You are on page 1of 20

REVIEWS

Metal halide perovskite nanostructures


for optoelectronic applications and the
study of physical properties
Yongping Fu   1, Haiming Zhu   2, Jie Chen1,3, Matthew P. Hautzinger   1, X.-Y. Zhu4
and Song Jin   1*
Abstract | Nanostructures of inorganic semiconductors have revolutionized many areas of
electronics, optoelectronics and photonics. The controlled synthesis of semiconductor
nanostructures could lead to novel physical properties, improved optoelectronic device
performance and new areas for exploration. Lead halide perovskites have recently excited the
photovoltaic research community owing to their high solar-​conversion efficiencies and ease of
solution processing; they also hold great promise for optoelectronic applications, such as light-​
emitting diodes and lasers. In this Review , we summarize recent developments in the synthesis
and characterization of metal halide perovskite nanostructures with controllable compositions,
dimensionality , morphologies and orientations. We examine the advantageous optical properties,
improved stability and potential optoelectronic applications of these 1D and 2D single-​crystal
perovskite nanostructures and compare them with those of bulk perovskites and nanostructures
of conventional semiconductors. Studies in which perovskite nanostructures have been used to
study the fundamental physical properties of perovskites are also highlighted. Finally , we discuss
the challenges in realizing halide perovskite nanostructures for optoelectronic and photonic
applications and offer our perspectives on future opportunities and research directions.

The properties and functions of semiconductor nano­ optical and electrical properties, such as high photolu­
structured materials and devices are highly dependent on minescence quantum yields (PLQYs), strong anisotropic
dimensionality, which affects the electron–matter and absorption and/or emission, higher exciton binding ener­
light–matter interactions1. Over the past three decades, gies and widely tuneable bandgaps, and thus could offer
there has been substantial progress in the synthesis, pro­ many exciting and potentially unique opportunities for
cessing and characterization of conventional inorganic optoelectronics. In this Review, we survey the synthesis,
semiconductor (for example, elemental, III–V and metal characterization, properties and optoelectronic applica­
chalcogenide) nanostructures, including nanocrystals, tions of perovskite nanostructures and microstructures,
1
Department of Chemistry, 1D nanowires (NWs) and 2D quantum wells. The major with a focus on 1D nanowires and microwires and 2D
University of Wisconsin– motivation underlying these efforts is to elucidate how quantum wells. Note that we define a structure as a nano­
Madison, Madison, WI, USA.
the dimensionality, size, architecture and composition structure if it has at least one dimension that is <1 µm and
2
Department of Chemistry,
influence the properties of such nanostructures in order a structure as a microstructure if the minimum dimen­
Zhejiang University,
Hangzhou, China.
to design nanostructures with new and/or enhanced sion is >1 µm. We also examine nanostructuring as an
3
International Research
properties and applications2. Semiconductor nanostruc­ approach to accessing metastable polymorphs of bulk
Center for Renewable Energy, tures could potentially revolutionize technological appli­ perovskites and highlight how surface functionalization
State Key Laboratory of cations, such as integrated optoelectronic and photonic can increase the structural stability of perovskite nano­
Multiphase Flow in Power devices, biosensors and energy conversion and storage structures. For discussions on colloidal perovskite
Engineering, Xi’an Jiaotong
devices3,4. nanocrystals, we refer readers to recent reviews11,12.
University, Shaanxi, China.
Metal halide perovskites have emerged as a class of
4
Department of Chemistry,
Columbia University,
semiconductors for high-​performance solution-​processed Structural chemistry and properties
New York, NY, USA. optoelectronics, such as photovoltaics5,6, lasers7–10 and Owing to their crystal structure (Box 1), halide per­
*e-​mail: jin@chem.wisc.edu light-​emitting diodes (LEDs). Compared with bulk per­ ovskites (ABX3, where A is a cation, B is Pb2+ or Sn2+ and
https://doi.org/10.1038/ ovskites and nanostructures of conventional semiconduc­ X is a halide) have vastly different chemistry and physics
s41578-019-0080-9 tors, halide perovskite nanostructures show advantageous from those of conventional inorganic semiconductors.

Nature Reviews | Materials


Reviews

For example, to achieve high performance in semicon­ be easily fabricated from polycrystalline perovskite thin
ductors, it is usually necessary to obtain ultralow den­ films deposited from solution at low temperature and
sities of impurities and defects by carefully controlled are highly efficient despite the high density of defects. It
crystal growth. By contrast, optoelectronic devices can appears that intrinsic point defects in perovskites do not
act as detrimental trap states13. This defect tolerance is
one of the most advantageous photophysical properties
Box 1 | Crystal structures of 2D ruddlesden–Popper and 3D metal halide perovskites
of perovskite semiconductors and is attributed to the
3D perovskites have the general structural formula aBX3 (see the figure, panel a), where a is unusual defect physics and the highly polarizable crystal
a cation, such as methylammonium (CH3NH3+, Ma+), formamidinium (NH2CHNH2+, Fa+) or lattice in the presence of charge carriers14–17. Calculations
Cs+; B is a metal (typically Pb2+ or sn2+); and X is a halide. the crystal structure consists of a have revealed that the most dominant defects occur at
3D network of corner-​sharing [BX6]4− octahedra with a 12-fold coordinated a cation energies near the band edges or within the bands18 owing
occupying the site in the middle of the cube surrounded by 8 octahedra. Compounds with
to the unique antibonding character of the valence band
the aBX3 stoichiometry often have multiple structural polymorphs and complex phase
behaviour, with the crystal structures differing in the connectivity of the [BX6]4− octahedra.
maximum and spin–orbital coupling. In addition, halide
the Goldschmidt tolerance factor (t) provides an empirical rule for predicting the perovskites have soft and polar lattices19 with Young’s
stability and lattice distortion of perovskite structures: t = (rA + rX) ∕  2 (rB + rX), in which moduli approximately ten times lower than those of Si or
ra, rB and rX are the effective radii of the a, B and X ions, respectively. a t value between 0.8 GaAs (ref.20). The soft, polar lattices exhibit soft phonon
and 1 is favourable for cubic (0.9 < t < 1) or distorted (0.8 < t < 0.9) perovskite structures; modes, strong anharmonicity and dynamic disordering.
however, larger (>1) or smaller (<0.8) t values usually result in non-​perovskite structures Moreover, the softness of the perovskite lattice results in
(see the figure, panel b). although there are several aBX3 compounds with various a strong electron–phonon coupling and leads to the for­
cations, only Cs+, Ma+ and Fa+ meet the requirement for the adoption of a 3D perovskite mation of large polarons21, which have been proposed
structure. Other cations are either too small (such as K+ and rb+) or too large (for example, to screen the Coulombic potential and reduce scattering
dimethylammonium (DMa+), ethylammonium (ea+), guanidinium (Ga+) and acetamidinium
with other carriers and charged defects.
(aa+)). Moreover, perovskites at the limit of the tolerance factor requirement, such as
FaPbi3 (t ≈ 1) and CsPbi3 (t ≈ 0.8), are metastable at room temperature.
Compared with conventional semiconductors, the
ruddlesden–Popper (rP) layered perovskites have the general formula (La)2(a)n−1BnX3n+1, distinct surface properties of perovskites that result from
where La+ is a long-​chain alkyl or aromatic ammonium cation and n is an integer (see the the ionic nature and defect tolerance have considerable
figure, panel c). these structures can be regarded as 2D slabs sliced from the corresponding implications for the physical properties of perovskite
cubic 3D perovskite along the (100) crystallographic plane. the extended 2D unit layers of nanostructures. Owing to the increased surface-​to-
inorganic sheets consist of corner-​sharing metal halide octahedra interspaced by the large volume ratio in nanostructures, surface recombination
La+ cations. each inorganic layer of a rP phase has the same thickness, indicated by the n results in a substantial loss of charge carriers and sig­
value. when n is infinite, the structure becomes a 3D perovskite of aBX3. when n is small, nificantly influences the optoelectronic properties of
the structure exhibits intrinsic quantum and dielectric confinement effects owing to the semiconductor nanostructures. However, the surface
few-​unit-cell thickness of the high-​dielectric inorganic layers, which are separated by
recombination velocities of perovskites are several
organic layers with a lower dielectric constant. therefore, these rP perovskite phases can
be considered as bulk assemblies of tuneable isolated quantum wells of 3D perovskites.
orders of magnitude lower than those of conventional
semiconductors22,23. For conventional semiconductor
a b nanostructures, mid-​gap states created by dangling
1.1
bonds on the surface need to be passivated with chem­
ical ligands or protective shells to ensure high perfor­
1.0
mance. By contrast, the surface defects of perovskites
formed by undercoordinated lead atoms create only
3D perovskite
0.9
structure shallow traps24 and can be readily passivated under
t

growth conditions. These features minimize the detri­


0.8 mental effects of an increase in surface area on the pho­
tophysical properties of perovskite nanostructures. As
0.7 an example, single-​crystal MAPbI3 (MA+  = methylam­
A monium) NWs with a diameter of a few hundred nano­
B K+ Rb+ Cs+ MA+ FA+ DMA+ EA+ GA+ AA+ metres exhibit long carrier lifetimes (0.1–1 µs) and long
ABX3 X Size of A cation carrier diffusion lengths (on the order of 10 µm)10,25,26,
substantially exceeding those of most conventional
c RP layered perovskites 3D perovskite semiconductors27,28. However, although perovskites are
more defect tolerant, they are not defect impervious, as
exemplified by the fact that surface ligand functiona­
lization can enhance the light emission properties of
perovskite nanocrystals.
Despite the simplicity of the crystal structure, halide
LA+ ··· perovskites exhibit an enormous variety of structural
modifications and variants with tailored properties.
Bandgap engineering has been a powerful approach
for designing semiconductor nanostructures with
n=1 desired optical properties for specific applications.
Compared with conventional semiconductors, one of
n=2 n=∞
the most attractive features of perovskites is their fac­
n=3 ile bandgap engineering, which can be achieved by

