segregation. ACS Energy Lett. 2018, 3, 428−435.
1358−1364. (330) Gharibzadeh, S.; Abdollahi Nejand, B.; Jakoby, M.; Abzieher,
(314) Saidaminov, M. I.; Kim, J.; Jain, A.; Quintero-Bermudez, R.; T.; Hauschild, D.; Moghadamzadeh, S.; Schwenzer, J. A.; Brenner, P.;
Tan, H.; Long, G.; Tan, F.; Johnston, A.; Zhao, Y.; Voznyy, O.; et al. Schmager, R.; Haghighirad, A. A.; et al. Record Open-Circuit Voltage
Suppression of Atomic Vacancies Via Incorporation of Isovalent Small Wide-Bandgap Perovskite Solar Cells Utilizing 2D/3D Perovskite
Ions to Increase the Stability of Halide Perovskite Solar Cells in Heterostructure. Adv. Energy Mater. 2019, 9, 1803699.
(331) Ramírez Quiroz, C. O.; Spyropoulos, G. D.; Salvador, M.;
Ambient Air. Nat. Energy 2018, 3, 648−654.
Roch, L. M.; Berlinghof, M.; Darío Perea, J.; Forberich, K.; Dion-
(315) Park, Y. H.; Jeong, I.; Bae, S.; Son, H. J.; Lee, P.; Lee, J.; Lee,
Bertrand, L.-I.; Schrenker, N. J.; Classen, A.; et al. Interface Molecular
C.-H.; Ko, M. J. Inorganic Rubidium Cation as an Enhancer for
Engineering for Laminated Monolithic Perovskite/Silicon Tandem
Photovoltaic Performance and Moisture Stability of HC(NH2)2PbI3
Solar Cells with 80.4% Fill Factor. Adv. Funct. Mater. 2019, 29,
Perovskite Solar Cells. Adv. Funct. Mater. 2017, 27, 1605988.
1901476.
(316) Min, H.; Kim, M.; Lee, S.-U.; Kim, H.; Kim, G.; Choi, K.; Lee,
(332) Köhnen, E.; Jošt, M.; Morales-Vilches, A. B.; Tockhorn, P.; Al-
J. H.; Seok, S. I. Efficient, Stable Solar Cells by Using Inherent
Ashouri, A.; Macco, B.; Kegelmann, L.; Korte, L.; Rech, B.;
Bandgap of Α-Phase Formamidinium Lead Iodide. Science 2019, 366,
Schlatmann, R.; et al. Highly Efficient Monolithic Perovskite Silicon
749−753. Tandem Solar Cells: Analyzing the Influence of Current Mismatch on
(317) Gao, F.; Zhao, Y.; Zhang, X.; You, J. Recent Progresses on Device Performance. Sustain. Energy Fuels 2019, 3, 1995−2005.
Defect Passivation toward Efficient Perovskite Solar Cells. Adv. Energy (333) Sahli, F.; Werner, J.; Kamino, B. A.; Bräuninger, M.; Monnard,
Mater. 2020, 10, 1902650. R.; Paviet-Salomon, B.; Barraud, L.; Ding, L.; Diaz Leon, J. J.;
(318) Li, X.; Dar, M. I.; Yi, C.; Luo, J.; Tschumi, M.; Zakeeruddin, S. Sacchetto, D.; et al. Fully Textured Monolithic Perovskite/Silicon
M.; Nazeeruddin, M. K.; Han, H.; Grätzel, M. Improved Performance Tandem Solar Cells with 25.2% Power Conversion Efficiency. Nat.
and Stability of Perovskite Solar Cells by Crystal Crosslinking with Mater. 2018, 17, 820−826.
Alkylphosphonic Acid Ω-Ammonium Chlorides. Nat. Chem. 2015, 7, (334) Hou, Y.; Aydin, E.; De Bastiani, M.; Xiao, C.; Isikgor, F. H.;
703−711. Xue, D.-J.; Chen, B.; Chen, H.; Bahrami, B.; Chowdhury, A. H.; et al.
(319) Li, X.; Zhang, W.; Wang, Y.-C.; Zhang, W.; Wang, H.-Q.; Efficient Tandem Solar Cells with Solution-Processed Perovskite on
Fang, J. In-Situ Cross-Linking Strategy for Efficient and Operationally Textured Crystalline Silicon. Science 2020, 367, 1135.
