Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

nature reviews chemistry https://doi.org/10.

1038/s41570-023-00492-z

Review article Check for updates

Stabilization of photoactive phases


for perovskite photovoltaics
Xueping Liu    1, Deying Luo    2  , Zheng-Hong Lu    2, Jae Sung Yun    1, Michael Saliba    3,4  , Sang Il Seok    5 
& Wei Zhang    1 
Abstract Sections

Interest in photovoltaics (PVs) based on Earth-abundant halide perovs­­ Introduction

kites has increased markedly in recent years owing to the remarkable The demand for phase stability
properties of these materials and their suitability for energy-efficient and The promise of FAPbI3 light
scalable solution processing. Formamidinium lead triiodide (FAPbI3)-rich absorbers
perovskite absorbers have emerged as the frontrunners for commercializa­ The challenges of stabilization
tion, but commercial success is reliant on the stability meeting the high-
Fabrication and stabilization
est industrial standards and the photoactive FAPbI3 phase suffers from strategies
instabilities that lead to degradation — an effect that is accelerated under Upscaling towards
working conditions. Here, we critically assess the current understanding commercialization

of these phase instabilities and summarize the approaches for stabilizing Outlook
the desired phases, covering aspects from fundamental research to device
engineering. We subsequently analyse the remaining challenges for state-
of-the-art perovskite PVs and demonstrate the opportunities to enhance
phase stability with ongoing materials discovery and in operando analysis.
Finally, we propose future directions towards upscaling perovskite
modules, multijunction PVs and other potential applications.

Non-photoactive phase Photoactive phase

Decrease Gibbs free energy

Increase energy barrier

Advanced Technology Institute, University of Surrey, Guildford, UK. 2Department of Materials Science and
1

Engineering, University of Toronto, Toronto, Ontario, Canada. 3Institute for Photovoltaics (IPV), University
of Stuttgart, Stuttgart, Germany. 4Helmholtz Young Investigator Group FRONTRUNNER, IEK5-Photovoltaik,
Forschungszentrum Jülich, Jülich, Germany. 5Department of Energy Engineering, School of Energy and Chemical
Engineering, Ulsan National Institute of Science and Technology, Ulsan, South Korea.  e-mail: deying.luo@
utoronto.ca; michael.saliba@ipv.uni-stuttgart.de; seoksi@unist.ac.kr; wz0003@surrey.ac.uk

Nature Reviews Chemistry


Review article

Introduction light harvesting and achievable PCEs of the perovskite PVs, the band-
A milestone in the development history of perovskite photovoltaics gap of the absorber needs to be judiciously optimized (Fig. 1a). In a
(PVs) was the discovery of formamidinium lead triiodide (FAPbI3) per- single-junction PV, the ideal bandgap range to achieve the maximum
ovskite absorbers, in which the archetypal methylammonium (MA) theoretical PCE should, in principle, be restricted to 1.1–1.4 eV, accord-
cations are replaced by formamidinium (FA) cations on the A-site of ing to the defined Shockley–Queisser (S–Q) limit23. The bandgap of the
the ABX3 crystal structure1,2 (Box 1). Compared with the MA-based polycrystalline FAPbI3 film (~1.5 eV) is narrower than that of the proto-
perovskites, FAPbI3 absorbers show better thermal stability and nar- type MAPbI3 (~1.6 eV) and all-inorganic CsPbI3 (~1.7 eV) counterparts
rower bandgaps, which are beneficial for enhancing lifetime of the and reasonably close to the optimal bandgap region2,24,25, as shown in
device and maximizing the efficiency of solar energy harvesting3,4. Fig. 1b. Sn–Pb-alloyed perovskites can achieve a bandgap as narrow as
The past decade has seen reports of perovskite PVs using FAPbI3-rich 1.1 eV, but the spontaneous oxidation of Sn in conjunction with high
compositions (Box 1) with ever-increasing power conversion efficien- trap densities imposes constraints on pursuing higher device perfor-
cies (PCEs), culminating in a certified PCE of 25.7% in single-junction mances26. The distinctive material properties — such as optoelectronic
devices, comparable to commercially available crystalline silicon conversion and processability — render FAPbI3 extremely attractive for
PVs5,6. These advancements have encouraged researchers to move the best-performing perovskite PVs5 (Fig. 1c), yet stabilizing the black
from l­ ab­oratory-scale single-junction and multijunction devices photoactive phase remains a tough challenge.
towards modules7,8 (Box 1), and we predict that perovskite PVs will be a Indeed, the origin of the poor phase stability (Fig. 1d) of the FAPbI3
game-changer in the future PV market. is associated with the large Goldschmidt tolerance factor (t) (Box 2).
Despite these fascinating features, pristine FAPbI3 absorbers A common assessment of the formation possibility and stability of a
exhibit phase instability as a result of external stressors — including perovskite crystal phase is to examine whether the empirical t value,
light, heat, electrical bias, moisture and oxygen — during device manu- governed by the combined effect of ionic sizes in the ABX3 lattice, lies
facturing and operation against external stressors. In general, the in the range of the ideal cubic structure (between 0.9 and 1.0)27. On the
photoactive phase of pristine FAPbI3 (also known as the ‘black phase’) is basis of perfect sphere approximations of all ions used in the FAPbI3,
thermodynamically metastable and occurs only at temperatures above the theoretically calculated t (~0.99) fits this geometric constraint
390 K; below this temperature, a non-photoactive phase (also known marginally11 (Fig. 1e). However, the non-centrosymmetric structure
as the ‘yellow phase’) starts to dominate9–11. Non-photoactive phase of FA cation and the dynamic rotation within the [PbI6]4− octahedron
impurities existing within polycrystalline perovskite films on micro- framework presumably lead to a t-value outside the ideal range9,28.
scales and nanoscales tend to seed device degradation under normal Consequently, the cubic atomic stacking of FAPbI3 naturally requires
working conditions12. Efforts to improve the phase stability of FAPbI3 at a higher formation enthalpy and is prone to transformation from a
room temperature have been initiated by precisely controlling chemi- photoactive phase (cubic) to a non-photoactive phase (hexagonal) at
cal compositions through lattice site ion alloying, forming FAPbI3-rich room temperature. Compared with the large ionic size of FA+ (2.53 Å)
perovskites13,14. These alloyed perovskites have shown great potential in FAPbI3, the smaller MA+ (2.17 Å) results in the t value of MAPbI3 at the
to maintain photoactive phases at room temperature with favourable lower edge of the ideal range, while even smaller Cs+ (1.67 Å) makes
formation enthalpy (ΔHf; Box 2). However, lower photocurrent — owing CsPbI3 tend to form an asymmetric orthorhombic phase14. Alloying
to a blue shift of the bandgap — and the occurrence of phase segrega- multication or anion compositions in lattices to form a FAPbI3-rich
tion of the alloyed perovskites under working conditions compromise perovskite, as we defined in Box 1, has been broadly utilized as a strat-
device performance15,16. Other methods, such as solvent-coordinated egy to stabilize the photoactive phase at room temperature through
intermediate phases17, additives18,19 and strain engineering20, can help tuning the t-values11. Nevertheless, such alloying approaches bring
circumvent these issues; yet, the long-term stability remains subop- new challenges, including bandgap widening (beyond 1.5 eV), chemical
timal. Indeed, the operational stability of perovskite PVs remains an composition inhomogeneity and phase segregation under operational
order of magnitude lower than the required product lifetime of at least stressors29,30 (Fig. 1f). Consequently, increasing attention has focused
25 years, as dictated by the International Electrotechnical Commission on stabilizing photoactive phases without widening bandgaps and
61215:2016 tests qualification21,22. Hence, the most pressing challenge inducing phase segregation31,32.
in enhancing the operational stability of perovskite PVs is stabilizing
the photoactive phases to pass the industrial lifetime assessment tests. The promise of FAPbI3 light absorbers
In this Review, we comprehensively investigate the causes for Crystallographic modification by substituting MA+ with FA+ on the
the generation of non-photoactive phases of FAPbI3-rich perovskites, A-site of the perovskite lattice bestows upon FAPbI3 perovskite a more
encompassing the inherent material characteristics that facilitate the symmetrical crystal structure and exceptional optoelectronic prop-
phase instability when exposed to stressors under working conditions. erties5,33. Compared with MA-based and Cs-based perovskites, the
We then critically assess the effective means to stabilize the photoac- charge-carrier lifetime of FAPbI3 perovskite is significantly increased3,
tive phases while highlighting the opportunities they might offer in which may arise from the large polarons formed by faster reorientation
the search for commercially viable perovskite PVs. Subsequently, the motion of the FA+ in response to photogenerated carriers34,35. However,
effect of various scale-up processes on the perovskite film stability is it remains a matter of debate whether the polaron formation is more
analysed. Finally, we extend the discussion towards impending future closely related to dipolar [BX6]4− sublattice or A-site cations36. Further-
challenges and propose potential directions to boost perovskite PVs more, C=N double bond in FA+ (NH2CH = NH2+) is stronger than the C−N
towards a wider set of applications. single bond in MA+ (CH3−NH3+) owing to a sizeable contribution from
the conjugated π bond; meanwhile, the protonated FA+ in this case is
The demand for phase stability less acidic than MA+, enabling deprotonation to be more challeng-
A light absorber can only absorb photons with energies greater than ing37, resulting in improved thermal and light stability of the FAPbI3
or equal to its bandgap (Eg) to generate charge carriers. To maximize perovskites3. Conversely, the relatively strong acidity causes MA+ to

Nature Reviews Chemistry


Review article

Box 1

Rapid progress in device performances for halide perovskite


photovoltaics
Perovskite semiconductors are an emerging family of photo­ single-junction and multijunction configurations) and illustrates
voltaic materials with the general formula of ABX3. Monovalent the rapid progress in device performance6,162. Promising laboratory-
cations, such as formamidinium (FA+, NH2CH = NH2+), methyl­ scale PV technology has been transferred to upscaled devices
ammonium (MA+, CH3−NH3+), Cs+, Rb+ or a combination thereof, and minimodules with unprecedented PCEs. Part c of the figure
occupy the A-site at the corners of a cube; an octahedral inorganic shows representative record efficiencies with respect to the area
complex of the type [BX6]4−, in which X is a halide anion and B is a (aperture or designated) for large-scale single-junction perovskite
divalent cation (such as Pb2+ or Sn2+), sits at the body centre of the modules, perovskite–perovskite and perovskite–Si tandem
cube. Part a of the figure shows a schematic of the perovskite PV modules. Most recently, single-junction minimodules (active
crystal structure with a general formula ABX3. Formamidinium area of 23.9 cm2) with a PCE of 22.7% and monolithic perovskite–Si
lead triiodide (FAPbI3)-rich perovskites are typically formed upon tandem submodules (active area of 274.2 cm2) with a PCE of 26.8%
incorporating foreign halide perovskites (for example, MAPbBr3 or have been demonstrated7,162. The cost-effective features and
CsPbI3) into the FAPbI3 host matrix at a molar ratio of less than 20%. extended long-term operational stability of perovskite PVs (early
The world has witnessed the rapid advancement of state-of-the-art devices operated for just a few minutes, but devices have now
single-junction and perovskite-based tandem photovoltaics (PVs) been operated for thousands of hours) show promise for the
on a laboratory scale (active area no more than 1.0 cm2). Part b next-generation PVs111,197. These accomplishments reported to
of the figure is a chart of reported power conversion efficiencies date are exclusively correlated with the development of FAPbI3-rich
(PCEs) against year for laboratory-scale perovskite PVs (including perovskites.

a X: I–, Br–, Cl– b 35 Small-area perovskite solar cells (<0.1 cm2) 32.5% HZB
[BX6]4– Large-area perovskite solar cells (~1 cm2)
B: Pb , Sn
2+ 2+
Perovskite–perovskite tandem solar cells (<0.1 cm2)
30 28.0% NJU
Perovskite–Si tandem solar cells (~1 cm2)
A: FA+, MA+, Cs+ Perovskite–CIGS tandem solar cells (~1 cm2) 25.7% UNIST
25
Efficiency (%)

24.2% 23.7% U. Sci. Tech.


20
HZB

15
ABX3
10 12.3% 13.7% 10.4% KHU
ZJU MIT
3.8% Toin U 10.9% IBM
5

2008 2010 2012 2014 2016 2018 2020 2022 2024


Year

c
b
a c 28
Perovskite solar modules
Perovskite–perovskite tandem modules
26 Perovskite–Si tandem modules Oxford PV*

24
EPFL*
Efficiency (%)

22 UNIST
KIER/EPFL
NJU* KIER/EPFL
KAUST
20 WUT UNC
DICP WUT
18 UNC MS* Panasonic*
WUT
16
10 100 1,000
Area (cm2)

Nature Reviews Chemistry


Review article

(continued from previous page)

CIGS, copper–indium–gallium selenide; DICP, Dalian Institute Toin University of Yokohama; UNC, University of North Carolina
of Chemical Physics; EPFL, École Polytechnique Fédérale de at Chapel Hill; UNIST, Ulsan National Institute of Science and
Lausanne; HZB, Helmholtz-Zentrum Berlin; IBM, International Technology; U. Sci. Tech., University of Science and Technology
Business Machines; KAUST, King Abdullah University of Science of China; WUT, Wuhan University of Technology; ZJU, Zhejiang
and Technology; KHU, Kyung Hee University; KIER, Korea Institute University. The symbol * represents certified efficiency. PV modules
of Energy Research; MIT, Massachusetts Institute of Technology; can be classified into minimodules (<200 cm2), submodules
MS, Microquanta Semiconductor; NJU, Nanjing University; Toin U., (200–800 cm2) and panel modules (>800 cm2)162.