www.nature.com/natrevmats
Reviews

controlling stoichiometry and/or dimensionality 29. method for thin-​film preparation, the underlying con­
For example, changing the halide in ABX3 enables the version pathways have only recently been elucidated10,32.
emission to be tuned from deep blue to near-​infrared At low concentrations of the organic halide salt (AX), the
(IR, 400–820 nm), and changing the metal element organic halide rapidly intercalates into the metal halide,
further extends the emission up to 960 nm. Moreover, resulting in the direct formation of ABX3. However, at
2D quantum-​confined perovskites can be obtained in high concentrations of AX, the solution conversion can
a bulk form — that is, 2D Ruddlesden–Popper (RP) also proceed via a slow dissolution–recrystallization
phases (Box 1) — through the introduction of larger process (Fig. 1a) that involves the formation of a soluble
ammonium cations, such as butylammonium (BA+). intermediate:
Quantum confinement can further shift the bandgap
by >1 eV. Therefore, by changing the well thickness and PbX 2(s) + 2X −(sol) → [PbX 4]2− (sol)
composition, it is possible to tune the emission across the
entire UV to the near-​IR spectrum. Such broad bandgap
tunability has rarely been demonstrated in conventional [PbX 4]2− (sol) + A+(sol) → APbX 3(s) + X −(sol)
semiconductors.
Mechanistic insight into the dissolution–recrystalli­
1D perovskite nanowires zation process enabled the first direct solution growth of
Semiconductor NWs exhibit 1D charge and photon-​ single-​crystal perovskite NWs with suitable dimensions
transport properties that can be exploited for next-​ for NW lasers10. For example, MAPbI3 NWs were syn­
generation optoelectronic and photonic devices. They thesized by placing a lead acetate (PbAc2) film in contact
are also excellent systems for exploring phenomena at with a highly concentrated MAI solution10. A thin poly­
the nanoscale. To date, semiconductor NWs have pri­ crystalline layer of MAPbI3 rapidly forms on the PbAc2
marily been composed of group IV, II–VI and III–V surface and serves as the seed to initiate NW growth.
semiconductors, which have strong and primarily This initial perovskite film also acts as a diffusion bar­
covalent bonds. The ability to synthesize high-​quality rier that prevents further direct reaction between the
single-​crystal semiconductor NWs with controlled solution-​phase MAI and the unreacted PbAc2 under­
diameters, orientation, chemical composition and neath. However, slow dissolution of the unreacted PbAc2
phase has enabled breakthroughs in various functional and the initial perovskite layer to produce the soluble
devices fabricated through bottom-​up approaches3,4. complex ion [PbI4]2− is thermodynamically favourable
Central to this success has been the discovery and devel­ at high concentrations of MAI. The slow generation of
opment of vapour–liquid–solid growth, which enables [PbI4]2− ions thus maintains a low supersaturation con­
1D anisotropic growth of various NW compositions2. dition that promotes 1D anisotropic crystal growth33.
In vapour–liquid–solid growth, a catalytic nanodroplet Spontaneous formation of single-​crystal hollow tubes
of a eutectic liquid alloy adsorbs a vapour precursor to has also been observed, which suggests that the catalyst-​
supersaturation levels, subsequently inducing 1D aniso­ free anisotropic growth of perovskite NWs may be
tropic growth at the liquid–solid interface between the driven by screw dislocations33.
catalyst and semiconductor. The dissolution–recrystallization NW growth
Compared with covalent inorganic semiconduc­ method is generally applicable to a wide range of per­
tors, ionic perovskite lattices are much softer and more ovskites with different cations and/or anions34–36. The
dynamically disordered. The ionic nature results in high keys to successful NW growth using this approach
solubility in water and polar solvents as well as a low include choosing a solvent with a high AX solubility
energy barrier for crystal formation30, which enables but low solubility for the perovskite products, ensur­
low-​temperature synthesis and facile ion exchange. For ing that the concentration of AX is high enough to
example, unlike conventional semiconductors, which chemically transform the solid lead precursors into the
require high-​temperature growth, highly crystalline corresponding perovskites and selecting a suitable lead
perovskites can form within seconds even at room tem­ precursor that can rapidly form an effective perovskite
perature. These unique features of perovskites provide barrier and seeding layer to help maintain low super­
new opportunities for simple and facile nanostructure saturation and induce NW growth. For example, single-​
growth. In this section, we review the various synthetic crystal CsPbBr3 NWs with rectangular cross sections
approaches for growing 1D perovskite NWs and their and well-​defined end facets can be readily grown by
arrays with controlled composition and orientation and reacting a PbX2 film with a CsBr solution in methanol35
discuss the mechanisms that underpin high-​quality 1D (Fig. 1b,c). By adjusting the ratios of different halides
anisotropic growth. or A cations in the precursor solutions, it is possible
to obtain alloyed NWs of mixed halides or mixed A
Solution-​phase recrystallization growth. One attrac­ cations with controlled compositions 34,37. Vertical
tive feature of perovskites is the possibility of using low-​ MAPbBr3 NW arrays can also be grown by depositing
temperature solution-​processing methods to prepare a PbAc2 film on an indium tin oxide–glass substrate
high-​quality single crystals and thin films. Facile solu­ coated with poly(3,4-ethylenedioxythiophene):poly­
tion conversion of metal halides into organic–inorganic styrene sulfonate38. However, with the solution-​phase
hybrid perovskites can be achieved by dipping metal recrystallization method, there is limited control over
halide films into solutions containing organic halide the morphology and size of the products; for example,
salts31. Despite the widespread use of such a conversion microplates are also produced.

Nature Reviews | Materials


Reviews

a Perovskite NW b c

[PbX4]2–

PbAc2
Glass substrate 10 μm 1 μm 1 μm

d e
160

Thickness of PbI2
120

plates (nm)
MAX PbX2 NWs 80
Gas flow or
nanoplates
40

Heating element 40 80 120 160 200 240 280


Thickness of MAPbI3 plates (nm)

f g

)
(100
0)
(01

1]
[00
(110)

10 μm 500 μm

PDMS template
h i

μm
10
Silicon 10
substrate μm

50 nm 0.1 M MABr·PbBr2 50 μm
in DMF

Fig. 1 | Growth and characterization of aPbX3 perovskite NWs and nanoplates. a | Illustration of the growth of
APbX3 (A is an organic cation and X is a halide) perovskite nanowires (NWs) via the dissolution–recrystallization process
(Ac− is acetate). b | Scanning electron microscopy (SEM) image (left) of CsPbBr3 nanostructures grown by the dissolution–
recrystallization process illustrated in panel a. Magnified SEM image (right) showing an array of NWs with flat end facets
and rectangular cross sections. c | Low-​resolution transmission electron microscopy (TEM) image of a representative
CsPbBr3 NW. The inset is the corresponding selected area diffraction pattern along the [110] zone axis (ZA) of the
orthorhombic perovskite structure. d | Illustration of the reaction setup for growing perovskite NWs and nanoplates by
the vapour-​phase conversion method (MA+  = methylammonium). e | Thickness of a series of PbI2 plates before (images
above the line) and after (images below the line) being converted into MAPbI3 using the vapour-​phase conversion method.
The morphology of the nanostructures is preserved after conversion. f | Optical microscopy image (left) of CsPbBr3 NW
networks grown on muscovite mica by vapour-​phase epitaxy. The inset shows a corresponding 2D fast Fourier-​transform
image and demonstrates that these NWs are oriented with a hexagonal symmetry. Illustration of the morphology of these
CsPbBr3 NWs (right). g | SEM image of an array of CsPbBr3 NWs grown on a sapphire substrate. h | TEM image of
colloidal CsPbBr3 NWs synthesized by the hot injection method. The inset shows the emission colour upon UV excitation.
i | Patterned templates can be used to confine perovskite growth and to fabricate perovskite NW arrays on arbitrary
substrates. In the first step (left), a polydimethylsiloxane (PDMS) template with rectangular grooves is pressed into a stock
solution of MABr·PbBr2 in N,N-​dimethylformamide (DMF), which fills the voids. Following evaporation of the solvent, the
PDMS template is removed, leaving the NW array on the silicon substrate. The SEM image (right) shows the morphology
of the resultant horizontally aligned MAPbBr3 NW array. Panel a is adapted from ref.10, Springer Nature Limited.
Panels b and c are adapted with permission from ref.35, ACS. Panel e is adapted with permission from ref.41, Wiley-​VCH.
Panel f is adapted with permission from refs43,48, ACS. Panel g is adapted with permission from ref.52, ACS. Panel h is
adapted with permission from ref.56, ACS. Panel i is adapted with permission from ref.63, ACS.