Stable Methylammoniun Lead Iodide Solar Cells. Nat. Commun. (335) Xu, J.; Boyd, C. C.; Yu, Z. J.; Palmstrom, A. F.; Witter, D. J.;
2018, 9, 3806. Larson, B. W.; France, R. M.; Werner, J.; Harvey, S. P.; Wolf, E. J.;
(320) Zheng, X.; Chen, B.; Dai, J.; Fang, Y.; Bai, Y.; Lin, Y.; Wei, H.; et al. Triple-Halide Wide−Band Gap Perovskites with Suppressed
Zeng, X. C.; Huang, J. Defect Passivation in Hybrid Perovskite Solar Phase Segregation for Efficient Tandems. Science 2020, 367, 1097−
Cells Using Quaternary Ammonium Halide Anions and Cations. Nat. 1104.
Energy 2017, 2, 17102. (336) Kim, D.; Jung, H. J.; Park, I. J.; Larson, B. W.; Dunfield, S. P.;
(321) Best Research-Cell Efficiency Chart; National Renewable Xiao, C.; Kim, J.; Tong, J.; Boonmongkolras, P.; Ji, S. G.; et al.
Research Laboratory, 2019; https://www.nrel.gov/pv/cell-efficiency. Efficient, Stable Silicon Tandem Cells Enabled by Anion-Engineered
html. Wide-Bandgap Perovskites. Science 2020, 368, 155−160.
(322) Anaya, M.; Lozano, G.; Calvo, M. E.; Míguez, H. Abx3 (337) Han, Q.; Hsieh, Y.-T.; Meng, L.; Wu, J.-L.; Sun, P.; Yao, E.-P.;
Perovskites for Tandem Solar Cells. Joule 2017, 1, 769−793. Chang, S.-Y.; Bae, S.-H.; Kato, T.; Bermudez, V.; et al. High-
(323) Bush, K. A.; Palmstrom, A. F.; Yu, Z. J.; Boccard, M.; Performance Perovskite/Cu(in, Ga)Se2 Monolithic Tandem Solar
Cheacharoen, R.; Mailoa, J. P.; McMeekin, D. P.; Hoye, R. L.; Bailie, Cells. Science 2018, 361, 904−908.
C. D.; Leijtens, T.; et al. 23.6%-Efficient Monolithic Perovskite/ (338) Forgács, D.; Gil-Escrig, L.; Pérez-Del-Rey, D.; Momblona, C.;
Silicon Tandem Solar Cells with Improved Stability. Nat. Energy Werner, J.; Niesen, B.; Ballif, C.; Sessolo, M.; Bolink, H. J. Efficient
2017, 2, 17009. Monolithic Perovskite/Perovskite Tandem Solar Cells. Adv. Energy
(324) Fu, F.; Feurer, T.; Weiss, T. P.; Pisoni, S.; Avancini, E.; Mater. 2017, 7, 1602121.
Andres, C.; Buecheler, S.; Tiwari, A. N High-Efficiency Inverted Semi- (339) Liao, W.; Zhao, D.; Yu, Y.; Shrestha, N.; Ghimire, K.; Grice,
Transparent Planar Perovskite Solar Cells in Substrate Configuration. C. R.; Wang, C.; Xiao, Y.; Cimaroli, A. J.; Ellingson, R. J. Fabrication
Nat. Energy 2017, 2, 16190. of Efficient Low-Bandgap Perovskite Solar Cells by Combining

AY https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Formamidinium Tin Iodide with Methylammonium Lead Iodide. J. Irradiance Performance of Metal Halide Perovskites for Concentrator
Am. Chem. Soc. 2016, 138, 12360−12363. Photovoltaics. Nat. Energy 2018, 3, 855.
(340) Kim, D. Y.; Guijt, E.; Si, F. T.; Santbergen, R.; Holovský, J.;
Isabella, O.; van Swaaij, R. A. C. M. M.; Zeman, M. Fabrication of
Double- and Triple-Junction Solar Cells with Hydrogenated
Amorphous Silicon Oxide (a-SiOx:H) Top Cell. Sol. Energy Mater.
Sol. Cells 2015, 141, 148−153.
(341) Ross, R. T.; Nozik, A. J. Efficiency of Hot-Carrier Solar Energy
Converters. J. Appl. Phys. 1982, 53, 3813−3818.
(342) Kahmann, S.; Loi, M. A. Hot Carrier Solar Cells and the
Potential of Perovskites for Breaking the Shockley−Queisser Limit. J.
Mater. Chem. C 2019, 7, 2471−2486.
(343) Hirst, L. C.; Fujii, H.; Wang, Y.; Sugiyama, M.; Ekins-Daukes,
N. J. Hot Carriers in Quantum Wells for Photovoltaic Efficiency
Enhancement. IEEE J. Photovolt. 2014, 4, 244−252.
(344) Harada, Y.; Kasamatsu, N.; Watanabe, D.; Kita, T.