dissociate readily and brings additional proton migration37, which may and stable perovskite PV devices11. In this case, tuning the Gibbs free
be accelerated by light illumination38. Notably, the lower polarity and energy and phase transition barrier are considered effective means to
orientational mobility of FA+ relative to MA+ potentially increase the stabilize the photoactive phase.
activation energy (Ea) of ion migration, thereby contributing to rela-
tively stable PV devices39,40. Additionally, the higher Young’s modulus Phase segregation under working conditions
(Table 1) of FAPbI3, in contrast to MAPbI3, is thought to benefit from Beyond the aforementioned phase transition, phase segregation
strengthening the structural–functional integrity against mechanical of FAPbI3-rich perovskites also has a decisive role in dictating the
deformation and mitigating strain propagation under stress, which is long-term stability of PV devices under working conditions. Phase seg-
more pronounced in flexible PV devices41–43. regation at the nanoscale has been revealed as responsible for the large
open-circuit voltage (Voc) deficits observed for perovskite PVs in early
The challenges of stabilization studies10. However, a recent study revealed that the Voc deficit in mixed-
Phase polymorphism of the FAPbI3 halide perovskites is primarily due to a high density of traps before halide
Pristine FAPbI3 is a polymorphic crystalline material. There are four segregation is initiated, which dictates non-radiative recombination48.
crystal phases that exist over a broad range of phase transition tem- In the meantime, efficient charge funnelling formed by phase segre-
peratures (Tc), including the cubic photoactive phase (α-FAPbI3, 390 K), gation into low-bandgap iodine-rich domains was observed, which
hexagonal non-photoactive phase (δ-FAPbI3, 293 K) and two low- selectively enhanced the quantum yield of photoluminescence at low
temperature photoactive phases, that is, tetragonal phase (β-FAPbI3, energies, and partially compensated for the Voc deficit associated with
140 K) and orthorhombic phase (γ-FAPbI3, 91 K)44 (Fig. 1d and Table 1). the narrow-bandgap minority phases49. Additionally, halide segrega-
Among these phases, only the cubic phase with high symmetry is tion was demonstrated to be detrimental for photocarrier extraction
desirable for PV applications. Yet, the thermodynamically metasta- as a result of the funnelled photocarriers being constrained within the
ble cubic phase (α-FAPbI3) can only be transformed at temperatures narrow-bandgap iodine-rich phase, in which the current extraction effi-
above Tc, that is, 390 K (standard atmospheric pressure)9. The change ciency is reduced50. The phase segregation in FAPbI3-rich perovskites is
in Gibbs free energy (ΔG) can describe the thermodynamic phase tran- thought to occur spontaneously and slowly before exposure to external
sition between two crystalline phases45, which is associated with the stressors, including light, heat and bias. A qualitative phase diagram
changes of enthalpy (ΔH) and the Helmholtz free energy (ΔF), that is, of the thermodynamic equilibrium energies in mixed iodide–bromide
ΔG = ΔH + ΔF, where ΔF = −TΔS, ΔS is the entropy change and T is the alloys at 300 K helps to rationalize these observations; a temperature-
temperature (variations in volume and pressure are negligible)9. dependent miscibility gap was revealed to act as a thermodynamic
As the formation enthalpy (ΔHf ) of the cubic phase is about 70 meV driving force for spontaneous halide segregation51. Another potential
(per formula unit cell) higher than that of the hexagonal phase, the source for phase segregation is predicted to occur from local crystal
Gibbs free energy of the cubic phase at absolute zero (0 K) is higher orientation mismatch, producing a local lattice structural strain dur-
than that of the hexagonal phase9,11,39 (Fig. 1g). When the temperature ing the fabrication process52. As presented in Fig. 2a, the disordered
increases, the change in Gibbs free energy between a cubic phase and a domains with nanoscale inhomogeneity landscapes or multiple phases
hexagonal phase originates from the contribution of entropy induced in perovskite alloys support this perspective, in which the spontaneous
by the rotational motion of the FA cations46. Specifically, when the FA phase segregation causes periodically ordered chemically segregated
cations interact with the surrounding [PbI6]4− octahedra through the domains, such as iodide-rich and bromide-rich regions53. The additional
formation of secondary hydrogen bonds47, the thermally activated energies supplied by the external stressors can further overcome the
orientation of the FA cations results in high isotropy and thus high phase segregation barriers by varying the activation energy of ion
entropy. Hence, when the temperature is higher than the Tc (390 K), migration, which has long been the origin of phase segregation under
the increased entropy reduces the Gibbs free energy of the cubic phase working conditions. Hence, analysing the key challenges associated
to a position well below that of the hexagonal phase, benefitting the with phase segregation under working conditions is paramount to
formation of the cubic phase9 (Fig. 1g). Nevertheless, when cooling enhancing both device efficiency and long-term stability.
to room temperature, the FA cations are preferentially orientated Typically, the FAPbI3-rich perovskites can be stabilized with a
in the hexagonal FAPbI3 phase. At this point, the metastable cubic negative Gibbs free energy difference of mixing (ΔGmix) at room tem-
perovskite with a higher Gibbs free energy can readily overcome the perature51,54. However, ΔGmix can become positive and overcome the
phase-transition energy barrier and spontaneously transform back kinetic barriers of phase segregation if perturbed by light illumina-
to the energetically stable and asymmetric hexagonal phase9. Such tion10,55. Density-functional theory calculations revealed that the mixed-
a phase transition of cubic to hexagonal must be avoided in efficient halide perovskites under light illumination have sufficient Helmholtz

Nature Reviews Chemistry


Review article

free energies (ΔFmix) to overcome the entropically driven metastable (0.1–0.5 eV)72,73. Under the applied electrical bias, perovskite lattice
homogeneous states to phase-segregated states51, and the resulting parameters are also observed to change, relating to the migrated ions
perovskite absorbers demonstrate dual-peak emission from band- in the perovskite74. With considerations of ion migration and defects
to-band recombination10,56 (Fig. 2b). Further studies indicated that predominantly present on grain boundaries and surfaces75, the local
the phase segregation on a nanoscale might arise from the localized surface potential analysis demonstrated a higher concentration of
lattice strain induced by the strengthened electron–phonon coupling the positively charged species at grain boundaries than that in grain
after light illumination57. This lattice strain was found in the high spatial centres, which facilitates the formation of iodine-rich phases owing
overlap between the lattice and single-charge density distribution, as to the preferential drift of iodine towards grain boundaries76. These
shown in Fig. 2c57–59. It has also been reported that the phase segregation crystal terminations, therefore, served as channels for the occurrence
under light illumination depends on the type and stoichiometric ratio of phase segregation in perovskite films77. In addition, the different
of A-site compositions, which is related to the formation enthalpy of
mixing (ΔHmix) variation60. For example, the ΔHmix of the mixed A-site
FA1−xCsxPbI3 at 300 K is much larger than that of FA1−xMAxPbI3 or FA1−x−y
CsxMAyPbI3, which is likely to be segregated into FA-rich and Cs-rich
phases under continuous illumination, followed by the formation of
Box 2
non-photoactive hexagonal FAPbI3 and orthorhombic CsPbI3 (ref. 54).
The overall gain in the ΔGmix of FA1−xMAxPbI3 and FA1−x−yCsxMAyPbI3 Key terms
perovskites has been experimentally confirmed to stabilize the pho-
toactive phase during synthesis and operation under illumination13,54. Phase transition: in perovskites, phase transitions between different
A growing consensus is that the broad tail states associated with the crystal variants (including cubic, tetragonal, orthorhombic and
electronic disorder would be enhanced for the perovskites under light hexagonal, as shown in Table 1) are usually achieved through
illumination, increasing defects such as halide vacancy (Vx+) and inter- crystallographic transitions, which involve sudden crystal lattice
stitial (Xi−)61. The interstitial defects usually form metastable neutral rearrangement or a change in symmetry4.
defects (Xi0) by trapping photogenerated electrons or holes and gen-
erating volatile X2 (I2 or Br2) through bimolecular processes62. Finally, Phase segregation: refers to the spatial compositional and
these increased defects, in turn, intensify phase segregation. structural heterogeneity (such as iodine-rich and bromine-rich
The heat effect associated with continuous light illumination is regions, formamidinium-rich and Cs-rich regions) upon light
another predominant factor that accelerates phase segregation. This illumination in mixed-halide or mixed A-site cation perovskites4,10.
is related to overcoming the kinetic barriers to phase segregation by
localized heating or excess energy from hot-carrier thermalization near Formation enthalpy (ΔHf): the standard formation enthalpy is
the illuminated surface to activate halide migration63, thus facilitating defined as the change in enthalpy when one mole of a substance is
phase segregation64,65. Analogously, the light illumination-induced formed from its constituent elements in their standard states, under
heating (55–65 °C for PV panels under outdoor working conditions) has standard conditions11,198.
been shown to activate the lattice expansion66,67 (Fig. 2d). The induced
lattice expansion was proposed to decrease the activation energy of Shockley–Queisser (S–Q) limit (also known as detailed balance
ion migration and be detrimental to phase segregation68. However, theory): describes the maximum theoretical efficiency of a single-
a linear dependency of illumination intensity threshold (Iexc,threshold) junction solar cell when only band-to-band radiative recombination
on temperature indicates that the Iexc,threshold for halide segregation exists in a solar cell23.
increases significantly with increasing temperature65 (Fig. 2e). Never-
theless, which factor dominates the process remains debatable: photo- Goldschmidt tolerance factor (t) and octahedral factor (μ): the
activated halide segregation or thermally activated halide remixing65. probable crystal structure and stability of ABX3 perovskite can be
Hence, decoupling the effects of thermal-driven and photo-driven evaluated by the Goldschmidt tolerance factor t and octahedral
segregation under working conditions is urgently needed. Moreover, factor μ199. t is defined as the ratio of the distance A−X to the
heat-induced decomposition of volatile components of the perovskites distance B−X in an idealized solid-sphere model:
under continuous light illumination, such as MA, preferentially induces
vacancy defects on the film surface, which likewise promotes the phase t = (rA + rX )/[√ 2(rB + rX )],
segregation process69. Some evidence suggests that the migration of VI+
is indeed assisted by coupling with thermally disordered MA+ (ref. 40). where rA, rB and rX are the ionic radii values of the corresponding
The electrical bias, another key stressor, was also thought to components in the ABX3 structure. μ is defined as the ratio of rB/rX
promote chemical inhomogeneities and phase segregation70. Light for the [BX6]4−.
illumination and electrical bias are always concurrently imposed on
the PV devices when assessing their operational stability. The effect Bond energy (E): characterizes the strength of chemical bonds and
of the electrical bias on phase segregation can be analysed simply by can be quantified by the magnitude of energy required to break the
evaluating the ion migration as it is thought to coincide when phase bond. The higher the bond energy, the more stable the structure,
segregation occurs. For example, direct evidence for ion migration was and vice versa.
provided by monitoring I2− and Br2− elemental evolution in the biased
condition, enabling ions to move under the drive of an applied electric Strain: describes the deformation strength of a material when
field71 (Fig. 2f). A-site cations behave similar to halide ion migration, subjected to external or internal stress122.
albeit with higher migration barriers (0.4–0.6 eV) than halide ions

Nature Reviews Chemistry


Review article

a 2.5
UV Visible Infrared b 40
Silicon S–Q limit (300 K)
35 32.9% 33.6% 31.6%
FAPbI3
2.0 MAPbI3 30.5%
28.6%
Irradiance (W m–2 nm–1)

30
CsPbI3
25 Indirect
1.5

PCE (%)
20
1.0 15

10 GaAs c-Si
0.5 CIGS FAPbI3
5 c-InP MAPbI3
CdTe CsPbI3
0.0 0
500 1,000 1,500 2,000 2,500 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Wavelength (nm) Bandgap (eV)

c 28 FAPbI3 d e
Cs/MA/FA/I/Br 151 K
24 0.8~0.9 0.9~1 >1 FAPbI3
0.70 Orthorhombic Ideal cubic Hexagonal
FAPbBr3
20
285 K FAPbCl3
0.65
MAPbI3
Cubic phase (α) Tetragonal phase ( β)

Octahedral factor (µ)


16 MAPbBr3
PCE (%)

0.60
MAPbI3 MAPbCl3
Slowly

Ra coo

12
140 K
cool
390 K

CsPbI3
91 K
pi l

0.55
dl

42 CsPbBr3
y

8 3
K CsPbCl3
0.50
4 MDAPbI3
DMAPbI
Moisture 0.45
0 EAPbI3
2008 2010 2012 2014 2016 2018 2020 2022 GAPbI3
Year 0.40
0.80 0.85 0.90 0.95 1.00 1.05
FAPbI3 CsPbI3 X-site alloying Hexagonal phase (δ) Orthorhombic Tolerance factor (t)
MAPbI3 A-site alloying A-site and X-site alloying phase (γ)

f Phase homogeneous state Phase segregation state g With oriented FA+ With isotropic FA+

Driving
force Br–
Cubic
I–

Hexagonal
T=0 T = Tc T > Tc

Fig. 1 | The evolution of perovskite photovoltaics and characteristic prop- cubic and hexagonal FAPbI3 phases at different temperatures9. The dashed yellow
erties of FAPbI3. a, Optical absorption ranges for silicon and several halide line represents the Gibbs free energy landscape when temperature is 0 K, whereas
perovskites in the solar irradiance spectrum. b, The maximum achievable power the solid blue line represents the Gibbs free energy landscape ascribed to the
conversion efficiency (PCE) as a function of the bandgap for single-junction entropy. FA cations have isotropic orientations in the cubic photoactive FAPbI3
photovoltaics according to the Shockley–Queisser (S–Q) theory. c, The evolution phase above the phase transition temperature (Tc) but acquire a preferred orienta-
trend of perovskite compositions and the increasing prominence of FAPbI3-rich tion in the hexagonal non-photoactive FAPbI3 phase when the temperatures are
perovskites. d, Temperature dependence of phase transition of the crystal struc- lower than Tc. DMA, dimethylammonium; EA, ethylammonium; GA, guanidinium;
ture of the FAPbI3 absorbers. e, Correlations between perovskite components MDA, methylenediammonium; UV, ultraviolet. Part f adapted from ref. 30,
with tolerance factor (t) and octahedral factor (μ). f, Illustration of alloyed perovs- Springer Nature Limited. Part g adapted with permission of AAAS from ref. 9
kite lattices evolving from phase homogeneity to phase segregation under driving © The Authors, some rights reserved; exclusive licensee AAAS. Distributed
forces, such as light and bias30. g, Diagram of the Gibbs free energy variation of the under a CC BY-NC 4.0 licence.

migration and hopping rates of iodine and bromide defects have pro- to its original homogeneous-phase state after eliminating or relaxing
ceeded to generate a vertical gradient distribution in terms of the external stressors10,79. For instance, when halide-segregated perovskite
iodide and bromide ratio, which is another possible driving force for film is placed in the dark, the driven enthalpy of phase segregation may
halide segregation78. be counteracted by mixing entropy and lattice strain relaxation, bringing
Beyond the driving force for phase segregation we discussed earlier, it back to a well-mixed state10. Furthermore, locally isolated halides in the
another intriguing and debatable topic around phase segregation is its mixed iodide–bromide perovskite could be redistributed to an entirely
reversibility: whether phase-segregated inhomogeneous film can return homogeneous state at high carrier–polaron densities30. This observation