www.nature.com/natrevmats
Reviews

Vapour-​phase conversion method. The low formation growth, which requires a metal catalyst, NWs of CsPbX3
energy of halide perovskites implies that they are highly can be grown by direct vapour growth without a catalyst,
defective. Schottky-​type defects are exclusively observed which eliminates concern about metal contamination in
with a predicted concentration of >0.4% in MAPbI3 the NW products.
(ref.39). Such high concentration of vacancies enables Recently, vapour-​phase growth has enabled addi­
ions to readily migrate in the lattice through a vacancy-​ tional control of the orientation of perovskite NWs by
mediated diffusion mechanism40. Consequently, ion dif­ using the epitaxial relationship between the perovskite
fusion coefficients in MAPbI3 are remarkably high, for lattice and various substrates. Facile heteroepitaxial
example, ~10−12 cm2 s−1 for I− at room temperature. The growth of perovskite NWs, microwires and nanoplates
high ion mobility presents opportunities for designing on various substrates, such as mica43,46,47, sapphire48,
innovative synthesis approaches. As a notable example, sodium chloride49,50 and oxide perovskites51, has been
MAPbI3 thin films can be easily synthesized by expos­ demonstrated. Examples include the heteroepitaxial
ing PbI2 films in gaseous MAI at elevated temperatures. growth of horizontally oriented CsPbX3 (X− = Cl−, Br−
Compared with the solution-​recrystallization method, or I−) interconnected NW networks with hexagonal
a major difference of the gas-​phase reaction is that the symmetry on mica substrates43,46 (Fig. 1f). The CsPbX3
formation of volatile lead iodide complex ions is highly NWs exhibit a triangular cross section and grow along
unfavourable. The solid–gas reaction instead proceeds the [001] crystallographic orientation of the cubic per­
by the diffusion of MAI through the perovskite layer ovskite phase, exposing side walls that correspond to the
that is initially formed to react with the remaining PbI2 {100} facets. The anisotropic growth of these horizontal
underneath. The reaction temperature is selected to be wires may be due to asymmetric lattice mismatch, which
below the decomposition temperature of the perovskites, leads to unrestricted growth along the [001] direction
but sufficiently high pressures of gaseous MAI and fast but limits the width in the [110] direction43. The lattice
ion diffusion kinetics drive the chemical transformation. mismatch between CsPbCl3 and CsPbI3 is as large as
Vapour-​phase conversion has been applied to syn­ ~10%; therefore, the facile epitaxy of various CsPbX3
thesize perovskite nanostructures by reacting pre-​grown perovskites on mica suggests that the soft perovskite
lead halide nanostructures with gaseous MAI (ref.41) lattice can tolerate a larger lattice mismatch than tradi­
(Fig. 1d). For example, PbI2 nanoplates were grown on tional semiconductors. In addition, the facile synthesis
mica by physical vapour deposition and then converted of large-​scale NW arrays or interconnected networks by
into MAPbI3 nanoplates by intercalating MAI into the epitaxial growth could facilitate direct integration into
layered PbI2 structures. After conversion, the thick­ practical devices48,52.
nesses of the nanoplates increased by a factor of 1.8, but Graphoepitaxy on a faceted sapphire plane is another
the lateral dimensions and morphology were retained approach used to guide the growth of CsPbBr3 NWs with
(Fig. 1e). Perovskite NWs can also be synthesized using the lengths on the scale of millimetres48,52 (Fig. 1g). Owing to
vapour-​phase approach, with the crucial step being strain accumulation, conventional NWs (for example,
the growth of PbI2 NWs. Although the single-​crystal CdS) grown by graphoepitaxy often have a high degree
PbI2 NWs are grown by vapour-​phase deposition, the of non-​uniformity in their crystallographic orientation
converted perovskite NWs are polycrystalline with and rougher surfaces than NWs grown on flat surfaces.
randomly oriented domains and exhibit much rougher By contrast, guided CsPbBr3 NWs have the same crys­
surfaces than those of solution-​grown NWs, probably tallographic orientation and surface smoothness as the
owing to the strain caused by volume expansion during analogous NWs grown on flat substrates. Again, this is
the conversion42. However, the polycrystalline nature attributable to the soft perovskite lattice, which enables
may in turn facilitate the conversion to the perovskite the perovskite NWs to easily compensate for the strain
by improving ion diffusion through grain boundaries. and hence adapt to different substrate morphologies.
Vapour-​phase conversion is also a promising method for
synthesizing heterostructures of perovskites with other Colloidal nanowire synthesis. Following the success
materials of the colloidal synthesis of perovskite nanocrystals53,
morphological engineering into 1D NWs has been
Direct vapour-​phase growth. Traditional semiconduc­ achieved by modifying the reaction conditions. These
tor NWs formed by high-​temperature vapour-​phase colloidal perovskite NWs generally have much narrower
growth are typically of superior crystalline quality diameters than the NWs synthesized by other methods.
and have lower defect densities than those grown by NWs of MAPbI3 were first observed among nanocrystals
solution methods. Direct vapour-​phase growth of synthesized by an antisolvent recrystallization method54.
hybrid perovskites is more challenging because ther­ Owing to the ionic nature of perovskites, they are highly
mal decomposition can occur before vaporization. soluble in polar aprotic solvents but almost insoluble in
However, vapour-​phase synthesis is an attractive option the less polar antisolvent. Upon adding a perovskite
for growing nanostructures of more thermally stable solution into an antisolvent, the decrease in solubility
all-​inorganic perovskites43–45. CsPbX3 has a lower melt­ leads to high supersaturation and the rapid precipitation
ing point (~450 °C) than conventional semiconductors of perovskite nanocrystals. However, with this method,
(for example, ~1,400 °C for Si and ~1,200 °C for GaAs), there is poor control of the morphology of the prod­
which enables CsPbX3 NWs to be synthesized through uct54. The hot injection method is a more controllable
a relatively low-​temperature physical vapour deposition synthetic route for the preparation of colloidal per­
process. Moreover, unlike vapour–liquid–solid NW ovskite NWs55–57, including high-​aspect-ratio CsPbBr3

Nature Reviews | Materials


Reviews

NWs with a diameter of ~10 nm and lengths of up to MAPbI3 can be readily converted into NWs by DMF
10 μm (Fig. 1h), as well as quantum wires with a diame­ treatment71,72, and submillimetre-​scale MAPbI3 micro­
ter ~2.2 nm less than the exciton Bohr radius (~7 nm). wires have been obtained within several minutes by
These ultrathin quantum wires show significantly blue-​ adding a small amount of DMF during the solution-​
shifted photoluminescence (PL) and a larger Stokes shift phase recrystallization73. In the latter case, microwires
than that of bulk-​like NWs (>10 nm). of an intermediate phase consisting of PbI2–MAI–DMF
Colloidal NWs of perovskites are highly dynamic, adducts rapidly formed and were converted into
as demonstrated by the morphology evolution during the perovskite phase upon isolation from solution. The
the reaction55, which is probably attributable to dynamic resulting MAPbI3 products retain the 1D morphology
ligand binding to the surfaces. In another method, of the intermediate adducts. Despite the facile synthesis,
CsPbX3 NWs were synthesized via transformation of NWs grown via the formation of intermediate adducts
the CsPbX3 nanocrystals that initially form under pro­ are polycrystalline and exhibit a rough surface and poor
longed ultrasonication of the solid perovskite precursors uniformity, which make them less suitable for optical
in a mixture of solvent and ligand58. Although the NW and photonic applications.
growth mechanism in colloidal synthesis is not fully
understood, the surface ligands have a crucial role in Ion exchange of existing perovskite nanowires. Owing
enabling a 1D morphology59. The dynamic structural to the large surface-​to-volume ratio, ion-​exchange reac­
transformation between different morphologies implies tions in NWs are generally faster than those in bulk
a complicated and kinetically driven NW growth mecha­ materials. In metal chalcogenide NWs, examples of
nism. Recent studies suggest that 1D growth may occur complete chemical transformations are limited to NWs
through self-​assembly and oriented attachment of the with small diameters (up to several tens of nanometres).
initial cuboidal nanocrystals58. The long-​term retention Partial exchange reactions that result in heterogeneity
of structural integrity and optical properties of these col­ are very common in NWs74. However, the dynamic
loidal NWs during storage or processing is a problem for and defective halide perovskite lattice promotes the
their potential practical application. ion-​migration kinetics. Consequently, ion-​exchange
reactions in halide perovskites are surprisingly fast,
Space-​confined nanowire growth. Space-​confined even at room temperature75, and can occur at solid–
growth is a promising route to creating arrays of per­ liquid, solid–gas and solid–solid interfaces. For example,
ovskite NWs with controlled dimensions and precise starting from wires of CsPbBr3 (with diameters of up to
positions. Dense vertical perovskite NW arrays were 1 µm), the conversion into CsPbI3 or CsPbCl3 wires or
formed inside anodized aluminium oxide templates mixed halide alloys by anion exchange was completed
by confining the crystallization of perovskites during within a few minutes35,56. This anion exchange is an alter­
either solution-​phase or vapour-​phase deposition60,61. native approach to accessing a series of NWs of metasta­
The templates may further act as stabilizing scaffolds to ble perovskite phases, including CsPbBr3−xIx and CsPbI3,
protect perovskites from moisture and air. Anodized alu­ which are difficult to directly synthesize owing to phase
minium oxide templates can also be used to extrude the instability. The NWs formed through anion exchange
perovskite solution to yield free-​standing quasi-​vertical showed well-​preserved morphologies and uniform alloy­
NW arrays by applying a pressure gradient62. Large-​scale ing of halides in the alloyed NWs, which is an indication
oriented perovskite NW arrays on arbitrary substrates of the soft perovskite lattice. Ionic exchange provides a
were fabricated by combining a top-​down microfabrica­ powerful method for accessing NWs with arbitrary
tion technique with bottom-​up solution growth63–65. For compositions and for constructing heterostructures.
example, a polydimethylsiloxane template with rectan­
gular grooves was used to spatially confine the crystal­ Nanowire heterostructures. Creating functional het­
lization of MAPbBr3, leading to directional NW growth erojunctions in perovskite NWs is important for con­
along the grooves (Fig. 1i). The dimensions, positions and structing advanced optoelectronic and photonic circuits.
density of the NWs can be controlled by adjusting the The diversity of perovskites could enable facile heteroep­
groove templates and the precursor concentration. This itaxial integration of multifunctional materials in single
space-​confined strategy has also enabled the growth NWs. Moreover, bandgap mismatch at the junction gives
of nanoplate arrays 66 and large-​area single-​c rystal rise to the possibility of controlling and manipulating the
films with up to submillimetre lateral dimensions and generation, recombination and transportation of charge
thicknesses of tens to hundreds of nanometres67. carriers. The investigation of charge-​carrier dynamics
across heterointerfaces is also crucial to the design and/or
Nanowire growth via intermediate adducts. Aniso­­­tropic improvement of solar cells and optoelectronic devices.
crystallization of MAPbI3 from N,N-​dimethylformamide Despite the promise of multifunctional perovskite
(DMF) or dimethylsulfoxide solution is commonly heterostructures, the facile ion migration in 3D per­
observed and probably proceeds via formation of an ovskites imposes challenges in creating stable hetero­
intermediate adduct. Direct evaporation from a thin junctions. Varying the reactants during growth can yield
layer of saturated solution of MAPbI3 on a glass sub­ stable and sharp heterojunctions in traditional semicon­
strate often yields microwires and NWs68. By modifying ductor NWs2,3. By contrast, 3D perovskites can rapidly
the crystallization conditions, this inherent anisotropic exchange their halide anions with the surrounding ions,
growth can be controlled to prepare large-​area aligned which leads to the formation of homogeneous alloys.
MAPbI3 microwire arrays69,70. Moreover, thin films of A spatially selective conversion of CsPbBr3 NWs into

www.nature.com/natrevmats
Reviews

e
100 p–n heterojunction
a c
80
Y-CsSnI3 B-CsSnI3

Current (pA)
60 EC
1.3 eV
40 EF
2.5 eV
20
EV
3 μm 10 μm 0

–3 –2 –1 0 1 2 3
b d Voltage (V)
495 600
– Electron f
+ Hole
490 550 Surface potential (mV)
Height (nm)

485 500 – –
– – – – – – – 5 μm
CB
480 450 VB + + + + + + + + +
400
1 μm
475
Narrow bandgap Wide bandgap
3.0 4.5 6.0 7.5 9.0
Length (µm)

Fig. 2 | Growth and characterization of perovskite NW heterostructures. a | Optical image of a CsPbBr3/CsPbCl3


nanowire (NW) heterostructure with a sharp junction under laser excitation. b | Surface potential profile (red) across the
junction in the NW heterostructure shown in panel a with the corresponding atomic force microscopy height profile
(black). c | Optical image of a CsPbBr3/CsPbCl3 NW heterostructure with a gradient junction under laser excitation.
d | Illustration of the charge-​carrier-funnelling process in a halide-​gradient NW heterostructure. e | Current–voltage curve
of a p–n heterojunction in a CsSnI3 NW, showing current-​rectifying functionality. The p–n heterojunction is formed through
a localized phase transition between a black perovskite phase (B-​CsSnI3, p-​type) and a yellow non-​perovskite phase
(Y-​CsSnI3, n-​type). The inset shows the energy band alignment under thermal equilibrium. f | The p–n heterojunctions within
the CsSnI3 NW described in panel e can be observed using scanning electron microscopy (top) and cathodoluminescence
microscopy (bottom). The dark–bright segments correspond to the Y-​CsSnI3 and B-​CsSnI3 phases, respectively , within a
CsSnI3 NW. EC, energy of the conduction band (CB) minimum; EF, Fermi energy ; EV, energy of the valence band (VB)
maximum. Panels a and b are adapted with permission from ref.76, PNAS. Panel c is adapted with permission from ref.77,
ACS. Panel d is adapted with permission from ref.78, ACS. Panels e and f are adapted with permission from ref.80, PNAS.