Nanosecond-Scale Hot-Carrier Cooling Dynamics in One-Dimen-
sional Quantum Dot Superlattices. Phys. Rev. B: Condens. Matter
Mater. Phys. 2016, 93, 115303.
(345) Nguyen, D.-T.; Lombez, L.; Gibelli, F.; Boyer-Richard, S.; Le
Corre, A.; Durand, O.; Guillemoles, J.-F. Quantitative Experimental
Assessment of Hot Carrier-Enhanced Solar Cells at Room Temper-
ature. Nat. Energy 2018, 3, 236.
(346) Price, M. B.; Butkus, J.; Jellicoe, T. C.; Sadhanala, A.; Briane,
A.; Halpert, J. E.; Broch, K.; Hodgkiss, J. M.; Friend, R. H.; Deschler,
F. Hot-Carrier Cooling and Photoinduced Refractive Index Changes
in Organic−Inorganic Lead Halide Perovskites. Nat. Commun. 2015,
6, 8420.
(347) Flender, O.; Klein, J. R.; Lenzer, T.; Oum, K. Ultrafast
Photoinduced Dynamics of the Organolead Trihalide Perovskite
CH3NH3PbI3 on Mesoporous TiO2 Scaffolds in the 320−920 nm
Range. Phys. Chem. Chem. Phys. 2015, 17, 19238−19246.
(348) Li, M.; Bhaumik, S.; Goh, T. W.; Kumar, M. S.; Yantara, N.;
Grätzel, M.; Mhaisalkar, S.; Mathews, N.; Sum, T. C. Slow Cooling
and Highly Efficient Extraction of Hot Carriers in Colloidal
Perovskite Nanocrystals. Nat. Commun. 2017, 8, 14350.
(349) Fu, J.; Xu, Q.; Han, G.; Wu, B.; Huan, C. H. A.; Leek, M. L.;
Sum, T. C. Hot Carrier Cooling Mechanisms in Halide Perovskites.
Nat. Commun. 2017, 8, 1300.
(350) Guo, Z.; Wan, Y.; Yang, M.; Snaider, J.; Zhu, K.; Huang, L.
Long-Range Hot-Carrier Transport in Hybrid Perovskites Visualized
by Ultrafast Microscopy. Science 2017, 356, 59−62.
(351) Jiménez-López, J.; Puscher, B. M.; Guldi, D. M.; Palomares, E.
Improved Carrier Collection and Hot Electron Extraction across
Perovskite, C60, and TiO2 Interfaces. J. Am. Chem. Soc. 2020, 142,
1236.
(352) Miller, O. D.; Yablonovitch, E.; Kurtz, S. R. Strong Internal
and External Luminescence as Solar Cells Approach the Shockley−
Queisser Limit. IEEE J. Photovolt. 2012, 2, 303−311.
(353) Pazos-Outón, L. M.; Szumilo, M.; Lamboll, R.; Richter, J. M.;
Crespo-Quesada, M.; Abdi-Jalebi, M.; Beeson, H. J.; Vrućinić, M.;
Alsari, M.; Snaith, H. J.; et al. Photon Recycling in Lead Iodide
Perovskite Solar Cells. Science 2016, 351, 1430−1433.
(354) Kirchartz, T.; Staub, F.; Rau, U. Impact of Photon Recycling
on the Open-Circuit Voltage of Metal Halide Perovskite Solar Cells.
ACS Energy Lett. 2016, 1, 731−739.
(355) Fang, Y.; Wei, H.; Dong, Q.; Huang, J. Quantification of Re-
Absorption and Re-Emission Processes to Determine Photon
Recycling Efficiency in Perovskite Single Crystals. Nat. Commun.
2017, 8, 14417.
(356) Ahrenkiel, R.; Dunlavy, D.; Keyes, B.; Vernon, S.; Dixon, T.;
Tobin, S.; Miller, K.; Hayes, R. Ultralong Minority-Carrier Lifetime
Epitaxial Gaas by Photon Recycling. Appl. Phys. Lett. 1989, 55, 1088−
1090.
(357) Lin, Q.; Wang, Z.; Snaith, H. J.; Johnston, M. B.; Herz, L. M.
Hybrid Perovskites: Prospects for Concentrator Solar Cells. Adv. Sci.
2018, 5, 1700792.
(358) Wang, Z.; Lin, Q.; Wenger, B.; Christoforo, M. G.; Lin, Y.-H.;
Klug, M. T.; Johnston, M. B.; Herz, L. M.; Snaith, H. J. High

AZ https://dx.doi.org/10.1021/acs.chemrev.0c00107
Chem. Rev. XXXX, XXX, XXX−XXX

You might also like