Nature Reviews Chemistry


Review article

is considered that the sufficiently high carrier density eliminates the interface80 (Fig. 2h). The fast transformation kinetics is evident at the
strain gradient driving force that induces phase segregation by saturat- localized grain boundaries of small-sized perovskite grains owing to
ing large-area polarons and uniformly deforming the lattice30. However, the increased contact area of moisture and perovskites. When MA+ and
such an explanation remains speculative as it has not been fully verified in Cs+ components are present in the FAPbI3-rich perovskites, their varied
other alloyed compositions, such as the mixed A-site cation perovskites. polarization natures and charge carrier mobilities enable photoactive
On the contrary, an irreversible phase reconstruction process occurs as phase failure in multistep pathways when exposed to a humid working
a chain reaction after phase segregation under continuous light irradia- environment84. Typically, the failure products of the photoactive phase
tion; the charge flow from bromide-rich to iodine-rich domains induces may include: (i) wide-bandgap hydrated phase, that is, PbI2-CH3NH3I-
the expulsion of iodide from the lattice, followed by forming molecu- H2O; (ii) needle-like δ-CsPbI3 and δ-FAPbI3 and (iii) PbI2 (refs. 84,85).
lar iodine79. Overall, further efforts will be necessary to ascertain the Hexagonal polytypes are prevalent during the degradation process
­underlying reasons for phase segregation and its reversibility. in mixed-cation and mixed-halide alloys, such as 2H δ-phase (1D), 4H
and 6H phases (3D)86, which are most detrimental to the operational
Photoactive phase failure at ambient conditions stability. Moreover, moisture and an electrical bias render polarization
In addition to the stressors mentioned earlier, FAPbI3-rich perovskites irreversibly on a timescale of minutes, leading to MA+ drift and perma-
also tend to suffer from cubic-to-hexagonal phase transition and chemi- nent degradation of perovskite films to PbI2 within hours87. Although
cal degradation under moisture and oxygen stressors80,81. In a humid moisture has a decisive role in the photoactive phase failure process,
environment, the FA+ rotational movement is inhibited, allowing the light and bias greatly accelerates this process.
preferential formation of the hexagonal FAPbI3 phase82. Furthermore, In addition, the high affinity of halide vacancies for oxygen mol-
the strong hydrogen bonding between FA+ and water disrupts the weak ecules enables their preferential adsorption and rapid diffusion into
interaction between FA+ and surrounding I− in the [PbI6]4− sublattice perovskite films, where they then form highly reactive superoxide spe-
and induces irreversible degradation from nanoscale hexagonal phase cies (O2−) under light, reaching complete saturation within 10 min81,88.
impurities to PbI2 (ref. 83) (Fig. 2g). Moisture-triggered phase transition The formation of high-density superoxide in pristine MAPbI3 and FAPbI3
kinetics in FAPbI3 strongly depends on perovskite grain sizes, relative demonstrates this oxygen-induced degradation mechanism81 (Fig. 2i).
humidity and moisture exposure time80,84. A (pseudo) first-order reac- Superoxide ions contribute to the formation of extra mobile ions
tion kinetics model established for α-FAPbI3 and δ-FAPbI3 substantiates and ion migration channels, which facilitate the degradation of per-
that a rapid α-to-δ phase transition occurs mediated by a ‘liquid-like’ ovskite films under working conditions89. In this context, fabricating

Table 1 | Lattice parameters and material properties of different perovskites41,44,169–173

Material Crystal structure Tc (K) Space Lattice Octahedral Crystal Electronic Thermal expansion Young’s
group constants (Å) tilting volume (Å3) bandgap (eV) coefficient (α) (K−1) modulus (GPa)

FAPbI3 Cubic (α) 390 Pm3m a = b = c = 6.36 a0a0a0 257.5 1.43–1.50 20.3 × 10−5 16.5
α = β = γ = 90°
Tetragonal (β) 140 P4/mbm a = b = 8.92 a0a0c+ 298.4 ~1.62 20.2–20.6 × 10−5 –
c = 6.33 α = β = 90°
γ = 120°
Orthorhombic (γ) 91 P4/mbm a = 8.88 Disordered tilts – ~1.74  –  –
c = 6.28

Hexagonal (δ) 293 P63 /mmc a = b = 8.66 α = β = 90° 513.3 ~2.17 4.2 × 10−5  –
c = 7.90 γ = 120°
MAPbI3 Cubic (α) 330 Pm3m a = b = c = 6.39 a0a0a0 251.6 1.55–1.60 5.8 × 10−5 14.8
α = β = γ = 90°
Tetragonal (β) 293 I4/mcm a = c = 8.84 a0a0c− 990.0 ~1.62 6.1 × 10−5 14.3
b = 12.70 α = β = γ = 90°
Orthorhombic (γ) 165 Pnma a = 8.84 a+b−b− 959.6 ~1.69 – –
b = 12.58
c = 8.56
CsPbI3 Cubic (α) 593 Pm3m a = b = c = 6.15 α = β = γ = 90° 222.0 ~1.73 11.8 × 10−5 21.6

Tetragonal (β) 533 P4/mbm a = b = 5.97 β = 8°6 243.0 ~1.94 – 19.2


c = 6.27 δ = 0°
Orthorhombic (γ) 448 Pbnm a = 8.63 a0a0a0 947.3 ~1.67 11.8 × 10−5 18.3
b = 8.96 β = 14.75°
c = 12.64 δ = 9.68°
Orthorhombic (δO) 293 Pnma a = 10.40 b−b−a+ 887.8 ~2.82 11.8 × 10−5 –
b = 4.79 β = 11.5°
c = 17.80 δ = 9.9°

Nature Reviews Chemistry


Review article

a b t = –3 min c
t = 2 min
40
t = 3 min
t = 6 min

PL counts (arb. units)


30

20

10

0.95 1.00 1.05


Br:Pb (norm.) 0
600 700 800
Wavelength (nm)

d e f I2– Br2– SnO– reference


100 0.1
1.082
Peak position, Q (Å–1)

Segregation

Intensity/normalized
Iexc, threshold (µW cm–2)

80
Thermal
expansion
1.080
60 0.01

40 No segregation
1.078

20 0.001
20 25 30 35 40 45 50 55 290 300 310 320 330 340 350 360 370 0 2 4 6 8
Temperature, T (°C) Temperature (K) Fluence (1021 ions m–2)

g h α-Phase δ-Phase
i MAPbI3 FAPbI3
PbI2 Water soaking
1.0 1.8

δ phase ~50% RH, 3 days CsMAFA


0.8
Intensity (arb. units)

1.6
Normalized intensity

y = 1 – e–0.042t
Rb
I(t)/I(t0)

~30% RH, >90 days 0.6 R2 = 0.986


1.4
k = 0.042 h–1
0.4 y = e–0.042t
N2, 14 days R2 = 0.986 Cl
1.2
0.2

Fresh Cd
1.0
0.0
10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40 45 0 20 40 60
2 θ (°) Time (h) Time (min)

FAPbI3-rich polycrystalline films with large grains and low defect Optimizing the formation pathway
densities can mitigate ambient atmosphere erosion. Using effective During the perovskite fabrication process, the constituents of the pre-
encapsulation techniques is expected to overcome the challenge of cursor solution strongly influence the preferential transition pathways
atmospheric susceptibility damage90. by modulating the formation energies of photoactive phases. The
enthalpy of formation of the high-temperature cubic phase is greater
Fabrication and stabilization strategies than that of the hexagonal phase from the regular dimethylformamide
The aforementioned energetics control the stability of the photoactive (DMF)-based precursor solution. Thus the hexagonal phases begin to
phase of FAPbI3-rich perovskites and guide the various phase transition form in the solvent-coordinated intermediate phases and speed up
pathways. Thus, the key to achieving an ideal phase equilibrium is to as the solvent evaporates. The cubic phase can be formed upon high-
decrease the formation enthalpy of the photoactive phase during the temperature annealing91. When introducing intermediate phases with
fabrication process, while controlling the bulk and surface Gibbs free favourable energies, it enables the photoactive phases to form below
energy of the photoactive phase, and increasing the energy barriers the phase transition temperature and avoids phase inclusions92. To this
(ΔGc) for the transition from photoactive to non-photoactive phase end, solvent engineering by changing the composition of the interme-
and phase segregation. diate phase in the early crystallization stage has proven to be crucial in

Nature Reviews Chemistry


Review article

Fig. 2 | External stimuli-induced photoactive phase failure mechanisms f, Time-of-flight secondary ion mass spectrometry depth profiles showing the
of FAPbI3-rich perovskites under working conditions. a, Normalized Br:Pb relative secondary ion intensity of iodine and bromine elements evolution across the
ratio map of the FA0.79MA0.16Cs0.05Pb(I0.83Br0.17)3 film extracted by tracking the film depth in the biased condition71. g, Evolution of X-ray diffraction patterns of
characteristic Pb Lα, I Lα and Br Kα lines in nano-X-ray fluorescence53 shows the the FAPbI3 films stored at different relative humidity (RH) levels and photographs
local nanoscale heterogeneity in alloyed perovskites. b, The heterogeneous of the corresponding thin films83. h, Kinetics plots of 1H signal intensity build-
compositions segregate into two distinct phases demonstrated by dual-peak up of −CH sites (FA cations) in δ-FAPbI3 and the simultaneous 1H signal intensity
emission in the in situ photoluminescence (PL) spectrum evolution for X-site decay for the same sites in α-FAPbI3 as a function of exposure time to moisture in
alloyed perovskite under 400 nm pulsed illumination56. c, Snapshot of the 99% the dark at 85% RH80. i, Normalized PL intensity of hydroethidine aliquots in which
isosurface of excess charge density taken from the Gibbs free energy calculations perovskite films were aged under light and oxygen blow at 610 nm, representing
and molecular dynamics simulations58. The driving force for local phase the yield of superoxide generation81. Part a reprinted from ref. 53, Springer Nature
segregation was found to be the strain resulting from the high spatial overlap Limited. Part b reprinted from ref. 56, CC BY 4.0. Part c reprinted with permission
between the lattice and single-charge density distributions. d, The position of from ref. 58, American Chemical Society. Part d reprinted with permission of
the (110) peak in the wide-angle X-ray scattering data of perovskite in the dark AAAS from ref. 67 © The Authors, some rights reserved; exclusive licensee AAAS.
as a function of temperature, showing thermally induced lattice expansion with Part e adapted with permission from ref. 65, American Chemical Society. Part f
increasing temperature67. e, The linear temperature dependency of excitation adapted with permission from ref. 71, Royal Society of Chemistry. Part g adapted
intensity threshold (Iexc,threshold) for halide segregation in MAPbBr1.5I1.5 films confirms with permission from ref. 83, Wiley. Part h reprinted with permission from ref. 80,
the interplay between thermally driven mixing and photo-driven segregation65. American Chemical Society. Part i reprinted from ref. 81, Springer Nature Limited.

acquiring the photoactive phase of FAPbI3-rich perovskites. Strongly appropriate compressive strain (2.4%) thermodynamically favours
polar aprotic solvents, such as dimethyl sulfoxide (DMSO), benefitted the epitaxial growth of single-crystal α-FAPbI3 on a heterogeneous
from their higher donor number (DN, which is a measure of the coor- MAPbCl1.5Br1.5 single-crystal substrate, resulting in increased lattice
dination properties of solvents and solutes) and a shorter Pb–O bond tetragonality (Fig. 3c). Therefore, the applied compressive strain has
length (2.386 Å) compared with DMF (2.431 Å)93, exhibiting stronger a notably positive impact on reducing the lattice mismatching with
coordination ability with the species of perovskite precursors, in par- the substrate99. Nevertheless, the epitaxial growth of phase-pure per-
ticular PbI2 (refs. 19,94). As a result, stable PbI2·DMSO is preferentially ovskite polycrystalline films is still constrained by the limited choices
formed rather than solvent-coordinated perovskite complexes when of underlying materials. Future work must expand the contact layer
DMSO is mixed with the DMF host solvent, dominating the nucleation library to avoid the concomitant growth of harmful phase impurities100.
of hexagonal phases95. This preferred route allows DMSO intercalated Alternatively, developing ‘van der Waals epitaxy’, epitaxy proceeds
in PbI2 to be easily replaced by high-affinity FAI and forms hexagonal through Wan der Waals interactions with large degrees of freedom on
FAPbI3 nuclei, which are then converted to cubic FAPbI3 after annealing lattice mismatches, would be another choice for growing photoactive
(Fig. 3a; ΔHf2), establishing DMSO as one of the most prominent pro- perovskites101.
cessing solvents96,97. Motivated by the pathway associated with forma-
tion enthalpy, other promising phase transition routes are explored. Enhance thermodynamic stability
More stable intermediate states induced by stronger polar solvents, Solid-state alloying, which brings the Goldschmidt tolerance factor to
for instance, N-methylpyrrolidone91, tetramethylene sulfoxide98 and near-optimal values (~0.95), has been thought to be a general design
methylamine formate17, have been reported to assist the direct trans- protocol for forming thermodynamically stable FAPbI3-rich perovskites
formation into photoactive phases from PbI2·solvent intermediate with lower Gibbs free energy of mixing (ΔGmix)11 (Fig. 3d). Reducing
states (Fig. 3a; ΔHf1) while avoiding the energetic pathway involving the lattice constant and increasing bond energy of FAPbI3-rich perovs-
non-photoactive phase which occurs in the DMSO or DMF coordinated kites realized negative formation enthalpy of mixing (ΔHmix) at ~300 K
complexes98 (Fig. 3a; ΔHf2). These intermediate phases, through strong (refs. 13,54). Such reduction in the Gibbs free energy of mixing is also
Pb–O bond and additional N–H···I hydrogen interaction, suppress the thought to be inseparable from the contribution of mixing entropy
generation of hexagonal nuclei during early crystallization, resulting in (ΔSmix); favourable gains in ΔHmix and ΔSmix reduce the phase transition
the formation of the photoactive phases in desirable perovskite films. temperature of δ-to-α as low as 100 °C, such as Cs–FA, MA–FA–I–Br and
Another effective means to facilitate the formation of the photoac- Rb–Cs–MA–FA–I–Br alloys15,102,103. As for the role of substitutable Rb+,
tive phase is associated with strain-controlled heteroepitaxial growth. further research is required to elucidate whether it causes octahedral
The strain originating from the hetero-interface with a higher energy shrinkage15, modifies the grain boundary104 or, conversely, forms extra
state owing to the additional contribution of surface or interfacial Rb-rich phase impurities105. Besides, these FAPbI3-rich perovskites
energy drives local nucleation and epitaxial growth of the cubic FAPbI3 with alloyed small-sized ions suffer widened bandgaps, non-negligible
(ref. 20). A representative example was demonstrated in a layered chemical inhomogeneity on multiscales and phase segregation29,53,
perovskite-templated epitaxial growth that facilitates the phase trans- although the underlying mechanism is still under debate48. Recently,
formation of hexagonal-to-cubic FAPbI3 (ref. 20). The induced hetero- trace alloy incorporation in FAPbI3 has been shown to have a beneficial
interface tensile strain and the increased entropy of FA+ in hexagonal effect on the photoactive phase and thus improve the balance between
FAPbI3 delay the growth of the hexagonal phase, leading to tenfold device performance and stability31,106. For example, a feasible solution
enlarged FAPbI3 perovskite crystals with a reduced defect density and was demonstrated by incorporating a low concentration of MAPbBr3
preferred orientation. Since the as-grown α-FAPbI3 is orientationally (0.8 mol%) to successfully stabilize α-FAPbI3 (ref. 106). Furthermore, the
aligned with the layered perovskite, the heterogeneous interface with partial substitution of FA cations with methylenediammonium cations
applied compressive strain acts as a nucleation site and triggers the (MDA2+), which has a slightly larger ionic radius and more H groups
phase transition of hexagonal to cubic (Fig. 3b). Besides, applying an than FA, was found to sustain the original narrow bandgap of FAPbI3