CsPbBr3/CsPbCl3 NW heterostructures was achieved by perovskite phase of CsSnI3 is readily observed under
carrying out a local anion-​exchange reaction76. The het­ cathodoluminescence microscopy (Fig. 2f).
erostructure features two-​colour emission with a sharp
interface (Fig. 2a). An abrupt change of ~190 mV in the 2D quantum wells
relative surface potential is observed across the hetero­ 2D quantum wells are another type of nanostruc­
junction interface, confirming that the two segments ture with promise for optoelectronic applications81,82.
have distinct electronic properties (Fig. 2b). However, Quantum-​confined perovskites present different opto­
solid-​state halide interdiffusion across the heterojunc­ electronic features from those of bulk 3D perovskites,
tion can readily occur, and this phenomenon can be used such as stronger excitonic character and anisotropic
to produce axial NW heterostructures with a bandgap absorption and emission. RP perovskites (Box 1) can be
energy gradient77,78 (Fig. 2c). For example, a NW with a regarded as self-​assembled quantum wells or superlat­
halide gradient, MAPbBrxI3−x, was obtained by placing a tices of 3D perovskites with an intrinsic quantum well
MAPbBr3 single crystal in contact with a MAPbI3 NW thickness of a single or a few unit cells. In this section,
at one end. The resulting NW heterostructure forms an we discuss the chemistry, optical properties and recent
energy funnel along the axial direction, which drives advances in the synthesis of 2D perovskites. Note that the
the unidirectional transport of charge carriers from the term 2D can refer here to the intrinsic 2D crystal struc­
wide-​bandgap region to the narrow-​bandgap region tures of RP perovskites or the morphological dimen­
(Fig. 2d). An interesting approach to creating heterojunc­ sions, such as nanosheets or nanoplates of perovskite
tions is to use light-​induced phase segregation in mixed materials (2D or 3D).
halide perovskites. For example, under photoexcitation,
a heterojunction forms between the I-​rich and Br-​rich Bulk Ruddlesden–Popper perovskites. The n = 1  RP
phases79 that result from the spinodal decomposition perovskites (where n indicates the thickness of the
of homogeneous MAPbBrxI3−x alloys. Very recently, a inorganic layers, Box 1) are the most extensively stud­
current-​rectifying p–n junction was created in a sin­ ied 2D perovskite structures, with numerous divalent
gle CsSnI3 NW (Fig. 2e) through a localized thermally metal and organic ammonium combinations demon­
driven phase transition80. Heterojunction formation strated29. However, the synthesis of phase-​pure RP per­
between the yellow non-​perovskite phase and black ovskites with higher n is nontrivial owing to the complex

Nature Reviews | Materials


Reviews

a Confined 2D b 80 c

+ –e– 60

PLQY (%)
E 40
Bulk 3D 2.5
2D perovskite
2.0
Organic emitter

Height (nm)
20 1.5
3D perovskite 1.0
Inorganic (bulk GaN) 0.5
+ – 0 Perovskite CQDs 0.0
–0.5
– Electron 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 0 1 2 3 4
E + Hole Diameter (μm)
1/FWHM (100 nm–1)

10 μm

Tune halide composition (PEA)2PbI4 Tune dimensionality (n)


e f h i
572 nm 640 nm 1
Intensity (arb. units)

Order 2 UC
0 1.3nm
1
3 UC
0
2 μm

PL and OD (arb. units)


1

4 UC 0 6 nm
400 500 600 700 800 0
Wavelength (nm) 1
Laser
g annealing 5 UC
0 2.5nm
1
Disorder
6 UC
0
1 2 μm
Cubes
0 10 nm
0
50 nm 350 400 450 500 550 600
Wavelength (nm)

Fig. 3 | Growth and characterization of 2D perovskite nanostructures. a | Illustration of quantum confinement and
dielectric confinement in Ruddlesden–Popper (RP) perovskites or nanoplates versus bulk crystals. Quantum-​confinement
effects increase the bandgap and the exciton binding energy of RP perovskites or colloidal nanoplates compared with
3D perovskites. Moreover, the exciton binding energy is further increased owing to dielectric confinement. In RP

perovskites or colloidal nanoplates, the electric field (E ) generated by charges in the inorganic framework (which has a
high dielectric constant) can extend into the organic alkylammonium spacers (which have a much lower dielectric
constant), in which it is screened less effectively. b | Comparison of the photoluminescence quantum yield (PLQY) against
the reciprocal of the full width at half maximum (FWHM) of the emission peak of a 2D (PEA)2PbBr4 (PEA+  = phenylethylam-
monium) single crystal and previously reported blue-​light emitters. c | Atomic force microscopy (AFM) image and height
profile of several (BA)2PbBr4 (BA+  = butylammonium) nanosheets synthesized by solvent evaporation. d | Optical images
of a series of individual microplates of (PEA)2(MA)nPbnI3n+1 (MA+  = methylammonium) with different halide compositions
and thicknesses (n) under laser excitation. The images show that it is possible to tune the emission colour and that
these microplates exhibit a strong waveguiding effect. e | Photoluminescence (PL) from monolayers of (BA)2(MA)Pb2I7
encapsulated by boron nitride. The structural reorganization caused by laser annealing of the (BA)2(MA)Pb2I7 monolayers
induces a shift in the excitonic energy. The black line is the PL of the ordered stage, and the blue line is the PL of the
disordered stage after surface relaxation. f | Illustration showing the order–disorder transition induced in monolayers of
(BA)2(MA)Pb2I7 by laser illumination. g | Transmission electron microscopy image of CsPbBr3 nanoplates with lateral
dimensions of 20 nm and thicknesses of a few unit cells (~3 nm) that emit cyan-​coloured light upon excitation (inset).
h | Optical absorption (dashed lines) and emission (solid lines) spectra of colloidal nanoplates of CsPbBr3 with controllable
thicknesses from 2 to 6 unit cells (UCs) and a CsPbBr3 nanocube. i | AFM topography images of ultrathin MAPbI3 nanosheets
that are a few nanometres thick and were prepared by vapour-​phase conversion of few-​layer-thick PbI2 nanosheets. CQD,
colloidal quantum dot; OD, optical density. Panel a is adapted with permission from ref.82, ACS. Panel b is adapted from
ref.86, Springer Nature Limited. Panel c is adapted with permission from ref.90, AAAS. Panel d (first two images on left-​hand
side) is adapted from ref.91, Springer Nature Limited. Panels e and f are adapted from ref.89, Springer Nature Limited.
Panel g is adapted with permission from ref.98, ACS. Panel h is adapted from ref.99, CC-​BY-4.0. Panel i is adapted with
permission from ref.102, ACS.

www.nature.com/natrevmats
Reviews

interplay of thermodynamics and kinetics between RP facilitate studies of their properties and their integration
phases with similar n values. Crystals of RP perovskites into optoelectronic devices. Therefore, new methods
of (BA)2(MA)n−1PbnI3n+1 with n ≤ 5 can be obtained by for the controlled growth of large-​area nanosheets of
using an off-​stoichiometry protocol83,84. This approach 2D perovskites that could rival the lateral dimensions
has also been used to obtain a series of Dion–Jacobson of more traditional 2D materials are needed.
2D perovskites in which the bilayer of monovalent RP perovskites are already inherently 2D quantum
long-​chain alkylammoniums (LA+) was replaced by a confined, and therefore, a decrease in the nanoplate
monolayer of dications85. thickness is not expected to produce further confine­
Owing to the strong quantum and dielectric confine­ ment effects. However, owing to the soft lattice of RP
ment (Fig. 3a), stable excitons with large binding ener­ perovskites, lattice deformation under external pertur­
gies of hundreds of meV can form in RP perovskites. bations (for example, pressure, strain or temperature)
The large exction binding energies give rise to intense can lead to varied optical and electronic properties,
PL with narrow peak widths and high radiative recom­ especially when the thickness is at the single-​layer
bination rates at room temperature, making RP per­ level. For example, a slight blue shift of the emission
ovskites promising light emitters. One notable example peak of (BA)2PbBr4 was observed upon decreasing
is (PEA) 2PbBr 4 (PEA +  = phenylethylammonium), the number of layers90, which was attributed to lattice
which exhibits a PLQY of up to 79% and a linewidth of expansion. A reversible shift in excitonic energies can
20 nm in the deep-​blue region86. These figures of merit be induced by laser annealing of exfoliated sheets of
are superior to those of wide-​bandgap semiconductors (BA)2(MA)n−1PbnI3n+1 encapsulated by boron nitride89
(for example, GaN), organic emitters and 3D chloride-​based (Fig. 3e). This phenomenon was attributed to the revers­
perovskites with similar emission wavelengths (Fig. 3b). ible structural reorientation of the surface BA+ cations in
The optical bandgap of RP perovskites is mainly dictated the easily deformable lattice (Fig. 3f), which formed free
by n, as the band structure is primarily determined by and bound excitonic states.
the inorganic layer. Increasing the quantum well thick­ Quantum-​confined nanoplates of 3D perovskites
ness (n) from 1 to 5 progressively decreases the corre­ can be obtained by modifying the colloidal synthesis
sponding bandgap of (BA)2(MA)n−1PbnI3n+1 from 2.43 eV protocol for perovskite nanocrystals. Analogous to
to 1.83 eV. Although the LA+ cations do not contribute RP perovskites, colloidal nanoplates can be described
directly to perovskite band structures, they can sub­ by the formula (LA)2(A)n−1BX3n+1, where LA+ is a sur­
tly influence PL properties by distorting the inorganic face ammonium ligand. These 2D nanostructures can
lattice29 and through electron–phonon interactions. be considered as free-​standing monolayers or multi­
Electron–photon interactions are stronger in more rigid ple layers of the RP perovskites with a specific n value.
2D perovskites, which results in faster non-​radiative Decreasing the thickness down to the length of the
recombination and a lower PLQY86. exciton Bohr radius leads to quantum-​confinement
effects. Quantum-​confined nanoplates were initially dis­
2D nanostructures. RP perovskites with an intrinsic 2D covered as a side product in the synthesis of MAPbBr3
crystal structure are a new class of solution-​processable nanocrystals92. The absorption and emission spectra
2D materials with excellent optical properties rarely of the nanoplates showed multiple blue-​shifted peaks
observed in existing classes of 2D materials, such as associated with different values of n. Since these initial
strong PL emission spanning from the UV to the near-​IR studies, significant progress has been made in con­
spectral regions. Starting from phase-​pure RP perovskite trolling the nanoplate dimensions and increasing the
single crystals, mechanical exfoliation is a convenient PLQY, and a wide range of perovskite compounds has
method for accessing thin nanosheets with precise been demonstrated93–98. For example, colloidal nano­
n values for fundamental studies87–89. However, the plates of MAPbBr3 with varying thicknesses have been
intrinsic softness of the perovskite lattice makes it diffi­ synthesized by tuning the ratio of LA+ to MA+ (ref.93).
cult to maintain structural integrity during exfoliation. According to the proposed mechanism for colloidal
Direct crystallization is a more robust method for obtain­ nanoplate growth, LA+ binds to the surfaces, which
ing high-​quality nanosheets with uniform thickness and confines crystal growth in one dimension after nucle­
a smooth surface. Ultrathin nanosheets of (BA)2PbBr4 ation. All-​inorganic perovskite nanoplates have also
with thicknesses of a few layers and lateral dimensions been demonstrated. Modifying the protocol reported for
of several micrometres (Fig. 3c) were grown in a solvent-​ the growth of cubic CsPbBr3 nanocrystals by lowering the
evaporation process90. A ternary solvent system was used reaction temperature led to the formation of CsPbBr3
to tune the solubility and crystallization rate, which is nanoplates with a thickness of a few unit cells98 (Fig. 3g).
crucial to the crystal growth. The thinnest nanosheet However, initial syntheses often produced nanoplates
obtained had a thickness of 1.6 nm, which corresponds with different thicknesses, leading to broad PL emis­
to one monolayer. In another work, solution-​phase trans­ sion. Only recently has the synthesis of monodispersed
port growth was used to form nano­plates of (PEA)2PbX4 nanoplates of a single thickness been achieved99. The
(X− = Br−, I−) with a well-​defined rectangular geometry, PL emission of the CsPbBr3 nanoplates shifts from
lateral dimensions of up to several tens of micrometres greenish-​blue (497 nm) to deep blue (432 nm) when the
and nanoscale thicknesses of several hundred nano­ thickness is decreased from 6 to 2 layers (Fig. 3h). High
metres91. The emission colour can be readily tuned by densities of surface defects owing to under-​coordinated
halide alloying or varying the n value (Fig. 3d). Increasing lead atoms result in low PLQYs. However, the PLQY can
the lateral dimensions of these nanosheets would be increased to nearly unity by repairing these defects100.