Nature Reviews Chemistry


Review article

a b c (104) Cl0.60Br2.40 Cl1.50 Br1.50


0.70
Compositions 20 nm 5 nm
+
α (111)
solvents
+ 105
additives ΔGc2

Intensity (a.u.)
ΔGc1 LP (002)

QZ (Å–1)
0.65
PbI2• solvent ΔHf1 [110 ZA] α-FAPbI3
Intermediates 7.24 Å α-FAPbI3 α-FAPbI3
ΔHf2 LP(002)
101

3.11 Å 3.61 Å
α (002) 0% –1.2% –2.4%
α-FAPbI3 (111) 0.60
0.1 0.20 0.1 0.20 0.1 0.20
δ-FAPbI3 QX (Å–1)

d Metastable phase
e f EDTA-induced tilted octahedra
[001]c
ΔGc (α δ)
Free energy

ΔGc
(δ α)
α-FAPbI3
δ-FAPbI3

Stable phase
Free energy

0.7 Å–1

Increase ΔGc
Decrease ΔG Tetragonal phase

g h 1.2 At T = 0 K i
–3.920 Vacuum
0.9 Tensile
Activation energy (eV)

Compressive –3.940
0.6
E (eV)

–3.960

0.3 235 240 245 250


0.651
Hexagonal LP Cubic Nitrogen
0.0
0.648

–0.3 0.645
14.6 14.8 15.0 15.2 15.4
d (A) 247 248 249 250
Compressive strain / Lattice volume (Å3)
/ ΔGc (α δ) ΔGc (δ α) ΔH

Fig. 3 | Stabilized photoactive perovskites with optimal bandgaps. a, The FAPbI3 was formed through surface-bound ethylenediaminetetraacetic acid
differences in Gibbs free energies for the transition from either intermediate (EDTA)-templated crystallization, identified by scanning electron diffraction
adduct or δ-FAPbI3 to α-FAPbI3 (ref. 91). b, The formed hetero-interface patterns near the [001]c zone axis. The superstructure reflections for P4/mbm
between the layered perovskite and α-FAPbI3 in heteroepitaxial growth induces FA-rich perovskites are highlighted with white arrows in scanning electron diffraction
compressive strain, such that the δ-FAPbI3 to α-FAPbI3 phase transition starts at patterns121. g, Schematic of spatially compressive strain engineering. h, Formation
the hetero-interface, and the nucleus formed may grow towards the grain bulk20. enthalpies of different phases and energy barriers from cubic to hexagonal
c, Reciprocal spatial mapping of the (104) asymmetric reflection as a function of the structure under strain-free and strained conditions, respectively20. ΔH, formation
magnitude of compressive strain for α-FAPbI3 films between 0% and −2.4% (ref. 99). enthalpy of the cubic phase relative to the hexagonal phase; d, interlayer spacing
Qx and Qz denote the X-ray scattering vector components in the in-plane and out- of the crystal structure; E, energy of the phase. i, Activation energies for phase
of-plane directions, respectively. d, A schematic (upper panel) shows the Gibbs segregation of alloyed perovskite (FAPbI3)0.8(MAPbBr3)0.2 as a function of lattice
free energy relationship between α-FAPbI3 and δ-FAPbI3 at room temperature. volume under vacuum and nitrogen atmospheres, respectively52. Part a adapted
The metastable α-FAPbI3 with higher Gibbs free energy tends to transform into the with permission of AAAS from ref. 91 © The Authors, some rights reserved;
δ-FAPbI3 with lower Gibbs free energy spontaneously. A schematic (lower panel) exclusive licensee AAAS. Parts b and h adapted from ref. 20, CC BY 4.0. Part c
demonstrates energy-managed phase stabilization strategies by decreasing the reprinted from ref. 99, Springer Nature Limited. Part d adapted with permission
Gibbs free energy difference (ΔG) of the photoactive phase or increasing the of AAAS from ref. 39 © The Authors, some rights reserved; exclusive licensee
energy barrier (ΔGc) for the transition from the photoactive to non-photoactive AAAS. Part e reprinted with permission of AAAS from ref. 113 © The Authors, some
phase39. e, Density-functional theory calculations reveal the strong preferential rights reserved; exclusive licensee AAAS. Part f adapted with permission of AAAS
interaction between lead cations and sulfate anions in the FAPbI3 when the PbI2- from ref. 121 © The Authors, some rights reserved; exclusive licensee AAAS. Part i
terminated surfaces are in contact with PbSO4 (ref. 113). f, Octahedral tilt-stabilized reprinted from ref. 52, Springer Nature Limited.

Nature Reviews Chemistry


Review article

and to avoid the phase segregation issue induced by mixed halide31,107. Increase the energy barrier for phase transition
In-depth studies show that the strong interaction of the MDA2+ with the Having obtained photoactive phases for FAPbI3-rich perovskites, main-
lattice can prevent the phase transition from the cubic to the hexagonal, tenance of their long-term photoactivity under working conditions
offering outstanding phase stability at room temperature. Besides, the becomes crucial to deploying perovskite PVs. The key to achieving this
thermodynamically stable cubic phase can also be achieved by incor- goal is to increase the phase transition energy barrier, ΔGc, from the pho-
porating halogenated organics into the precursor solution, such as toactive to the non-photoactive phase (Fig. 3d, bottom). The alloying
methylammonium chloride (MACl)18,97. One major improvement stems strategies described earlier have led to the production of thermody-
from the formation of ‘intermediate state’ cubic phase FAPbI3-MACl namically stable photoactive phases and a substantial increase in the
before annealing, preventing undesirable hexagonal phase formation18. energy barrier, preventing unexpected phase transitions11. The alloyed
Another benefit is that the degassing of the lightweight Cl− improves FAPbI3-rich perovskites should, in principle, present cubic or pseudo-
the orientation of the cubic FAPbI3 after annealing, and the residual MA+ cubic structures118–120. However, a recent study through nanostructure
(~5 mol%) contributes to the tolerance factor, thus helping reduce the atomic-level characterization techniques has disclosed that FAPbI3-rich
formation enthalpy of the photoactive phase18,32. photoactive perovskites are non-cubic and exhibit ~2° octahedral tilt-
The strategies discussed in this section work to optimize the per- ing associated with the asymmetric tetragonal structure121. This result
ovskite composition and are realized as a result of changes in the Gibbs was corroborated by the presence of the hexagonal phase impurities
free energy of the perovskite bulk material. As the total Gibbs free in the infrequently localized regions without apparent octahedral tilt.
energy of the perovskite is composed of the bulk and surface energies, By comparison with the facile cubic-to-hexagonal transition with a low
researchers have paid more attention to the film surface energy39. Exotic energy barrier height (26 meV), the tetragonal-to-hexagonal transition
additives that are absent at lattice sites (such as K+, Na+, F−, SCN− or has a higher energy barrier (86 meV) and is less likely to occur under
HCOO−) have been actively investigated to strengthen film surfaces similar conditions. The same minor octahedral tilt-stabilized phase has
and grain boundaries at the microscopic level, which offers a solution to been constructed by introducing the surface-bound ethylenediamine-
synthesizing thermodynamically stable photoactive perovskites with- tetraacetic acid (EDTA) chelating agent (Fig. 3f). Although the impact
out bandgap widening18,32,92,108–110. For example, manipulating surface of EDTA on device performance has not yet been investigated, this
Gibbs free energy by using pseudo-halide thiocyanate anions (SCN−) surface coordination approach presents an avenue for the stabilization
has been attempted, which is beneficial in triggering the transition from of desirable photoactive phases.
face-sharing hexagonal to edge-sharing intermediates (4H polytypes of Spatial strain management (Fig. 3g), including microdomain and
FAPbI3) and ultimately corner-sharing cubic FAPbI3 (ref. 92). The edge- macrodomain, is another effective approach to mitigate photoactive
sharing intermediates with iodides replaced by SCN– on the surface of FAPbI3 phase degradation99,122. It has been reported that compres-
hexagonal FAPbI3 drive ions rearrangement and stabilize corner-sharing sive strain applied to a perovskite film can compensate for the nega-
cubic FAPbI3 from surfaces to the bulk. Similar work has also revealed tive impact of tensile strain in the cubic phase lattice, thus raising the
that iodide vacancy defects at grain boundaries and perovskite surfaces transition barrier height of the cubic-to-hexagonal phase transition
can be mitigated with the addition of 2% formate (HCOO−) — a pseudo- by ~0.15 eV and favouring thermodynamic stability below the phase
halide anion — yielding perovskite films with favourable surface ener- transition temperature20 (Fig. 3h). From a microscopic perspective,
gies32. Although these approaches have considerably improved film the mismatched sizes of the FA cation and the octahedral [PbI6]4− units,
quality and device performance, the debate on the roles of these small along with its dynamic disordered nature, is susceptible to microstrain-
foreign ions and their spatial distributions has continued. Further break- induced lattice distortions107, crystal lattice defects81 and lattice ori-
throughs have demonstrated that the surface energy of perovskite films entation mismatch at grain boundaries123. These phenomena are the
can be tuned by forming extra chemical bonds with perovskites77, such thermodynamic driving forces that promote the phase transition to
as ionic bonds111, hydrogen bonds109 or coordination interactions112. For relax the local microstrain. For example, an anisotropic strained lattice
instance, the Pb–S bond, which is much stronger than the Pb–I bond, in the (111) plane was believed to be the direct driving force, acting as
has been successfully used to suppress ion migration and stabilize the the nucleation site for the (001) plane of the δ-FAPbI3 (ref. 124). Partially
perovskite interface, as identified by surface energetics from density- substituting lattice sites with small-sized ions induces lattice shrink-
functional theory calculations and experiment conclusions111,113 (Fig. 3e). age and has a crucial role in compensating the microstrain, stabilizing
The formed compact crystalline PbSO4 phase is well matched to the the metastable pseudo-cubic or cubic phase107. From a macroscopic
underlying perovskite lattice, avoiding the formation of strained per- perspective, the different thermal expansion coefficients of the per-
ovskites and improving phase stability. It is worth noting that decreasing ovskite (~20 × 10−5 K−1) and the adjacent layers (~1 × 10−5 K−1) result in
the dimension of 3D perovskite to 0D nanometre-sized quantum dots tensile strain125. The cubic phase can be effectively maintained by com-
(QDs) demonstrates better phase stabilization than perovskite bulk pensating the tensile strain with the in-plane compressive strain. This
materials owing to the significantly reduced surface energy114. FAPbI3 can be achieved by reducing the interface energy of cubic FAPbI3 with
QD films with stable phases at room temperature have been achieved, the substrate after optimization of the annealing process or contact
which benefit from the dominant contribution of the surface energy and layers99,126. Nevertheless, future efforts to understand how mechanical
lattice contraction115. Nevertheless, so far, perovskite QDs-based solar characteristics influence phases and optoelectronic properties are
cells have yet to match polycrystalline thin-film counterparts in terms of still required.
device efficiency and operational stability. Alternatively, incorporating
QDs such as Cs1−xFAxPbI3 and PbS onto the bulk perovskite surface offers Inhibit the phase segregation
additional benefits in surface energy management116,117. These encour- The key to tackling the long-standing concern of phase segrega-
aging results suggest that regulating the Gibbs free energy of the bulk tion under continuous light illumination is to reduce the possibility
and surfaces are promising directions to enhance the thermodynamic of ion and defect migration. To enhance the inherent phase stabil-
stability of FAPbI3-rich perovskites. ity of FAPbI3-rich perovskites, the judicious selection of lattice-site