Nature Reviews | Materials


Reviews

In addition to controlling the thickness, introducing based on these films have achieved power conversion
shorter ligands together with the longer ligands and var­ efficiencies (PCEs) of > 12% while exhibiting greatly
ying their ratios enable the growth of nanosheets with increased device stability compared with the 3D
larger lateral dimensions of up to a few micrometres101. MAPbI3-based devices when subjected to thermal, light
Generally, the shape and composition of colloidal nan­ and humidity tests113,114.
oplates evolve over time82, and thus, stabilizing them Another form of structural instability is exhibited
remains challenging. Moreover, the high PLQY and by metastable perovskites such as CsPbI3 and FAPbI3
strong confinement effects are often lost when they are (FA+  = formamidinium; Box 1), which undergo spon­
made into solid-​state films. taneous phase transitions into non-​perovskite poly­
It is nontrivial to prepare ultrathin nanosheets of 3D morphs at room temperature. These non-​perovskite
perovskites owing to the inherent cubic crystal structure. polymorphs are undesirable because their optoelectronic
The presence of LA+ in the growth solution is crucial to properties are inferior to those of the perovskite phases.
break the crystal symmetry of 3D perovskites into 2D RP Nanostructuring is an effective approach to gain access
phases and therefore to promote anisotropic 2D growth. to these metastable perovskite polymorphs because they
Nevertheless, several alternative approaches have been have a lower surface energy than the corresponding
developed that do not require LA+. For example, start­ stable non-​perovskite phases115. When the increasing
ing from PbI2 nanosheets with a thickness of a few lay­ energy contribution from the surface outweighs that
ers as templates, vapour-​phase conversion yielded 2D from the bulk as the size of crystallites decreases, the
quantum-​confined nanosheets of MAPbI3 with con­ metastable phase can potentially be thermodynamically
trollable thicknesses of a few nanometres102 (Fig. 3i). stabilized (Fig. 4c). In the initial report on the colloi­
Ultrathin 2D nanosheets of MAPbCl3 with thicknesses dal synthesis of CsPbI3 nanocrystals, the nanocrystals
of 8.7 nm and lateral dimensions of >20 µm have also formed in the perovskite structure but transformed
been directly synthesized by chemical vapour deposi­ into the non-​perovskite phase within a few days owing
tion103. These vapour-​phase methods have also been to aggregation116. A modified purification approach in
used to create vertical heterostructures of perovskites which the nanocrystals are isolated without full removal
with other 2D materials104,105. of the surface ligands can increase the structural stabil­
ity to several months117 (Fig. 4d). Solar cells containing
Phase stabilization of nanostructures thin films made of stabilized CsPbI3 nanocrystals have
Poor structural stability and chemical stability remain demonstrated PCEs of up to 10.8%. Theoretical calcu­
major concerns for the practical application of halide lations have confirmed that the perovskite structure
perovskites. The susceptibility of 3D perovskites to of CsPbI3 has a lower surface energy than the non-​
chemical and structural degradation stems from the perovskite structure, and nanocrystals of the perovskite
low formation energy and reconfigurable lattice of phase are thermodynamically more stable when their
these materials. The high surface-​area-to-​volume ratio size is <100 nm (ref.118). Similarly, colloidal nanocrys­
of nanostructures can increase the impact of surface tals of FAPbI3 that exhibited bright luminescence were
properties on the chemical properties and phase stabil­ stabilized in the cubic perovskite phase119,120.
ity. However, counterintuitively, nanostructures of halide In addition to the above size effect, surface ligand
perovskites can exhibit enhanced structural and chemi­ functionalization can also contribute to phase stabi­
cal stability owing to a surface energy effect and surface lization by reducing the surface energy of the nano­
ligand functionalization. structures. Through surface functionalization, it is
3D perovskites are particularly susceptible to possible to stabilize pure FAPbI3 in the perovskite phase
moisture-​induced degradation owing to the strong in single-​crystal nanostructures with up to micrometre
interaction between water and the perovskite surface. sizes121 (Fig.  4e). Without surface functionalization,
For example, MAPbI3 complexes with water to form FAPbI3 exclusively forms in the non-​perovskite phase
hydrated phases upon exposure to high humidity lev­ at room temperature. The surfaces of the stabilized
els106. RP perovskites107, colloidal nanoplates108 and nanostructures are terminated by LAI, which is similar
functionalized 3D perovskites109 are more resistant to to RP perovskites. The LA+ cations occupy the surface
humidity than prototypical 3D perovskites because the A sites, and the I− anions coordinate to the surface Pb2+
surface is terminated by long-​chain ammonium cations ions such that all of the [PbI6]4− octahedra are complete.
with the hydrophobic alkyl chains oriented outwards This charge-​balanced surface structure is an intuitive
(Fig. 4a). These surface ligands render the perovskite surface termination for halide perovskites, especially
surface more hydrophobic109 (Fig. 4b) and prevent mois­ in ammonium-​halide-rich environments. Calculations
ture from penetrating into the perovskite lattice. Recent show that surface functionalization can lower the sur­
calculations have confirmed that RP perovskites exhibit face energy of nanostructures (Fig. 4f), which causes the
high positive water adsorption energies, whereas 3D per­ crossover between the metastable perovskite phase and
ovskites have negative (that is, favourable) adsorption non-​perovskite phase to occur at larger crystal sizes
energies110. However, it should be noted that solution-​ (Fig. 4c). The smaller surface energy of the functionalized
processed thin films of RP perovskites (n > 1) consist nanostructures can be attributed to van der Waals inter­
of a mixture of RP perovskites with different n values actions between adjacent ligands on the surfaces. The
along with nanoscale crystallites of 3D perovskite111,112. functionalized FAPbI3 nanostructures did not exhibit
As n increases, the system can be considered as a 3D detectable phase conversion to the non-​p erovskite
perovskite with surface functionalization. Solar cells phase upon storage for several months. In addition to

www.nature.com/natrevmats
Reviews

a FA+-terminated surface PEA+-terminated surface e

2 μm

Surface functionalization
b

500 nm

10μm

c d f

Formation energy (eV)


–5
Metastable
Free energy

perovskite –6

Surface –7
(FA)2(FA)n–1PbnI3n+1
functionalization –8
6.2 Å
Non-perovskite 5 nm –9 (PEA)2(FA)n–1PbnI3n+1

Surface area or inverse crystal size –10


0 1 2 3 4 5 6 7 8 9
n

Fig. 4 | Chemical and structural stability in perovskite nanostructures. a | Illustrations of the (001) surface of cubic
FAPbI3 (FA+  = formamidinium) terminated by FA+ (left, a prototypical surface) and PEA+ (right, a functionalized surface;
PEA+  = phenylethylammonium). b | Photographs of a water droplet on a pristine MAPbI3 (MA+  = methylammonium) film (left)
and a MAPbI3 film functionalized with hexadecyltrimethylammonium on the surface (right), showing different contact
angles and wetting properties. c | Plot of the total free energy as a function of surface area (or inverse crystal size) for a non-​
perovskite, metastable perovskite and surface-​functionalized metastable perovskite. d | High-​resolution transmission electron
microscopy image of CsPbI3 nanocrystals stabilized in the cubic perovskite phase. e | Scanning electron microscopy images
of hexagonal FAPbI3 nanostructures grown without long-​chain alkylammonium (L A+) ligands (top) and cubic perovskite FAPbI3
nanostructures grown with L A+ ligands (bottom) at room temperature. The insets show the corresponding optical images
of the substrates. f | Calculated formation energies of (FA)2(FA)n−1PbnI3n+1 and (PEA)2(FA)n−1PbnI3n+1 as a function of the thickness of
the inorganic perovskite layer, n. The PEA+-terminated FA+-based perovskites always have a more negative formation energy
than their FA+-terminated counterparts, indicating that the former have a smaller surface energy. Panels a, e and f are
adapted with permission from ref.121, ACS, further permissions related to the material should be directed to the ACS.
Panel b is adapted from ref.109, Springer Nature Limited. Panel d is courtesy of Joseph M. Luther, University of Colorado, USA.

Optoelectronic applications
the thermodynamic stabilization effect, the hydropho­ The success of 3D perovskites in photovoltaics has moti­
bic surface of the functionalized FAPbI3 nanostructures vated the development of various high-​performance
could also mitigate moisture-​induced phase transforma­ optoelectronic devices129. Single-​crystal 1D and 2D
tion. The phase stability of CsPbI3 perovskite nanocrys­ nanostructures (or microstructures) of perovskites
tals122 and thin films123,124 exhibits strong dependence on offer potentially superior optoelectronic performance
the capping ligands, further indicating the importance and many exciting opportunities in comparison with
of surface chemistry on phase stabilization. Even more nanostructures of conventional semiconductors or the
robust phase stabilization can be achieved through bulk forms of 3D perovskites.
stronger surface ligand coordination with zwitterionic
molecules or polymers125,126. Perovskite nanowires for photonics and optoelec-
Preventing thermal-​induced and photoinduced tronics. Single-​crystal semiconductor NWs can serve
degradations are key challenges for the application of as optical gain media and provide the cavity for lasers
hybrid perovskite nanostructures. Upon above-​bandgap that could potentially outperform their thin-​film coun­
excitation or heating, hybrid perovskites decompose into terparts4. Since the pioneering work on UV lasing
the parent lead halide, gaseous organic amine and HX. from ZnO NWs130, optically pumped lasing has been
One feasible approach to suppressing the degradation demonstrated in various semiconductor NWs in the
process is to prevent the organic species from evapo­ UV to near-​IR spectral regions4. These NW lasers hold
ration. For example, encapsulating MAPbI3 microplates great promise as miniature coherent light sources for
or (PEA)2PbI4 nanosheets with hexagonal boron nitride a wide range of applications. However, the growth of
thin flakes and/or polycarbonates can greatly improve high-​quality NWs capable of lasing generally requires
their stability127,128. high temperatures and a vacuum, and the difficulty of

Nature Reviews | Materials


Reviews

realizing continuous-​wave operation131–133 and electri­ hybrid perovskite counterparts, showing stable lasing
cally injected lasing hinders potential application of emission with no measurable degradation after at least
NW lasers. 7.2 × 109 laser shots and 8 hours (Fig. 5d). The improved
Perovskite NWs have demonstrated outstanding stability of CsPbBr3 NWs under illumination is consist­
lasing performance, facilitating NW lasers capable of ent with their enhanced photostability and thermal sta­
continuous-​wave operation. The first perovskite NW bility compared with those of MAPbX3. Initial perovskite
laser was demonstrated in NWs made of MAPbI3 grown NW lasers were demonstrated with femtosecond pulsed
by solution-​phase recrystallization10. These single-​crystal excitation to avoid thermal degradation. Recently, under
NWs have smooth facets and subwavelength dimen­ continuous-​wave pumping and at low temperatures,
sions, which make them efficient nanoscale Fabry− lasing emission from CsPbBr3 NWs was observed at an
Pérot cavities. Under pulsed optical excitation, these excitation density of 6 kW cm−2 and with a quality factor
MAPbI3 NWs lase at room temperature with low las­ of 2,300 (ref.139). The achievement of continuous-​wave
ing thresholds of ~220 nJ cm−2, record-​breaking qua­ lasing is a step towards the development of electrically
lity factors of ~3,600 (Fig. 5a) and broad stoichiometric driven lasers.
wavelength tunability. The lasing threshold values for In addition to lasing applications, perovskite NWs
single-​crystal MAPbI3 NWs are substantially lower than can also function as efficient nonlinear emission sources,
those reported for lasing from polycrystalline MAPbI3 electro-​optic modulators and photodetectors. Owing
thin films and nanostructures9,134 and a family of near-​ to the large multiple-​photon absorption cross section
IR III–V semiconductor NW lasers135,136. The superior and low threshold required for population inversion,
lasing performance of perovskite NWs is attributed to multiple-​photon pumped lasing from perovskite NWs
exceptionally low rates of non-​radiative recombination has been achieved 140,141. Single-​c rystal perovskite
in these single-​crystal nanostructures. Moreover, the NWs can also function as active waveguides142, exhib­
MAPbI3 NWs showed extremely high PLQYs of up iting an ultralow propagation loss of 0.04 dB µm−1. The
to 87% without any intentional surface passivation. combination of the perovskite NW waveguides with
By contrast, to circumvent the strong non-​radiative sur­ electrical inputs has been explored for potential appli­
face recombination in GaAs NWs, surface passivation cation as electro-​optical modulators (Fig. 5e). The fabri­
by core–shell design or a cryogenic temperature is often cated NW electro-​optical device achieved a PL intensity
required135. Nevertheless, the state-​of-the-​art epitaxial modulation of up to 98.4% (Fig. 5f), outperforming most
GaAs–AlGaAs core–shell NW laser has a low PLQY inorganic and organic NWs. The effective modula­
of ~0.4%136. This suggests that the electronic properties of tion may result from a combined effect of the unique
MAPbI3 in single-​crystal NWs are much more tolerant ion migration and photon recycling in perovskites.
to surface defects than those of GaAs even though their Photodetectors fabricated with MAPbI3 NWs exhibit
bulk forms exhibit similar photophysical properties. The responsivities of up to 104 A W−1, detectivities approach­
lasing performance from polycrystalline MAPbX3 NWs ing 1012 Jones, a linear dynamic range of >150 dB and a
synthesized by vapour-​phase conversion137 was poorer fast photoresponse with a bandwidth of up to 0.8 MHz
than that of NWs fabricated by direct solution synthesis, (refs69,143). The high performance mainly originates from
probably owing to the inferior crystalline quality. The the low doping concentration (nearly intrinsic) and low
effects of morphology on lasing from perovskite nano­ density of trap states in high-​quality perovskite NWs,
structures have been established, for which we refer which result in very low dark current and high gain.
readers to a recent review138. Furthermore, perovskite
NWs have demonstrated continuous tunability of the Light-​emitting diodes. 3D perovskites are well known
lasing wavelength from 420 nm to 820 nm, which can for their high photovoltaic performance, which is
be readily achieved by controlling the anionic or catio­ enabled by the long carrier diffusion lengths and low
nic stoichiometry of APbX3 (Fig. 5b,c). The emission of exciton binding energies. However, the light-​emitting
NW lasers mostly comes from the two end facets with efficiency of perovskites is reduced owing to the com­
coherent interference under lasing conditions (Fig. 5c). petition between non-​radiative trapping processes and
Interestingly, the wide-​bandgap all-​inorganic perov­ slow bimolecular radiative recombination under low
skites CsPbCl3 and CsPb(Br,Cl)3 demonstrate efficient injected carrier densities (that is, <1015 cm−3), which is
lasing at room temperature in the violet–blue spectral the regime that typical LEDs operate in144. High light-​
region35,36, which is not possible for wide-​bandgap hybrid emitting efficiency can be obtained at high fluences, at
perovskites such as MAPbCl3. We suspect that wide-​ which radiative bimolecular recombination dominates
bandgap CsPbCl3, in which the ionic radii are favoura­ the carrier dynamics. Because the transition from bulk
bly matched in the perovskite structure, could be more 3D perovskites to nanostructures or 2D RP perovskites
stable and less defect-​prone than MAPbCl3 and FAPbCl3, with reduced dimensionalities introduces confine­
thus exhibiting superior light-​emitting properties. ment effects and increases the probability of radiative
Although MAPbX3 NWs exhibit excellent lasing recombination (Fig. 6a), nanostructuring is an effective
performance, they are unstable under prolonged illumi­ strategy to ensure high PL at low carrier densities by sup­
nation. In comparison, single-​crystal NWs of FAPbX3 pressing exciton dissociation and promoting radiative
and CsPbX3 grown by either solution-​phase or vapour-​ recombination145.
phase methods show similar lasing performance but Benefiting from the fundamental understanding of
exhibit increased stability34–36. In particular, all-​inorganic photophysics and synthetic advances, LEDs based on
CsPbBr3 NWs were substantially more robust than their solution-​processed 2D RP perovskites have achieved

www.nature.com/natrevmats
Reviews

excellent performance with high external quantum room-​temperature electroluminescence (EL)147. By con­
efficiencies (EQEs), high brightness, high colour purity verting polycrystalline thin films of (PEA)2PbBr4 into
and broad spectral tunability. The earliest LEDs using single-​crystal micrometre-​sized plates through a sol­
n = 1 perovskites were reported in the 1990s and could vent vapour-​annealing technique, the PLQY increased
function only at cryogenic temperatures146. Improving from 10% to 26%. Furthermore, colour-​pure violet
the crystal quality increased the PLQY and enabled LEDs were demonstrated at room temperature with the

a c

Intensity (arb. units)


40 554 nJ cm–2
3 580nJ cm–2

FWHM (nm)
604nJ cm–2
20 617 nJ cm–2
Intensity (arb. units)

630 nJ cm–2
d
2 0 1.0
400 500 600 700
Pump fluence δλ ≈ 0.22 nm

Lasing intensity
FAPbI3
(nJ cm–2)

(arb. units)
MAPbI3
1 MABr-stabilized
FAPbI3
CsPbBr3

0.1
782 783 784 785 786 787 788 789 0.01 0.1 1.0 10
Wavelength (nm) Illumination time (h)
b
x=3 CsPbClxBr3–x x=0, y=3 CsPbBryI3–y y =3 (MA,FA)Pb(Br,I)3
PL intensity (arb. units)

400 450 500 550 600 650 700 750 800 850
Wavelength (nm)
e f
Laser
12,000

PL intensity (arb. units)


0V 0

Perovs 12,000
k ite
15 V
0
V 12,000
Detector
30 V
0
Electric field

Fig. 5 | High-​performance nanoscale optoelectronic and photonic applications of perovskite NWs. a | Emission spectra
around the lasing threshold of a MAPbI3 (MA+  = methylammonium) nanowire (NW) pumped by a 402-nm-​pulsed laser
excitation (150 fs, 250 kHz). The inset shows the integrated emission intensity (grey circles) and full width at half maximum
(FWHM, blue triangles) as a function of pump fluence, indicating the lasing threshold at ~600 nJ cm−2. The FWHM (δλ) of the
lasing peak at 630 nJ cm2 is ~0.22 nm, corresponding to a quality factor of ~3,600. b | Broad wavelength-​tuneable lasing from
single-​crystal NWs of APbX3 (A+ is Cs+, MA+ or formamidinium (FA+); X− is a halide) perovskites with different compositions.
c | Optical images of CsPbX3 (X− = Cl−, Br− or I−) NW lasers with tuneable colour emission. The images show that emission
mostly comes from the two end facets with coherent interference under lasing operation. d | Integrated emission intensity
of representative MAPbI3, FAPbI3, MABr-​stabilized FAPbI3 and CsPbBr3 NW lasers as a function of illumination time under
continuous illumination of a 402-nm-​pulsed laser excitation at room temperature. e | Schematic of an electro-​optical
modulator device fabricated with a perovskite NW. f | Optical dark-​field images of the electro-​optical modulator device
under different external voltages (V), showing the modulation of the photoluminescence (PL) intensity under an applied
voltage. Panel a is adapted from ref.10, Springer Nature Limited. Panel b is adapted with permission from ref.10, Springer
Nature Limited, and refs34,35, ACS, further permissions related to the material should be directed to the ACS. Panel c is
adapted with permission from ref.35, ACS. Panel d is adapted with permission from refs34,35, ACS, further permissions
related to the material should be directed to the ACS. Panel f is adapted with permission from ref.142, Wiley-​VCH.