Nature Reviews Chemistry


Review article

compositions with low migration capability is required54. For instance, heterogeneities. It is thus important to assess the impact of scalable
using the less polar Cs+ instead of MA+ in FAPbI3 can reduce electron– processes on phase stability and film quality. The current scalable tech-
phonon coupling strength and thus suppress phase segregation57,58. The niques reported for manufacturing perovskite modules can be gener-
enhancement of phase stability was proposed to have another explana- ally classified into solution-phase deposition (including blade-coating,
tion: the migration of iodine vacancies coupled to dynamically oriented slot-die coating, spray-coating and inkjet printing), vapour-phase depo-
MA+ facilitates iodine migration, whereas less polar and orientationally sition methods (vacuum-evaporated deposition and chemical vapour
mobile Cs+ and FA+ slow down the halide migration40. This argument deposition)132 and hybrid deposition that combines solution-phase
is supported by the observed spatially homogeneous halides and Cs+ and vapour-phase approaches. Among these scalable processes, the
after incorporating small amounts of CsI (5 mol%)119. Note that higher Cs differences in perovskite crystallization kinetics could affect the crystal
fractions may be detrimental to compositional homogeneity. However, phases and thin-film quality, thereby influencing device performance.
a contrasting work demonstrated that the presence of the reorientation In the light of solution-phase deposition methods, the key to
of dipolar MA cation shifts the deep trap states to shallower levels owing maintaining stable perovskite absorbers with photoactive phases is
to its higher dipole moment capable of reorienting in response to local to eliminate the degradation sites induced by microscale fluid flow to
electrical field changes caused by adjacent deep traps127. Therefore, a the perovskite nuclei during the drying of large-area wet films with
small amount of MA cation alloying is considered to be beneficial for increased surface tension133. This is achievable by controlling the fluid
healing deep-level defects without adversely affecting photostabil- dynamics by manipulating the viscosity, concentration, solid–liquid
ity. In contrast to the common photo-induced phase segregation that contact angle and surface tension of the perovskite precursor solution.
occurs in mixed iodide−bromide alloys, significant inhibition of phase Incorporating surfactants (for example, l-α-phosphatidylcholine)
segregation in triple-halide alloys (iodine, bromine and chlorine) was is an effective route to lower surface tension and the contact angle
observed128. This may arise from the ability of Cl− to enter the perovs- because of the Marangoni effect134, which is prominent in overcoming
kite lattice (owing to its small size) and improving the thermodynamic uncontrollable increase in film thickness for blade-coating135. Fur-
stability of the triple-halide alloys. However, more evidence is required. thermore, pre-heating (70–145 °C) the substrate is another effective
Further enhancing the phase segregation barrier by increasing means to improve the wettability by promoting uniform removal of
the ion migration activation energy and reducing the number of ion the solvent132. A combined strategy associated with the substrate pre-
migration pathways can minimize the occurrence of localized phase heating and seed-assisted crystallization (KPb2Br5) has recently been
segregation. It was found that incorporating alkali cations, such as Rb+ developed for slot-die coating FAPbI3-rich films with compact and
and K+, increases the formation energy of mobile halide interstitial homogeneous surface coverage, producing stable minimodule (active
defects129, binding and immobilizing undercoordinated halides to area of 57.5 cm2) that retains 82% of their initial PCEs (T82) after 4,800-h
suppress unwanted migration130. Forming larger grains with fewer storage136 (Fig. 4a). Such an impressive long-term shelf life is compa-
grain boundaries that eliminate ion migration pathways is also consid- rable to spin-coating-based perovskite PV minimodules, offering the
ered an effective strategy to alleviate phase segregation. By analogy, promise of scaling-up modules. By contrast, the continuous-operation
single-crystal perovskites that possess reduced grain boundaries and stability of the minimodule manufactured by slot-die or blade-coating
ion migration have better thermal and photostability than the polycrys- is only half as stable as spin-coating137,138 (Fig. 4b). Therefore, there is
talline perovskite counterpart76,131. Hence, developing single-crystal- increasing demand for the understanding of fluid dynamics theory
based perovskites may be another potential approach to acquiring a and the design of suitable approaches for slot-die coating processes to
phase-stable structure. Moreover, recent work has demonstrated that actively overcome solution-phase scalable process challenges, such as
the activation energy of the phase segregation for FAPbI3-rich perovs- non-ideal film thickness caused by a mismatch of solution flow rate and
kites under operation in nitrogen atmospheres is positive, whereas it is coating speed during deposition, the formation of degradation sites
negative in vacuum52 (Fig. 3i), which inspires researchers to focus on the introduced by air entrainment defects or macroscopic voids caused
effect of the working atmosphere on the lifetime of perovskite devices. by discontinuous outflow during the coating process139.
In contrast to scalable solution-phase methods, solvent-free
Upscaling towards commercialization vapour-phase deposition is a high-throughput fabrication technique
Upscaling efficient and stable PVs from large-area devices to solar because it precisely controls the film thickness and shows compatibility
modules with suitable processing techniques has been the focus of with large-scale production140. Nevertheless, to date, and despite the
the current perovskite PV research activity. Although we have broadly availability of efficient upscaling minimodules (PCE of 19.87% with an
discussed a wide variety of approaches to resolving the phase stability aperture area of 14.4 cm2), vapour-phase deposition has only been suc-
and phase segregation challenges in the previous sections, these issues cessfully used to deposit stable perovskites on a small area141. Notably, a
could still occur when processing techniques are transferred from remarkable lifetime of small-area FAPbI3-rich PVs that has been gained by
laboratory-scale cells into industrially viable upscaled modules. The vacuum-evaporated deposition showed negligible efficiency loss (5%)
perovskite phase may differ markedly between small-area devices and after 20,000 h of storage (Fig. 4a). We note that the crystallization pro-
upscaled modules, an effect that is mainly attributed to differences in cess of vacuum-evaporated FAPbI3-rich perovskites is a solid reaction at
the uniformity and crystallization kinetics, as well as the additional high vapour pressures, especially the evaporation of FA-based halides,
degradation pathways introduced by module patterning design (known which is markedly different from solution-phase processes. As such,
as P1, P2 and P3 patterning lines)132. Herein, we examine the impact of paths to control photoactive phases of perovskites in a vacuum cham-
scalable processing techniques on storage and operational stability, ber during the dual-source or multiple-source vacuum co-evaporation
with changes to phase stability and film degradation. must be comprehensively considered. Motivated by this, appropriate
To realize commercialization of perovskite PVs, the first challenge homemade vacuum systems are encouraged to develop by vaporizing
that must be addressed is to uniformly coat a compact perovskite these perovskites in an ultra-high vacuum chamber. In addition, iodide-
absorber with photoactive phases and reduced spatial compositional based organic species present poor stability in an ultra-high vacuum

Nature Reviews Chemistry


Review article

a Dark storage b MPP under AM 1.5 G


104
T95 = 20,000 h, ref. 140 T80 = 5,950 h, ref. 189
T72 = 10,224 h, ref. 136

104
T100 = 7,200 h, ref. 174
T100 = 7,000 h, ref. 175 T90 = 5,040 h, T87 = 2,400 h, ref. 188
T100 = 4,536 h, T82 = 4,800 h,
ref. 176 T90 = 1,600 h, T86 = 2,000 h, ref. 191
ref. 186 ref. 136
T100 = 3,720 h, ref. 187 T96 = 3,600 h, ref. 142 T98 = 1,500 h, ref. 193 ref. 190
T80 = 1,157 h,
T85 = 3,600 h, T95 = 1,187 h,

Working lifetime (h)


T99 = 2,880 h, ref. 179 ref. 192
T80 = 1,500 h, T90 = 1,300 h, ref. 184
Storage lifetime (h)

ref. 185 ref. 137


T93 = 1,500 h, ref. 184 103 T94 = 1,056 h, ref. 138
T92 = 1,000 h, T80 = 1,000 h, T99 = 1,000 h, ref. 195 T91 = 1,000 h, ref. 194
ref. 178
103 ref. 182 ref. 183
T90 = 960 h, T75 = 1,000 h,
T80 = 500 h, ref. 144
ref. 180 ref. 177 T80 = 535 h, ref. 142
T90 = 600 h,
ref. 181 T80 = 388 h, ref. 142

Spin-coating Spray-coating T83 = 200 h,


ref. 143 Spin-coating Spray-coating
Blade-coating Evaporation
102 102 Blade-coating Evaporation
Slot-die CVD
Slot-die CVD

0.1 1 10 100 0.1 1 10 100


Area (cm2) Area (cm2)

Fig. 4 | Long-term lifetime as a function of area (aperture or designated) for tracking at maximum power point (MPP) under continuous AM 1.5 G illumination
laboratory-scale FAPbI3-rich perovskite photovoltaic devices and upscaling following the International Summit on Organic Photovoltaic Stability-L-2
modules produced by various manufacturing techniques. a, Storage lifetime protocol (light-soaking, 65–85 °C). CVD, chemical vapour deposition. Data from
of the perovskite photovoltaic (PV) cells following the International Summit refs. 137,138,142,144,184,188–195.  CVD, chemical vapour deposition; Tx, time
on Organic Photovoltaic Stability-D-1 protocol (dark, room temperature). Data when the power conversion efficiency (PCE) of the PV device decays to x% of the
from refs. 136,140,142,143,174–187. b, Working lifetime of perovskite PV cells initial PCE.

atmosphere, thereby increasing the risk of non-stoichiometric ratios optimal bandgap range defined by the S–Q limit. Compared with the
for resultant FAPbI3-rich perovskites. We note that phase impurities and FAPbI3 polycrystalline film (with a bandgap of ~1.50 eV), a single-crystal
defects caused by non-stoichiometry are typically the main channels for format shows more potential to expand to near-infrared response
ion and defect migration under operating conditions12. Thus, ongoing owing to its narrower bandgap (~1.43 eV)145. Additionally, single-crystal
deposition process improvements are required. Currently, the hybrid perovskite exhibits a lower defect density and fewer phase impurities,
processing technique that combines solution-phase (ultrasonic spray- which is beneficial for phase stabilization against external stimuli146.
coating) and chemical vapour deposition allows modules to have a shelf However, the perovskite PVs based on these single crystal formats
life (T96 = 3,600 h) close to that of solution-based methods142 (Fig. 4a). are still inferior to those of polycrystalline perovskite PVs131. When
Overall, the storage and working lifetimes of minimodules manufac- considering the upscaling techniques, single-crystal perovskite PVs
tured by vapour-phase deposition still lag far behind the solution-phase face a few challenges, such as the manipulation of film thickness and
processed PVs143,144. There is thus plenty of room to improve both the the time-consuming process of large bulk single crystal growth. Simi-
device efficiency and stability of perovskite PV devices. larly, the bandgap of single-crystal-like FAPbI3 powders is estimated to
be ~1.48 eV when processing from solution-phase methods5,107. Hence,
Outlook one possible route to narrowing bandgaps is fabricating preferentially
Preserving the optical activity of halide perovskites is a key factor in oriented single-crystal-like thin film with fewer phase impurities131,145.
achieving the long-term operational stability of perovskite PVs. In this Moreover, strain engineering offers another possibility of bandgap
section, we first analyse the remaining challenges and opportunities for tuning. For instance, by applying a compressive strain, the bandgap of
FAPbI3-rich perovskites, including further reductions to the bandgap, FAPbI3 film has been reduced by ~35 meV99. Therefore, we expect that
uncovering the origin of phase instability with cutting-edge machine there is still plenty of room to enhance the light harvesting of FAPbI3-
learning and in operando characterizations, followed by delineation rich perovskite PVs by reducing the bandgaps that they can eventually
of the stability metrics that must be met for commercial PV. We then achieve device efficiencies approaching the S–Q limit.
highlight the potential future research directions that boost this emerg- Another pressing challenge beyond efficiency enhancement is to
ing PV technology towards commercial deployments, such as upscaling meet the increasing demand for commercial PV stability standards55,84.
perovskite modules, exploring multijunction PVs and unlocking the Although ex situ characterizations are broadly used, in operando
potential of multiple practical applications. techniques can provide rich information on the operation of PVs.
Studying phase evolution against external stressors under operating
Challenges and opportunities conditions can create a coherent and complete picture of the phase-
If the efficiency of single-junction FAPbI3-rich perovskite PVs is to sur- related degradation mechanism of the perovskite PVs147,148 (Fig. 5a,b).
pass that of crystalline silicon PVs, the first challenge is to continue An atomic-level understanding of the phase instability and degradation
narrowing the bandgap value of the light absorber to approach the mechanism still relies on destructive characterization techniques, for

Nature Reviews Chemistry


Review article

a In operando characterizations b Instability mechanism c Machine learning and high-throughput screening


Electron beam
Phase impurity–lattice mismatch Theoretical calculation: rA, rB, rX, ∆H, f(x)
Ion beam Photon
Improve stability

Components Stable
Candidates
X-ray Detector
Decoupling of
light–bias–heat stressors Feedback
Perovskite film

Stable Unstable
photoactive
Working phase
conditions Detector

d Laboratory fabrication e Commercial production f Innovative applications

Precursors
Lab-to-Fab

Direction
Precursors
P1 P2 P3
ON
Precursors OFF
IoT

Spin-coating Slot-die coating Roll-to-roll


Indoor PVs Smart window
91
Fig. 5 | Current challenges and future outlook for stabilized perovskites in are upscaling modules through slot-die coating and flexible module processed
devices. a, Development of in operando characterization techniques could help by roll-to-roll coating196, respectively.) f, Schematic of innovative applications,
to reveal the origin of instability. b, An in-depth understanding of the mechanism such as indoor photovoltaics (PVs), thermochromic perovskites-based smart
of photoactive phase instability is essential. c, Advanced machine learning windows164. IoT, internet of things. Part e (left inset) adapted with permission of
and high-throughput screening are likely to accelerate the discovery of stable AAAS from ref. 91 © The Authors, some rights reserved; exclusive licensee AAAS
perovskites with ideal bandgaps with respect to the Shockley–Queisser limit. and (right inset) adapted from ref. 196, Springer Nature Limited. Part f adapted
d,e, The challenges from the lab-to-fab deposition techniques (insert figures with permission from ref. 164, Elsevier.

example, transmission electron microscopy, and we emphasize that thus preventing iodide oxidation152. Eliminating iodide-related deep-
such approaches hinder the practical structural analysis of phase insta- level traps of FAPbI3-rich perovskites by iodide management in the
bility evolution100,149. As an alternative, cryoelectron microscopy may perovskite precursor solution has provided hints of record-breaking
mitigate electron beam-induced damage with a much higher critical PCEs in perovskite PVs153. The key to improving intrinsic phase stability
electron dose (46 e− Å−2 at 1.49 Å spatial resolution for MAPbBr3) when is to obtain the FAPbI3 photoactive phase at room temperature with
compared with that at room temperature (~11 e− Å−2)150. Despite such lower Gibbs free energy than the non-photoactive phase, a feat that has
success, interpreting multidimensional signals is complicated, given been achieved for MAPbI3 (ref. 24). To obtain thermodynamically stable
the strong correlations between characteristic signatures (such as perovskite crystal structures, integrating versatile machine learning
photon, electron and ion signals) under various working conditions. with graded screening criteria into established workflows based on the
Recently, a reported machine learning workflow promises to separate phase energetics could accelerate the discovery of novel A-site cations
the charging effect from beam destruction during mass spectrom- to replace FA cations or brand-new perovskite compounds154 (Fig. 5c).
etry151. It promises to provide an in-depth understanding of the phase Furthermore, exploring available ligands or bridging agents to stabilize
instability — such as predicting the formation of undesirable hexagonal FAPbI3 photoactive phases will be an ongoing direction encouraged by
phase by-products under operating conditions12, thus helping to guide the recent success of the EDTA chelating agent121. In particular, the ani-
the rational design of efficient and stable perovskite PVs (Fig. 5b). sotropy of Young’s modulus allows the strain to propagate in a preferen-
The intrinsic instability of perovskite absorbers is a long-standing tial direction. The interplay of strain with defects, and the dislocations
challenge that must be circumvented. However, chemical degrada- as a result, may accelerate phase transition and chemical degradation155.
tion is also problematic but can be slowed down by industry-standard After rigorous and long-term operation, the strain propagation caused
encapsulation technology. Remarkably, avoiding FA+ deprotonation by lattice mismatch might become more apparent. Importantly, relax-
and I− oxidation from photochemical and electrochemical reactions ing deleterious strain can manage the propagation of strain in bulk
should be considered to enhance the chemical stability of perovskite and at the interfaces of perovskite films. Better understanding of this
materials. One approach involves incorporating additives with a higher process may well spark future phase engineering approaches.
acidity than FA+ to adjust the chemical environment. Redox reactions
of the FAPbI3-rich perovskites under light and electrical bias could also The keys to practical applications
be mitigated by tailoring the precursor solution. Previous studies have The ultimate destination of FAPbI3 PV technologies is commercial
shown that a Eu3+-Eu2+ ion redox shuttle can recover lead and iodide, deployment, where upscaling will be a critical step in manufacturing

Nature Reviews Chemistry


Review article

PV modules on the production line. It is well known that the nucleation the device stability of tandem and multijunction PVs based on FAPbI3-
and crystal growth of the solution-processed FAPbI3 perovskite films rich perovskites by improving phase stability and suppressing phase
are more complicated during upscaling. The present difficulty lies in segregation of each sub-cell perovskite absorber.
achieving spatially uniform films with photoactive phases (without pho-
toactive inhomogeneities) from laboratory-scale devices conducted Innovative applications of perovskite PVs
by spin-coating to upscaled modules based on scalable deposition Although phase instability is detrimental to long-term operation for
technologies (Fig. 5d,e), which will dictate the commercial viability regular PV applications under sunlight illumination, stimulated phase
of the modules. To reduce the lab-to-fab gap, understanding FAPbI3 transition is potentially interesting for innovative applications, such
crystallization kinetics on large-area substrates and optimization of as smart windows and sensors164. A key metric for such applications
the upscaling deposition process may help to fabricate metre-scale (and arguably more important than device performance) is cost–­
photoactive perovskite PV modules133. Lessons learnt from LaMer and effectiveness, and the easy fabrication of perovskites is helpful here.
Ostwald’s ripening models may help determine the feasible ways to tune Although regular PV applications require a single stable phase under
the spatial uniformity of photoactive phases on upscaling substrates156. sunlight illumination, these applications strongly rely on the reversible
Motivated by these, advances in upscaling perovskite submodules chemical and structural transformations in response to stimuli such
with a 15.3% PCE (with an aperture area of 205 cm2) have been achieved as light, electric fields and the chemical atmosphere. Hence, a critical
using slot-die upscaling coating technology, with suppression of the concern is reliable switching over thousands of cycles. Perovskite PV
hexagonal phase impurities157. Other upscaling techniques, such as cells show great potential for applications in indoor (not in sunlit) con-
the vapour-phase deposition method, have also been used to fabricate ditions. As a result, some prospective applications, including indoor
FAPbI3-rich perovskites. However, their performances lag behind those PVs and all-in-one smart windows (Fig. 5f), are already on the horizon
prepared using solution-based methods and come with an increased of commercialization164–167. Their uses have been extended into satel-
manufacturing cost133. In addition to the efficiency of upscaling PV lites and spacecraft168. Thus, exploring applications associated with the
modules, the operational lifetime associated with phase stability with controlled phase transitions of FAPbI3 or FAPbI3-rich perovskites under
FAPbI3-rich perovskites must also be evaluated by an industrial lifetime various working conditions is a future avenue for research.
assessment test, and the need to establish a link between operational In summary, we believe that exploring and screening perovskite
stability and phase stability through a scientific understanding of the materials with robust photoactive phases will eventually overcome
crystal structure evolution on atomic scales is an urgent one158. Despite their thermodynamic instability and withstand the harsh environment
the great emphasis on FAPbI3-based device stability investigations, for diverse applications. The near future will foresee the great prom-
previous research works have only barely passed the established indus- ise of perovskite PVs with unique advantages for commercialization,
try International Electrotechnical Commission 61215:2016 standards, creating an exciting next-generation energy-harvesting technology
whereby over 95% of their initial performance can be sustained after for a broad range of users.
1,000 h under harsh conditions, such as damp heat test (85% relative
humidity at 85 °C)22. Another possible standard worth considering is Published online: xx xx xxxx
the International Summit on Organic Photovoltaic Stability protocols,
based on the consensus among researchers in the FAPbI3 perovskite References
1. Kojima, A., Teshima, K., Shirai, Y. & Miyasaka, T. Organometal halide perovskites as
PVs on procedures for testing device reliability159. visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc. 131, 6050–6051 (2009).
Regarding practical applications, increasing device performance 2. Koh, T. M. et al. Formamidinium-containing metal-halide: an alternative material for
near-IR absorption perovskite solar cells. J. Phys. Chem. C 118, 16458–16462 (2014).
effectively strengthens cost-effective features for PV panels. Tandem
One of the earliest studies of FAPbI3 in halide perovskite photovoltaic cells.
and multijunction PV technologies tantalizingly offer prospects, 3. Eperon, G. E. et al. Formamidinium lead trihalide: a broadly tunable perovskite for
which can extend beyond the S–Q thermodynamic limit of single- efficient planar heterojunction solar cells. Energy Environ. Sci. 7, 982–988 (2014).
4. Liu, Z. et al. Efficient and stable FA‐rich perovskite photovoltaics: from material
junction PVs128,160. The facile fabrication and tunable bandgap of
properties to device optimization. Adv. Energy Mater. 12, 2200111 (2022).
FAPbI3-rich perovskites render them comparable with the existing 5. Min, H. et al. Perovskite solar cells with atomically coherent interlayers on SnO2
PVs with textured surfaces161. Nowadays, all-perovskite and monolithic electrodes. Nature 598, 444–450 (2021).
6. Best research-cell efficiency chart. National Renewable Energy Laboratory https://www.
perovskite-on-silicon tandem PVs delivered record efficiencies of nrel.gov/pv/assets/pdfs/best-research-cell-efficiencies.pdf (2022).
28.0% (~0.05 cm2) and 31.3% (~1.17 cm2), respectively6,162. Multijunction 7. Ding, Y. et al. Single-crystalline TiO2 nanoparticles for stable and efficient perovskite
PVs made from all-perovskite absorbers and perovskite-on-silicon modules. Nat. Nanotechnol. 17, 598–605 (2022).
8. Wang, R. et al. Prospects for metal halide perovskite-based tandem solar cells.
configurations lead to maximum theoretical efficiencies of 46.9% and Nat. Photon. 15, 411–425 (2021).
49.4% (ref. 163), respectively, which will be a considerable boost. Despite 9. Chen, T. et al. Entropy-driven structural transition and kinetic trapping in formamidinium
the vast promises, several pressing challenges must be solved before lead iodide perovskite. Sci. Adv. 2, e1601650 (2016).
A detailed study of the thermodynamic and kinetic mechanisms of the FAPbI3 phase
the full potential of perovskite tandem and multijunction PVs can be transition.
exploited. For example, the higher bromide fraction (over 20 mol%) 10. Hoke, E. T. et al. Reversible photo-induced trap formation in mixed-halide hybrid
is typically incorporated into the FAPbI3 matrix to widen the band- perovskites for photovoltaics. Chem. Sci. 6, 613–617 (2015).
One of the first reports of the light-induced phase segregation and its reversibility
gap approaching 1.75–2.0 eV for the top-cell absorbers in tandem in mixed halide perovskites.
architecture, but severe phase segregation will be induced163. For the 11. Li, Z. et al. Stabilizing perovskite structures by tuning tolerance factor: formation of
all-perovskite tandem and multijunction PV, the FAPbI3-rich Sn–Pb formamidinium and cesium lead Iodide solid-state alloys. Chem. Mater. 28, 284–292
(2015).
perovskite, which is necessary to achieve the optimum narrow band- Demonstration of the relationship between perovskite component alloying and
gap, presents further challenges with respect to obtaining a stable Goldschmidt tolerance factor, phase transition temperature, system free energy
photoactive phase in a real tandem PV. All of these effects together and perovskite film stability.
12. Macpherson, S. et al. Local nanoscale phase impurities are degradation sites in halide
have undoubtedly become the major roadblock for perovskite-based perovskites. Nature 607, 294–300 (2022).
PV applications. Hence, future efforts must move on to strengthening An excellent example illustrating the impact of phase purity on the stability of perovskites.