Nature Reviews | Materials


Reviews

a b electroluminescence peak centred at 410 nm and with


–2
3D perovskite 2D RP perovskite a narrow full width at half maximum of 14 nm owing
Small n Large n
–3 –– to strong quantum confinement. With perovskites of
– – higher n values (n > 1), multiple-​phase 2D RP per­

Energy (eV)

MoOx/Au
hν – hν –– ovskites with different values of n and 3D perovskite

TFB
–4

ITO
ZnO/PEIE
+
+ hν –
crystallites can be assembled in solution-​processed thin
–5
+ + films to form an energy-​cascade structure148–150 (Fig. 6b).
Perovskite MQWs ++
–6 ++
Spectroscopic studies revealed that these films function
Recombination – Electron
as an energy funnel that can guide the charge carriers
at defect + Hole to the component with the smallest bandgap (Fig. 6c,d)
and enable efficient radiative recombination even at low
c d
0 ns 567 nm excitation densities150,151 (Fig. 6e). The ultrafast energy
0.003 0.5 ns 610 nm transfer from wide-​b andgap to narrower-​b andgap
1 ns 642 nm
5 ns 0.01 695 nm
components occurs on the picosecond timescale. The
0.002 PL (static) energy-​cascade design has enabled high-​performance
LEDs with EQEs of up to 11.7% in the near-​IR region
∆T

∆T

0.001 (~760 nm). Owing to the presence of surface ligands


that stabilize the crystallites and impede ion migration
0.001 under an electric field, these devices based on 2D perov­
0.000
skites also showed greater operational stability than
3D-​perovskite-based devices.
500 600 700 800 0.1 1 10 100 1,000 More efficient LEDs with tuneable emission
Wavelength (nm) Delay time (ps) can be achieved by changing the composition of
e f (LA)2(A)n−1BnX3n+1 and by controlling the value of n
3D FAPbI3 film 1.0
2D (NMA)2(FA)Pb2I7 film (refs151–153). For example, the investigation of a series
of (PEA)2(FA)n−1PbnBr3n+1 films with various n values
60 0.8
showed that the n = 3 system exhibited the highest PLQY.
EL intensity (norm.)

Efficient green LEDs with the n = 3 perovskite achieved


0.6
40 a current efficiency of 62.4 cd A−1 and a high EQE of
PLQY (%)

14.36%152. Moreover, the EL can be tuned across a wide


0.4
spectral range using mixed halide compositions of dif­
20
0.2
ferent ratios149. Note that all of the LEDs fabricated with
stoichiometric (LA)2(A)n−1BnX3n+1 (n > 1) compositions
0.0
exhibit an EL peak close to that of the corresponding
0
10–1 100 101 102
400 600 800 1,000 3D perovskites. This is due to the inevitable formation
Lasing intensity (mW cm–2) Wavelength (nm) of 3D perovskite crystallites when using stoichiometric
synthesis protocols and the rapid energy transfer from
Fig. 6 | High-​performance leDs based on self-​assembled 2D perovskite nanostructures.
a | Comparison of charge-​carrier recombination in 3D perovskites (left) and 2D Ruddlesden– wide-​bandgap components to 3D perovskites. To fully
Popper (RP) perovskite films (right), which are assemblies of multiple quantum wells take advantage of the strong quantum confinement of
(MQWs) with various values of n (where n indicates the thickness of the inorganic perovskite 2D perovskites, excess LAX is needed to suppress the
layers). In 3D perovskite films, defect-​state trapping competes effectively with bimolecular growth of 3D perovskites and to produce small-​n RP
radiative recombination under low excitation densities, resulting in a relatively low perovskites154,155.
photoluminescence quantum yield (PLQY). In RP perovskite films, charge carriers are rapidly Although high-​performance green and red LEDs
localized from wide-​bandgap (small-​n) RP phases to the narrowest-​bandgap phase (large-​n based on perovskites have been achieved, it remains
RP phases or nanoscale crystallites of 3D perovskites), and thus the radiative recombination challenging to obtain the blue LEDs that are required
rate is faster than in bulk 3D perovskite films. b | Schematic representation of the flat-​band for lighting and displays. Mixed halide 3D perovskites
energy-​level diagram and cascade energy transfer in self-​assembled MQWs of a RP
(Br− and Cl−) can exhibit blue emission; however, the
perovskite film. The charger carriers are transferred downstream from the wider-​bandgap
RP phases to narrower-​bandgap phases, and the emission is mainly from the component unavoidable phase separation that occurs under electri­
with the narrowest bandgap. c | Transient absorption spectra of a (PEA)2(MA)2Pb3I10 (MA+ is cal bias and the formation of trap states associated with
methylammonium; PEA+ is phenylethylammonium) perovskite film at different timescales Cl− are limitations for achieving stable emission and high
show that the relative intensity of the four bleaching peaks evolves over time. d | Bleaching PLQY156. RP perovskites and colloidal nanoplates that
kinetics in the (PEA)2(MA)2Pb3I10 film in panel c probed at selected wavelengths, showing a exhibit strong quantum confinement are therefore prom­
cascade on the 100 fs to 100 ps timescale from higher-​energy phases to lower-​energy ising candidates for fabricating blue LEDs. By tuning the
phases. e | Excitation-​intensity-dependent PLQY of a 3D perovskite FAPbI3 film and a RP ratio of LABr:ABr:PbBr2 in the precursor solution, thin
perovskite film of (NMA)2(FA)Pb2I7 (FA+ is formamidinium; NMA+ is 1-(2-naphthylmethyl) films mainly composed of quantum-​confined perovskite
ammonium). f | Electroluminescence (EL) spectrum of a light-​emitting diode (LED) device grains can be obtained. These thin films enabled the fab­
fabricated with CsPbBr3 nanoplates. The inset shows a photograph of the LED device with
rication of tuneable blue–green LEDs with emissions in
blue-​light emission. ∆T, change in transmission; ν, photon energy ; h, Planck’s constant; ITO,
indium tin oxide; PEIE, polyethylenimine ethoxylated; TFB, poly(9,9-dioctyl-​fluorene-co-​ the range 460–520 nm (ref.154). The LED device fabricated
N-(4-butylphenyl)diphenylamine). Panel a is adapted from ref.144, Springer Nature Limited. with colloidal nanoplates of CsPbBr3 (with a thickness of
Panel b is adapted from ref.153, CC-​BY-4.0. Panels c and d are adapted from ref.148, Springer three unit cells) shows pure blue emission with a narrow
Nature Limited. Panel e is adapted from ref.149, Springer Nature Limited. Panel f is adapted peak centred at 460 nm (refs99,100) (Fig. 6f). However, the
from ref.99, CC-​BY-4.0. EQEs of the devices remain quite low (~0.1–1%).

www.nature.com/natrevmats
Reviews

The study of physical properties removing the uncertainty due to ensemble averaging
Single-​crystal nanostructures offer a platform for inves­ of inhomogeneous polycrystalline samples (Fig.  7b).
tigating the intriguing physical properties of diverse Compared with bulk single crystals (Fig. 7c), optically
perovskite materials. In polycrystalline samples, the thin and flat nanoplates may be more suitable for many
presence of grain boundaries along with the sensitiv­ spectroscopic studies, such as PL measurements, because
ity to preparation conditions may obscure the intrinsic reabsorption of the PL emission can be minimized.
properties of the perovskites157,158. By contrast, each indi­ Moreover, perovskite nanoplates have a submicrometre
vidual perovskite nanostructure is a single crystal with thickness that is similar to the optimal thickness of the
well-​controlled facets and no grain boundaries (Fig. 7a), perovskite layer in thin-​film solar cells. Additionally,

a b d
1D diffusion model
Laser
Emission
APbX3 Delay time
2 μm
2 μm
+ Carrier diffusion +
– –
c

Solid-state ion interdiffusion


10 mm – Electron
+ Hole
10 μm A′PbX′3

e f g h
1.0
k2 x 10–10 cm3 s–1 Excitation
10 3 8 30 1.0
Intensity (arb. units)

h
Delay time (ns)

k1 x 107 s–1

Intensity (arb. units)


25 0.8

Distance (μm)
6 5.3μm
0.5 20
k1 or k2

10 2
7.1 μm 0.6
4 8.8 μm 15
0.4
10.7μm 10
1
10 2 0.2
0.0 5 5 µm
5 μm 0.0
0
2.3 2.4 2.5 2.6 MA+ FA+ Cs+ MA+ FA+ Cs+ 460 480 500 520 540 560
Energy (eV) Cation Wavelength (nm)

i – Electron j UP k
FP cavity + Hole LP
Cavity photon
Exciton 2.32
Energy (eV)

Perovskites 2ħΩ
Energy

+
– Exciton 2.30

Cavity photon

Momentum Intensity 2.6 2.7 2.8 2.9


(arb. units) kz (10–3 Å–1)

Fig. 7 | studies of physical properties using single-​crystal perovskite diffusion in a MAPbI3 NW. The inset is an optical image of the NW; the red spot
nanostructures. a | Optical image of representative single-​crystal perovskite is the location of laser excitation. h | Normalized PL line mapping of a gradient
nanostructures with well-​controlled facets and subwavelength thicknesses CsPbBr3−xClx NW. The y axis indicates the distance from the edge of the
(as highlighted by the scanning electron microscopy image of a single plate CsPbCl3 microplate that is placed over the CsPbBr3 NW. The inset is a real-​
shown in the inset). b | Scanning electron microscopy image of a colour PL image of the NW under broad laser excitation. i | Schematic of an
polycrystalline thin film. The optically thin nanoplates are typically tens of exciton–polariton microcavity. j | Schematic showing the dispersion of a
micrometres in lateral dimensions and hundreds of nanometres in thickness, cavity photon, exciton, upper polariton (UP) and lower polariton (LP)
which make them a suitable model system for spectroscopic studies. branches. The black points indicate Fabry–Pérot modes in perovskite NWs;
c | Photograph of a bulk single-​crystal perovskite. d | Illustration of the these modes are shown on the energy−wavevector diagrams in panels j and k.
charge-​carrier diffusion process in a perovskite (APbX3, where A+ is Cs+, k | Emission spectrum of a CsPbBr3 NW (left) and the corresponding energy–
methyl­ammonium (MA+) or formamidinium (FA+) and X− is a halide) nanowire wavevector (kz) dispersion curve along with the exciton–polariton model fit
(NW) and the ion interdiffusion process in a perovskite heterostructure. (red solid line). ħ, reduced Plank’s constant; Ω, Rabi splitting energy ; FP, Fabry–
e | Time-​dependent photoluminescence (PL) spectra from a MAPbBr3 nanoplate, Pérot. Panels a (inset) and e are adapted with permission from ref.159, AAAS.
showing hot PL emission with a lifetime of ~102 ps. f | Comparison of the first-​ Panel b is adapted from ref.157, Springer Nature Limited. Panel c is adapted
order rate constant (k1) or carrier trapping rate constant and the second-​order with permission from ref.167, AAAS. Panel g is adapted with permission from
rate constant (k2) for electron–hole radiative recombination for MAPbBr3, ref.25, ACS. Panel h is adapted with permission from ref.77, ACS. Panel k is
FAPbBr3 and CsPbBr3 nanoplates163. g | The PL intensity image reveals carrier adapted with permission from ref.139, Wiley-​VCH.