Nature Reviews Chemistry


Review article

13. Saliba, M. et al. Cesium-containing triple cation perovskite solar cells: improved stability, A study exploiting voltage loss to reveal the relevance of its mechanism to light-
reproducibility and high efficiency. Energy Environ. Sci. 9, 1989–1997 (2016). induced halide segregation and defects.
14. Fu, C. et al. From structural design to functional construction: amine molecules in high‐ 49. Caprioglio, P. et al. Nano-emitting heterostructures violate optical reciprocity and enable
performance formamidinium‐based perovskite solar cells. Angew. Chem. Int. Ed. 134, efficient photoluminescence in halide-segregated methylammonium-free wide bandgap
e202117067 (2022). perovskites. ACS Energy Lett. 6, 419–428 (2021).
15. Saliba, M. et al. Incorporation of rubidium cations into perovskite solar cells improves 50. Knight, A. J., Patel, J. B., Snaith, H. J., Johnston, M. B. & Herz, L. M. Trap states, electric
photovoltaic performance. Science 354, 206–209 (2016). fields, and phase segregation in mixed‐halide perovskite photovoltaic devices.
16. Draguta, S. et al. Rationalizing the light-induced phase separation of mixed halide Adv. Energy Mater. 10, 1903488 (2020).
organic–inorganic perovskites. Nat. Commun. 8, 200 (2017). 51. Brivio, F., Caetano, C. & Walsh, A. Thermodynamic origin of photoinstability in the CH3NH
17. Hui, W. et al. Stabilizing black-phase formamidinium perovskite formation at room 3Pb(I1−xBrx)3 hybrid halide perovskite alloy. J. Phys. Chem. Lett. 7, 1083–1087 (2016).
temperature and high humidity. Science 371, 1359–1364 (2021). 52. Guo, R. et al. Degradation mechanisms of perovskite solar cells under vacuum and one
18. Kim, M. et al. Methylammonium chloride induces intermediate phase stabilization for atmosphere of nitrogen. Nat. Energy 6, 977–986 (2021).
efficient perovskite solar cells. Joule 3, 2179–2192 (2019). 53. Frohna, K. et al. Nanoscale chemical heterogeneity dominates the optoelectronic
19. Xiang, W. et al. Intermediate phase engineering of halide perovskites for photovoltaics. response of alloyed perovskite solar cells. Nat. Nanotechnol. 17, 190–196 (2022).
Joule 6, 315–339 (2022). 54. Schelhas, L. T. et al. Insights into operational stability and processing of halide perovskite
20. Lee, J.-W. et al. Solid-phase hetero epitaxial growth of α-phase formamidinium active layers. Energy Environ. Sci. 12, 1341–1348 (2019).
perovskite. Nat. Commun. 11, 5514 (2020). 55. Li, N. et al. Microscopic degradation in formamidinium-cesium lead iodide perovskite
21. Mei, A. et al. Stabilizing perovskite solar cells to IEC61215:2016 standards with over solar cells under operational stressors. Joule 4, 1743–1758 (2020).
9000-h operational tracking. Joule 4, 2646–2660 (2020). 56. Motti, S. G. et al. Phase segregation in mixed-halide perovskites affects charge-carrier
22. Holzhey, P. & Saliba, M. A full overview of international standards assessing the long-term dynamics while preserving mobility. Nat. Commun. 12, 6955 (2021).
stability of perovskite solar cells. J. Mater. Chem. A 6, 21794–21808 (2018). 57. Bischak, C. G. et al. Tunable polaron distortions control the extent of halide demixing
23. Shockley, W. & Queisser, H. J. Detailed balance limit of efficiency of p–n junction solar in lead halide perovskites. J. Phys. Chem. Lett. 9, 3998–4005 (2018).
cells. J. Appl. Phys. 32, 510–519 (1961). 58. Bischak, C. G. et al. Origin of reversible photoinduced phase separation in hybrid
24. Wang, K. et al. Isothermally crystallized perovskites at room-temperature. perovskites. Nano Lett. 17, 1028–1033 (2017).
Energy Environ. Sci. 13, 3412–3422 (2020). 59. Wright, A. D. et al. Electron–phonon coupling in hybrid lead halide perovskites.
25. Xiang, W., Liu, S. & Tress, W. A review on the stability of inorganic metal halide perovskites: Nat. Commun. 7, 11755 (2016).
challenges and opportunities for stable solar cells. Energy Environ. Sci. 14, 2090–2113 (2021). 60. Beal, R. E. et al. Structural origins of light-induced phase segregation in organic–
26. Hao, F., Stoumpos, C. C., Chang, R. P. & Kanatzidis, M. G. Anomalous band gap behavior inorganic halide perovskite photovoltaic materials. Matter 2, 207–219 (2020).
in mixed Sn and Pb perovskites enables broadening of absorption spectrum in solar 61. Motti, S. G. et al. Defect activity in lead halide perovskites. Adv. Mater. 31, 1901183 (2019).
cells. J. Am. Chem. Soc. 136, 8094–8099 (2014). 62. Motti, S. G. et al. Controlling competing photochemical reactions stabilizes perovskite
27. Jena, A. K., Kulkarni, A. & Miyasaka, T. Halide perovskite photovoltaics: background, solar cells. Nat. Photon. 13, 532–539 (2019).
status, and future prospects. Chem. Rev. 119, 3036–3103 (2019). 63. Liu, S., Zheng, F., Grinberg, I. & Rappe, A. M. Photoferroelectric and photopiezoelectric
28. Qiu, Z., Li, N., Huang, Z., Chen, Q. & Zhou, H. Recent advances in improving phase properties of organometal halide perovskites. J. Phys. Chem. Lett. 7, 1460–1465 (2016).
stability of perovskite solar cells. Small Methods 4, 1900877 (2020). 64. Vicente, J. R. & Chen, J. Phase segregation and photothermal remixing of mixed-halide
29. Huang, T. et al. Performance-limiting formation dynamics in mixed-halide perovskites. lead perovskites. J. Phys. Chem. Lett. 11, 1802–1807 (2020).
Sci. Adv. 7, eabj1799 (2021). 65. Elmelund, T., Seger, B., Kuno, M. & Kamat, P. V. How interplay between photo and thermal
30. Mao, W. et al. Light-induced reversal of ion segregation in mixed-halide perovskites. activation dictates halide ion segregation in mixed halide perovskites. ACS Energy Lett.
Nat. Mater. 20, 55–61 (2020). 5, 56–63 (2019).
31. Min, H. et al. Efficient, stable solar cells by using inherent bandgap of α-phase 66. Jacobsson, T. J., Schwan, L. J., Ottosson, M., Hagfeldt, A. & Edvinsson, T. Determination
formamidinium lead iodide. Science 366, 749–753 (2019). of thermal expansion coefficients and locating the temperature-induced phase
32. Jeong, J. et al. Pseudo-halide anion engineering for α-FAPbI3 perovskite solar cells. transition in methylammonium lead perovskites using X-ray diffraction. Inorg. Chem. 54,
Nature 592, 381–385 (2021). 10678–10685 (2015).
33. Pellet, N. et al. Mixed-organic-cation perovskite photovoltaics for enhanced solar-light 67. Rolston, N. et al. Comment on ‘Light-induced lattice expansion leads to high-efficiency
harvesting. Angew. Chem. Int. Ed. 53, 3151–3157 (2014). perovskite solar cells’. Science 368, eaay8691 (2020).
34. Kubicki, D. J. et al. Cation dynamics in mixed-cation (MA)x(FA)1−xPbI3 hybrid perovskites 68. Muscarella, L. A. et al. Lattice compression increases the activation barrier for phase
from solid-state NMR. J. Am. Chem. Soc. 139, 10055–10061 (2017). segregation in mixed-halide perovskites. ACS Energy Lett. 5, 3152–3158 (2020).
35. Zhu, H. et al. Screening in crystalline liquids protects energetic carriers in hybrid 69. Conings, B. et al. Intrinsic thermal instability of methylammonium lead trihalide
perovskites. Science 353, 1409–1413 (2016). perovskite. Adv. Energy Mater. 5, 1500477 (2015).
36. Zhu, H. et al. Organic cations might not be essential to the remarkable properties of band 70. Hu, J. et al. Tracking the evolution of materials and interfaces in perovskite solar cells
edge carriers in lead halide perovskites. Adv. Mater. 29, 1603072 (2017). under an electric field. Commun. Mater. 3, 39 (2022).
37. Ceratti, D. R. et al. Eppur si muove: proton diffusion in halide perovskite single crystals. 71. Domanski, K. et al. Migration of cations induces reversible performance losses over
Adv. Mater. 32, 2002467 (2020). day/night cycling in perovskite solar cells. Energy Environ. Sci. 10, 604–613 (2017).
38. Sadhu, S., Buffeteau, T., Sandrez, S., Hirsch, L. & Bassani, D. M. Observing the migration 72. Haruyama, J., Sodeyama, K., Han, L. & Tateyama, Y. First-principles study of ion diffusion
of hydrogen species in hybrid perovskite materials through D/H isotope exchange. in perovskite solar cell sensitizers. J. Am. Chem. Soc. 137, 10048–10051 (2015).
J. Am. Chem. Soc. 142, 10431–10437 (2020). 73. Oranskaia, A., Yin, J., Bakr, O. M., Bredas, J. L. & Mohammed, O. F. Halogen migration in
39. Lee, J. W., Tan, S., Seok, S. I., Yang, Y. & Park, N. G. Rethinking the A cation in halide hybrid perovskites: the organic cation matters. J. Phys. Chem. Lett. 9, 5474–5480 (2018).
perovskites. Science 375, eabj1186 (2022). 74. Heo, S. et al. Enhancement of piezoelectricity in dimensionally engineered metal‐halide
A comprehensive review of the critical role of A-site cations on structural stability, perovskites induced by deep level defects. Adv. Energy Mater. 12, 00181 (2022).
optoelectronic properties and device performance. 75. Meggiolaro, D., Mosconi, E. & De Angelis, F. Formation of surface defects dominates ion
40. Mosconi, E. & De Angelis, F. Mobile ions in organohalide perovskites: interplay of migration in lead-halide perovskites. ACS Energy Lett. 4, 779–785 (2019).
electronic structure and dynamics. ACS Energy Lett. 1, 182–188 (2016). 76. Tang, X. et al. Local observation of phase segregation in mixed-halide perovskite.
41. Ghosh, D., Aziz, A., Dawson, J. A., Walker, A. B. & Islam, M. S. Putting the squeeze on lead Nano Lett. 18, 2172–2178 (2018).
iodide perovskites: pressure-induced effects to tune their structural and optoelectronic 77. Xue, J., Wang, R. & Yang, Y. The surface of halide perovskites from nano to bulk. Nat. Rev.
behavior. Chem. Mater. 31, 4063–4071 (2019). Mater. 5, 809–827 (2020).
42. Chen, B. et al. Large electrostrictive response in lead halide perovskites. Nat. Mater. 17, 78. Barker, A. J. et al. Defect-assisted photoinduced halide segregation in mixed-halide
1020–1026 (2018). perovskite thin films. ACS Energy Lett. 2, 1416–1424 (2017).
43. Chen, H. et al. Flexible optoelectronic devices based on metal halide perovskites. 79. Li, Z. et al. Beyond the phase segregation: probing the irreversible phase reconstruction
Nano Res. 13, 1997–2018 (2020). of mixed-halide perovskites. Adv. Sci. 9, 2103948 (2021).
44. Fabini, D. H. et al. Reentrant structural and optical properties and large positive thermal 80. Raval, P. et al. Understanding instability in formamidinium lead halide perovskites:
expansion in perovskite formamidinium lead iodide. Angew. Chem. Int. Ed. 128, kinetics of transformative reactions at grain and subgrain boundaries. ACS Energy Lett. 7,
15618–15622 (2016). 1534–1543 (2022).
45. Huang, K. Statistical Mechanics 2nd edn (John Wiley & Sons, 1987). 81. Saidaminov, M. I. et al. Suppression of atomic vacancies via incorporation of isovalent
46. Woo, S. J., Lee, E. S., Yoon, M. & Kim, Y. H. Finite-temperature hydrogen adsorption and small ions to increase the stability of halide perovskite solar cells in ambient air.
desorption thermodynamics driven by soft vibration modes. Phys. Rev. Lett. 111, 066102 Nat. Energy 3, 648–654 (2018).
(2013). 82. Hong, M. J., Johnson, R. Y. & Labram, J. G. Impact of moisture on mobility in
47. Lee, J. W. et al. Dynamic structural property of organic–inorganic metal halide perovskite. methylammonium lead iodide and formamidinium lead iodide. J. Phys. Chem. Lett. 11,
iScience 24, 101959 (2021). 4976–4983 (2020).
48. Mahesh, S. et al. Revealing the origin of voltage loss in mixed-halide perovskite solar 83. Yun, J. S. et al. Humidity-induced degradation via grain boundaries of HC(NH2)2PbI3
cells. Energy Environ. Sci. 13, 258–267 (2020). planar perovskite solar cells. Adv. Funct. Mater. 28, 1705363 (2018).

Nature Reviews Chemistry


Review article

84. Ho, K., Wei, M., Sargent, E. H. & Walker, G. C. Grain transformation and degradation 117. Aynehband, S. et al. Solution processing and self-organization of PbS quantum dots
mechanism of formamidinium and cesium lead iodide perovskite under humidity and passivated with formamidinium lead iodide (FAPbI3). ACS Omega 5, 15746–15754 (2020).
light. ACS Energy Lett. 6, 934–940 (2021). 118. Charles, B. et al. Phase behavior and substitution limit of mixed cesium–formamidinium
85. Song, Z. et al. Perovskite solar cell stability in humid air: partially reversible phase lead triiodide perovskites. Chem. Mater. 32, 2282–2291 (2020).
transitions in the PbI2-CH3NH3I-H2O system. Adv. Energy Mater. 6, 1600846 (2016). 119. Correa-Baena, J.-P. et al. Homogenized halides and alkali cation segregation in alloyed
86. Gratia, P. et al. The many faces of mixed ion perovskites: unraveling and understanding organic-inorganic perovskites. Science 363, 627–631 (2019).
the crystallization process. ACS Energy Lett. 2, 2686–2693 (2017). 120. Rehman, W. et al. Photovoltaic mixed-cation lead mixed-halide perovskites: links
One of the first studies to reveal the role of hexagonal polytypes in the crystallization between crystallinity, photo-stability and electronic properties. Energy Environ. Sci. 10,
process of alloyed perovskites. 361–369 (2017).
87. Leijtens, T. et al. Mapping electric field-induced switchable poling and structural 121. Doherty, T. A. S., Nagane, S., Kubicki, D. J. & Jung, Y.-K. Stabilized tilted-octahedra
degradation in hybrid lead halide perovskite thin films. Adv. Energy Mater. 5, 1500962 halide perovskites inhibit local formation of performance-limiting phases. Science 374,
(2015). 1598–1605 (2021).
88. Aristidou, N. et al. Fast oxygen diffusion and iodide defects mediate oxygen-induced A study that revealed the photoactive phase of FAPbI3-rich perovskites is tetragonal.
degradation of perovskite solar cells. Nat. Commun. 8, 15218 (2017). 122. Liu, D. et al. Strain analysis and engineering in halide perovskite photovoltaics. Nat.
89. Lin, D. et al. Ion migration accelerated reaction between oxygen and metal halide Mater. 20, 1337–1346 (2021).
perovskites in light and its suppression by cesium incorporation. Adv. Energy Mater. 11, 123. Sánchez, S. et al. Thermally controlled growth of photoactive FAPbI3 films for highly
2002552 (2021). stable perovskite solar cells. Energy Environ. Sci. 15, 3862–3876 (2022).
90. Ma, S. et al. Development of encapsulation strategies towards the commercialization 124. Zheng, X. et al. Improved phase stability of formamidinium lead triiodide perovskite by
of perovskite solar cells. Energy Environ. Sci. 15, 13–55 (2022). strain relaxation. ACS Energy Lett. 1, 1014–1020 (2016).
91. Bu, T. et al. Lead halide-templated crystallization of methylamine-free perovskite 125. Xue, D. J. et al. Regulating strain in perovskite thin films through charge-transport layers.
for efficient photovoltaic modules. Science 372, 1327–1332 (2021). Nat. Commun. 11, 1514 (2020).
A detailed study of the effect of an additive on the formation pathway of the 126. Zhu, C. et al. Strain engineering in perovskite solar cells and its impacts on carrier
photoactive phase of FAPbI3-rich perovskites. dynamics. Nat. Commun. 10, 815 (2019).
92. Lu, H. et al. Vapor-assisted deposition of highly efficient, stable black-phase FAPbI3 127. Tan, H. et al. Dipolar cations confer defect tolerance in wide-bandgap metal halide
perovskite solar cells. Science 370, eabb8985 (2020). perovskites. Nat. Commun. 9, 3100 (2018).
93. Gutmann, V. Solvent effects on the reactivities of organometallic compounds. Coord. 128. Xu, J. et al. Triple-halide wide-band gap perovskites with suppressed phase segregation
Chem. Rev. 18, 225–255 (1976). for efficient tandems. Science 367, 1097–1104 (2020).
94. Hamill, J. C., Schwartz, J. & Loo, Y.-L. Influence of solvent coordination on hybrid 129. Qiao, L., Fang, W. H., Long, R. & Prezhdo, O. V. Extending carrier lifetimes in lead halide
organic–inorganic perovskite formation. ACS Energy Lett. 3, 92–97 (2017). perovskites with alkali metals by passivating and eliminating halide interstitial defects.
95. Petrov, A. A. et al. Crystal structure of DMF-intermediate phases uncovers the link Angew. Chem. Int. Ed. 132, 4714–4720 (2020).
between CH3NH3PbI3 morphology and precursor stoichiometry. J. Phys. Chem. C 121, 130. Abdi-Jalebi, M. et al. Maximizing and stabilizing luminescence from halide perovskites
20739–20743 (2017). with potassium passivation. Nature 555, 497–501 (2018).
96. Yang, W. S. et al. High-performance photovoltaic perovskite layers fabricated through 131. Alsalloum, A. Y. et al. 22.8%-Efficient single-crystal mixed-cation inverted perovskite solar
intramolecular exchange. Science 348, 1234–1237 (2015). cells with a near-optimal bandgap. Energy Environ. Sci. 14, 2263–2268 (2021).
97. Kim, M. et al. Conformal quantum dot–SnO2 layers as electron transporters for efficient 132. Wang, Y. et al. Printing strategies for scaling-up perovskite solar cells. Natl Sci. Rev. 8,
perovskite solar cells. Science 375, 302–306 (2022). nwab075 (2021).
98. Wang, S. et al. Superior textured film and process tolerance enabled by intermediate- 133. Park, N.-G. & Zhu, K. Scalable fabrication and coating methods for perovskite solar cells
state engineering for high-efficiency perovskite solar cells. Adv. Sci. 7, 1903009 and solar modules. Nat. Rev. Mater. 5, 333–350 (2020).
(2020). 134. Fanton, X. & Cazabat, A. M. Spreading and instabilities induced by a solutal Marangoni
99. Chen, Y. et al. Strain engineering and epitaxial stabilization of halide perovskites. Nature effect. Langmuir 14, 2554–2561 (1998).
577, 209–215 (2020). 135. Krechetnikov, R. & Homsy, G. M. Surfactant effects in the Landau–Levich problem. J. Fluid
An excellent example of growing α-FAPbI3 by epitaxy and applying compressive strain Mech. 559, 429–450 (2006).
to stabilize the photoactive phase. 136. Rana, P. J. S. et al. Alkali additives enable efficient large area (>55 cm2) slot‐die coated
100. Rothmann, M. U. et al. Atomic-scale microstructure of metal halide perovskite. Science perovskite solar modules. Adv. Funct. Mater. 32, 202113026 (2022).
370, eabb5940 (2020). 137. Yang, Z. et al. Slot-die coating large-area formamidinium–cesium perovskite film for
101. Ohring, M. Epitaxy. in Materials Science of Thin Films 417–494 (Elsevier, 2002). efficient and stable parallel solar module. Sci. Adv. 7, eabg3749 (2021).
102. Yi, C. et al. Entropic stabilization of mixed A-cation ABX3 metal halide perovskites for high 138. Deng, Y. et al. Defect compensation in formamidinium–caesium perovskites for highly
performance perovskite solar cells. Energy Environ. Sci. 9, 656–662 (2016). efficient solar mini-modules with improved photostability. Nat. Energy 6, 633–641
103. Jeon, N. J. et al. Compositional engineering of perovskite materials for high-performance (2021).
solar cells. Nature 517, 476–480 (2015). 139. Patidar, R., Burkitt, D., Hooper, K., Richards, D. & Watson, T. Slot-die coating of perovskite
104. Kubicki, D. J. et al. Phase segregation in Cs-, Rb- and K-doped mixed-cation solar cells: an overview. Mater. Today Commun. 22, 100808 (2020).
(MA)x(FA)1−xPbI3 hybrid perovskites from solid-state NMR. J. Am. Chem. Soc. 139, 140. Wang, S. et al. Over 24% efficient MA-free CsxFA1−xPbX3 perovskite solar cells. Joule 6,
14173–14180 (2017). 1344–1356 (2022).
105. Hu, Y., Aygüler, M. F., Petrus, M. L., Bein, T. & Docampo, P. Impact of rubidium and cesium 141. Li, H. et al. Sequential vacuum-evaporated perovskite solar cells with more than 24%
cations on the moisture stability of multiple-cation mixed-halide perovskites. ACS Energy efficiency. Sci. Adv. 8, eabo7422 (2022).
Lett. 2, 2212–2218 (2017). 142. Jiang, Y. et al. Negligible‐Pb‐waste and upscalable perovskite deposition technology
106. Yoo, J. J. et al. Efficient perovskite solar cells via improved carrier management. Nature for high‐operational‐stability perovskite solar modules. Adv. Energy Mater. 9, 1803047
590, 587–593 (2021). (2019).
107. Kim, G. et al. Impact of strain relaxation on performance of α-formamidinium lead iodide 143. Luo, L. et al. Large-area perovskite solar cells with CsxFA1−xPbI3−yBry thin films deposited
perovskite solar cells. Science 370, 108–112 (2020). by a vapor–solid reaction method. J. Mater. Chem. A 6, 21143–21148 (2018).
108. Tang, Z. et al. Modulations of various alkali metal cations on organometal halide 144. Qiu, L. et al. Hybrid chemical vapor deposition enables scalable and stable Cs-FA mixed
perovskites and their influence on photovoltaic performance. Nano Energy 45, 184–192 cation perovskite solar modules with a designated area of 91.8 cm2 approaching 10%
(2018). efficiency. J. Mater. Chem. A 7, 6920–6929 (2019).
109. Li, N. et al. Cation and anion immobilization through chemical bonding enhancement 145. Nazarenko, O., Yakunin, S., Morad, V., Cherniukh, I. & Kovalenko, M. V. Single crystals of
with fluorides for stable halide perovskite solar cells. Nat. Energy 4, 408–415 (2019). caesium formamidinium lead halide perovskites: solution growth and gamma dosimetry.
110. Zhao, Y. et al. Inactive (PbI2)2RbCl stabilizes perovskite films for efficient solar cells. NPG Asia Mater. 9, e373 (2017).
Science 377, 531–534 (2022). 146. Chen, L. et al. Toward long-term stability: single-crystal alloys of cesium-containing
111. Li, X. et al. Constructing heterojunctions by surface sulfidation for efficient inverted mixed cation and mixed halide perovskite. J. Am. Chem. Soc. 141, 1665–1671 (2019).
perovskite solar cells. Science 375, 434–437 (2022). 147. Yang, W. et al. In situ three‐dimensional observation of perovskite crystallization revealed
112. Chen, R. et al. Sulfonate-assisted surface iodide management for high-performance by two‐photon fluorescence imaging. Adv. Opt. Mater. 10, 2200089 (2022).
perovskite solar cells and modules. J. Am. Chem. Soc. 143, 10624–10632 (2021). 148. Qin, M. et al. Precise control of perovskite crystallization kinetics via sequential A-site
113. Yang, S. et al. Stabilizing halide perovskite surfaces for solar cell operation with doping. Adv. Mater. 32, 2004630 (2020).
wide-bandgap lead oxysalts. Science 365, 473–478 (2019). 149. Ran, J. et al. Electron‐beam‐related studies of halide perovskites: challenges and
114. Yuan, J. et al. Metal halide perovskites in quantum dot solar cells: progress and opportunities. Adv. Energy Mater. 10, 1903191 (2020).
prospects. Joule 4, 1160–1185 (2020). 150. Zhang, D. et al. Atomic-resolution transmission electron microscopy of electron beam-
115. Xue, J. et al. Surface ligand management for stable FAPbI3 perovskite quantum dot solar sensitive crystalline materials. Science 359, 675–679 (2018).
cells. Joule 2, 1866–1878 (2018). 151. Higgins, K. et al. Exploration of electrochemical reactions at organic–inorganic halide
116. Que, M. et al. Quantum-dot-induced cesium-rich surface imparts enhanced stability to perovskite interfaces via machine learning in in situ time‐of‐flight secondary ion mass
formamidinium lead iodide perovskite solar cells. ACS Energy Lett. 4, 1970–1975 (2019). spectrometry. Adv. Funct. Mater. 30, 2001995 (2020).

Nature Reviews Chemistry


Review article

152. Wang, L. et al. A Eu3+-Eu2+ ion redox shuttle imparts operational durability to Pb-I 182. Chen, R. et al. Crown ether‐assisted growth and scaling up of FACsPbI3 films for efficient
perovskite solar cells. Science 363, 265–270 (2019). and stable perovskite solar modules. Adv. Funct. Mater. 31, 2008760 (2020).
153. Yang, W. S. et al. Iodide management in formamidinium-lead-halide-based perovskite 183. Zhang, J. et al. Two-step sequential blade-coating of high quality perovskite layers for
layers for efficient solar cells. Science 356, 1376–1379 (2017). efficient solar cells and modules. J. Mater. Chem. A 8, 8447–8454 (2020).
154. Li, J., Pradhan, B., Gaur, S. & Thomas, J. Predictions and strategies learned from machine 184. Du, M. et al. Surface redox engineering of vacuum-deposited NiOx for top-performance
learning to develop high‐performing perovskite solar cells. Adv. Energy Mater. 9, 1901891 perovskite solar cells and modules. Joule 6, 1931–1943 (2022).
(2019). 185. Rolston, N. et al. Rapid open-air fabrication of perovskite solar modules. Joule 4,
155. Sun, S., Fang, Y., Kieslich, G., White, T. J. & Cheetham, A. K. Mechanical properties of 2675–2692 (2020).
organic–inorganic halide perovskites, CH3NH3PbX3(X=I, Br and Cl), by nanoindentation. 186. Feng, J. et al. High-throughput large-area vacuum deposition for high-performance
J. Mater. Chem. A 3, 18450–18455 (2015). formamidine-based perovskite solar cells. Energy Environ. Sci. 14, 3035–3043 (2021).
156. Liu, C., Cheng, Y.-B. & Ge, Z. Understanding of perovskite crystal growth and film 187. Leyden, M. R., Lee, M. V., Raga, S. R. & Qi, Y. Large formamidinium lead trihalide
formation in scalable deposition processes. Chem. Soc. Rev. 49, 1653–1687 (2020). perovskite solar cells using chemical vapor deposition with high reproducibility
157. Bu, T. et al. Modulating crystal growth of formamidinium–caesium perovskites for over and tunable chlorine concentrations. J. Mater. Chem. A 3, 16097–16103 (2015).
200 cm2 photovoltaic sub-modules. Nat. Energy 7, 528–536 (2022). 188. Jiang, Q. et al. Surface reaction for efficient and stable inverted perovskite solar cells.
158. Pescetelli, S. et al. Integration of two-dimensional materials-based perovskite solar Nature 611, 278–283 (2022).
panels into a stand-alone solar farm. Nat. Energy 7, 597–607 (2022). 189. Krishna, A. et al. Nanoscale interfacial engineering enables highly stable and efficient
159. Khenkin, M. V. et al. Consensus statement for stability assessment and reporting for perovskite photovoltaics. Energy Environ. Sci. 14, 5552–5562 (2021).
perovskite photovoltaics based on ISOS procedures. Nat. Energy 5, 35–49 (2020). 190. Su, H. et al. Stable perovskite solar cells with 23.12% efficiency and area over 1 cm2 by
160. Vos, A. D. Detailed balance limit of the efficiency of tandem solar cells. J. Phys. D Appl. an all-in-one strategy. Sci. China Chem. 65, 1321–1329 (2022).
Phys. 13, 839–846 (1980). 191. Liu, Z. et al. A holistic approach to interface stabilization for efficient perovskite solar
161. Lin, R. et al. All-perovskite tandem solar cells with improved grain surface passivation. modules with over 2,000-hour operational stability. Nat. Energy 5, 596–604 (2020).
Nature 603, 73–78 (2022). 192. Tong, G. et al. Scalable fabrication of >90 cm2 perovskite solar modules with >1000 h
162. Green, M. A. et al. Solar cell efficiency tables (version 60). Prog. Photovolt. Res. Appl. 30, operational stability based on the intermediate phase strategy. Adv. Energy Mater. 11,
687–701 (2022). 2003712 (2021).
163. Eperon, G. E., Hörantner, M. T. & Snaith, H. J. Metal halide perovskite tandem and 193. Zhou, T. et al. Crystal growth regulation of 2D/3D perovskite films for solar cells with both
multiple-junction photovoltaics. Nat. Rev. Chem. 1, 0095 (2017). high efficiency and stability. Adv. Mater. 34, 2200705 (2022).
164. Zhumekenov, A. A., Saidaminov, M. I., Mohammed, O. F. & Bakr, O. M. Stimuli-responsive 194. Bi, E. et al. Efficient perovskite solar cell modules with high stability enabled by iodide
switchable halide perovskites: taking advantage of instability. Joule 5, 2027–2046 (2021). diffusion barriers. Joule 3, 2748–2760 (2019).
A comprehensive review of harnessing the reversible chemical and structural 195. Roß, M. et al. Co-evaporated formamidinium lead iodide based perovskites with 1,000 h
transformation properties of perovskites to develop innovative applications. constant stability for fully textured monolithic perovskite/silicon tandem solar cells.
165. Hassan, Y. et al. Ligand-engineered bandgap stability in mixed-halide perovskite LEDs. Adv. Energy Mater. 11, 2101460 (2021).
Nature 591, 72–77 (2021). 196. Snaith, H. J. Present status and future prospects of perovskite photovoltaics. Nat. Mater.
166. She, X.-J. et al. A solvent-based surface cleaning and passivation technique for 17, 372–376 (2018).
suppressing ionic defects in high-mobility perovskite field-effect transistors. Nat. 197. Song, Z. et al. A technoeconomic analysis of perovskite solar module manufacturing
Electron. 3, 694–703 (2020). with low-cost materials and techniques. Energy Environ. Sci. 10, 1297–1305 (2017).
167. Xiao, Z. & Huang, J. Energy‐efficient hybrid perovskite memristors and synaptic devices. 198. Van de Walle, C. G. & Neugebauer, J. First-principles calculations for defects and
Adv. Electron. Mater. 2, 1600100 (2016). impurities: applications to III-nitrides. J. Appl. Phys. 95, 3851–3879 (2004).
168. Kirmani, A. R. et al. Countdown to perovskite space launch: guidelines to performing 199. Green, M. A., Ho-Baillie, A. & Snaith, H. J. The emergence of perovskite solar cells.
relevant radiation-hardness experiments. Joule 6, 1015–1031 (2022). Nat. Photon. 8, 506–514 (2014).
169. Deng, Y.-H. Perovskite decomposition and missing crystal planes in HRTEM. Nature 594,
E6–E7 (2021). Acknowledgements
170. Zhao, J. et al. Strained hybrid perovskite thin films and their impact on the intrinsic W.Z. thanks EPSRC standard research (EP/V027131/1) and Newton Advanced Fellowship
stability of perovskite solar cells. Sci. Adv. 3, eaao5616 (2017). (192097) for financial support. S.I.S. acknowledges financial support from the Basic Science
171. Steele, J. A. et al. Thermal unequilibrium of strained black CsPbI3 thin films. Science 365, Research Program (NRF-2018R1A3B1052820) through the National Research Foundation of
679–684 (2019). Korea (NRF) funded by the Ministry of Science, ICT and Future Planning (MSIP). The authors
172. Sutton, R. J. et al. Cubic or orthorhombic? Revealing the crystal structure of metastable acknowledge D. Liu for advice on figures and writing and T. Webb for revising the final
black-phase CsPbI3 by theory and experiment. ACS Energy Lett. 3, 1787–1794 (2018). manuscript.
173. Fadla, M. A., Bentria, B., Dahame, T. & Benghia, A. First-principles investigation on the
stability and material properties of all-inorganic cesium lead iodide perovskites CsPbI3 Author contributions
polymorphs. Phys. B Condens. Matter 585, 412118 (2020). X.L. wrote the first manuscript, and D.L., M.S., S.I.S. and W.Z. revised the manuscript and
174. Caprioglio, P. et al. Bi-functional interfaces by poly(ionic liquid) treatment in efficient supervised the project. D.L., Z.-H.L., J.S.Y., M.S., S.I.S. and W.Z. contributed to manuscript
pin and nip perovskite solar cells. Energy Environ. Sci. 14, 4508–4522 (2021). editing and comments.
175. Jang, Y.-W. et al. Intact 2D/3D halide junction perovskite solar cells via solid-phase
in-plane growth. Nat. Energy 6, 63–71 (2021). Competing interests
176. Ren, A. et al. Efficient perovskite solar modules with minimized nonradiative The authors declare no competing interests.
recombination and local carrier transport losses. Joule 4, 1263–1277 (2020).
177. Agresti, A. et al. Two-dimensional material interface engineering for efficient perovskite Additional information
large-area modules. ACS Energy Lett. 4, 1862–1871 (2019). Peer review information Nature Reviews Chemistry thanks the anonymous reviewers for their
178. Liang, Q. et al. Manipulating crystallization kinetics in high-performance blade-coated contribution to the peer review of this work.
perovskite solar cells via cosolvent-assisted phase transition. Adv. Mater. 34, e2200276
(2022). Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
179. Zheng, X. et al. Dual functions of crystallization control and defect passivation enabled published maps and institutional affiliations.
by sulfonic zwitterions for stable and efficient perovskite solar cells. Adv. Mater. 30,
e1803428 (2018). Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this
180. Chapagain, S. et al. Direct deposition of nonaqueous SnO2 dispersion by blade coating article under a publishing agreement with the author(s) or other rightsholder(s); author self-
on perovskites for the scalable fabrication of p–i–n perovskite solar cells. ACS Appl. archiving of the accepted manuscript version of this article is solely governed by the terms
Energy Mater. 4, 10477–10483 (2021). of such publishing agreement and applicable law.
181. Xing, Z. et al. A highly tolerant printing for scalable and flexible perovskite solar cells.
Adv. Funct. Mater. 31, 2107726 (2021). © Springer Nature Limited 2023

Nature Reviews Chemistry

You might also like