Nature Reviews | Materials


Reviews

the 1D geometry of NWs provides directional channels MAPbBr3. Similarly, a minority charge-​carrier diffu­
for charge and/or photon transport and ion migration sion length of up to 21 μm in MAPbI3 NWs under an
(Fig. 7d), which could enable much simpler modelling of electric field has been directly measured with scanning
their kinetics. However, nanostructures have a greater photocurrent microscopy26. The corresponding carrier
surface area than do bulk single crystals. Although the mobility was estimated to be 170 cm2 V−1 s−1 using a PL
greater surface area could be a drawback, the surface lifetime of 1 μs. The diffusion lengths and carrier mobil­
defects are more benign and/or easily passivated in ities for single-​crystal nanostructures are several to tens
nanostructures, and the effect on properties might be of times higher than the values found for polycrystal­
minimized. Therefore, nanostructures with well-​defined line thin films164,165 and are comparable to those of bulk
surface facets can be used as a more sensitive (rela­ single crystals166,167 (that is, on the order of 10 µm and
tive to single crystals) but more controlled (relative to 10–100 cm2 V−1 s−1, respectively), which suggests that
polycrystalline films) material platform to study surface the absence of grain boundaries substantially increases the
effects. carrier diffusion length. In another example of optical
studies on single-​crystal nanostructures, the tetragonal-​
Charge-​carrier properties. To study the charge-​ to-orthorhombic phase transition in a single MAPbI3
carrier properties of 3D perovskites, time-​resolved PL NW was studied using super-​resolution PL imaging168
measurements were conducted on nanoplates of 3D and indicated that the phase transition temperature may
perovskites with different cations (MA+, FA+ or Cs+). depend on the concentration and nature of local defects.
Single-​crystal nanostructures and/or microstructures
of diverse perovskites with different compositions and Ion exchange. Nanoscale heterostructures of perovskites
properties have enabled comparative studies of per­ provide model systems for mechanistic and kinetic
ovskite materials34–36. For example, hot fluorescence studies of solid-​state ion exchange77. The ion interdiffu­
emission at room temperature from energetic carriers sion kinetics of heterostructures fabricated by stacking
with lifetimes of ~102 ps was observed from hybrid CsPbCl3 microplates on top of CsPbBr3 NW networks
perovskites (Fig. 7e) but not from CsPbBr3 (ref.159). This have been visualized using spatially resolved PL meas­
observation suggests that in hybrid perovskites, there urements. Anion interdiffusion readily occurs between
is strong screening of the Coulomb potential owing to CsPbCl3 and CsPbBr3, which results in halide concen­
the formation of large polarons, which are probably of tration gradients in the NWs (Fig. 7h). Quantitative anal­
a ferroelectric nature15. The much reduced Coulomb ysis of the temperature-​dependent composition profiles
potential decreases hot electron cooling through elec­ yields an interdiffusion coefficient of ~10−13 cm2 s−1
tron–longitudinal-​optical-phonon scattering. Large at room temperature and a low activation energy of
polaron formation in MAPbBr3 and CsPbBr3 occurs 0.44 eV. By contrast, the activation energy reported
on timescales of 0.3 ps and 0.7 ps, respectively; only the for InxGa1−xAs/GaAs is 3.0 eV at 900–1,150 °C (ref.169).
former is competitive with longitudinal optical pho­ In contrast to the anions, the interdiffusion of the
non cooling, leading to the partial retention of excess A-​site cations in the MAPbBr3/CsPbBr3 system is much
electronic energy in MAPbBr3 (ref.160). Further cooling slower. These insights provide guidelines for designing
of hot polarons is slowed to the timescale of 100 ps perovskite heterostructures with tailored properties for
owing to the low thermal conductivity of perovskites161. practical applications.
Dynamic disordering of the cations was recently sug­
gested to be responsible for the suppressed thermal Light–matter interactions. Perovskite NWs have
conductivity of MAPbBr3 NWs compared with that of recently emerged as promising platforms for demon­
CsPbBr3 NWs162. In addition, by comparing the band-​ strating and manipulating strong light–matter interac­
edge carrier dynamics of single-​crystal nanoplates of tions. Strong coupling between excitons and confined
different perovskites, similar low carrier trapping rate cavity photons can result in the formation of exciton–
constants (∼107 s−1) and electron–hole radiative recom­ polaritons (Fig. 7i), which are of mixed exciton and pho­
bination rate constants (∼10−10 cm3 s−1) were found and ton character. The mass of an exciton–polariton can be
shown to be independent of the cation type163 (Fig. 7f). as low as 10−5 me (where me is the electron rest mass);
These results suggest that the remarkable properties the corresponding long de Broglie wavelength of these
of photoexcited carriers in perovskites are inherent to bosonic quasiparticles makes them exciting platforms
these materials, triggering intense studies and discus­ to achieve Bose–Einstein condensation at room tem­
sion regarding the mechanisms of their defect tolerance perature170,171. As the exciton–polariton condensate is a
and low recombination rates14–17. macroscopic quantum coherent state, light escaping the
Carrier transport properties in single-​crystal nanos­ cavity is coherent and laser-​like. This can be the basis
tructures of 3D perovskites have been studied using spa­ for low-​threshold lasing without population inversion.
tially resolved optical techniques. Using time-​resolved The robustness of an exciton–polariton is determined
and PL-​scanning imaging microscopy, the carrier dif­ by both the exciton binding energy and the Rabi split­
fusion processes in single-​crystal MAPbX3 NWs and ting energy (Fig. 7j). The latter quantifies the coupling
nanoplates25 were directly visualized and quantitatively strength between an exciton and a cavity photon.
characterized (Fig. 7g). Fitting the PL kinetics with a The spacings of the Fabry–Pérot modes in CsPbBr3
diffusion model yielded a carrier mobility and diffu­ NWs are not identical but decrease as the mode energy
sion length of 80 cm2 V−1 s−1 and 14 μm, respectively, increases44 (Fig. 7k), which is inconsistent with the pho­
for MAPbI3 and 34 cm2 V−1 s−1 and 6 μm, respectively, for tonic model. The decreasing mode spacing with energy

www.nature.com/natrevmats
Reviews

(that is, the negative curvature in Fig. 7k) can be attrib­ avenues for exploring RP perovskites with improved
uted to strong light–matter interactions. There are two optical properties and stability29. The structural stabili­
interpretations of this strong light–matter interaction. zation by surface ligands further expands the library of
In the classical interpretation, as the emission energy compositions and structural diversity121,123. The chemis­
approaches an electronic resonance, the refractive try of some of these perovskite or perovskite-​like mate­
index increases, which is reflected in the decrease in rials may be quite different from that of the 3D APbX3
mode spacing. In the alternative quantum-​mechanical perovskites. Given the structural diversity, there is much
explanation, for an exciton not in resonance with a cav­ scope for creative new nanomaterial syntheses.
ity photon energy, the two dispersion curves cross at a Considering the advantages of perovskite hetero­
momentum away from zero. The avoided crossing due structures, including greater control of the generation,
to the Rabi splitting energy results in a lower polariton recombination and transport of charge carriers and the
branch and an upper polariton branch (Fig. 7j). The lower possibility of multifunctional properties, we anticipate
polariton branch in the avoided crossing region, called increased efforts and progress in the fabrication of
the bottleneck region, has negative curvature and can functional perovskite heterostructures. Direct growth
account for the nonlinear dispersions139,172,173. Analysis of
of heterostructures of different 3D perovskites remains
the polariton dispersions reveals Rabi splitting energies difficult owing to the fast ion-​exchange kinetics. In par­
of hundreds of meV (for example, up to 656 meV for ticular, fast anion exchange fundamentally limits the
CsPbBr3), and the value increases as the dimensions of fabrication of a stable heterojunction with different hal­
the NWs decrease173. Notably, the Rabi splitting energies ides77. However, solid-​state cation exchange seems to
of perovskites are much larger than those of conventional be much slower. Heterojunctions between perovskites
semiconductor NWs (for example, ~25 mV for GaAs with different cations have not yet been reported,
and CdTe, ~200 meV for CdS and ~150 meV for ZnO), and the long-​term junction stability remains unclear.
which suggests strong light–matter interactions in per­ Heterostructures of RP perovskites are also of interest.
ovskite NWs173. The exciton–photon coupling strength In RP perovskites, ion migration is suppressed along
can be further increased in a MAPbBr3 NW/SiO2/Ag both in-​plane and out-​of-plane directions181,182, which
cavity owing to the strong localized excitation field and presents opportunities to fabricate atomically sharp
reduction in the effective mode volume induced by a junctions in either vertical or lateral arrangements183.
surface plasmon174. Recent studies have also demonstrated the feasibility of
There is, however, a need to be cautious in using the constructing heterostructures of 2D perovskites with
exciton–polariton interpretation. In addition to the usual other 2D-​layered materials184,185. It will be interesting
criteria of threshold behaviour in emission intensity and to explore the integration of perovskites with other
peak width, an exciton–polariton condensate is charac­ inorganic or organic semiconductors and functional
terized by both temporal and spatial coherence. A quan­ oxide perovskites51, and to investigate the emerging
titative measurement of a spatial coherent state should multifunctional properties.
have a characteristic dependence on excitation density Although perovskite NWs have shown great promise
for an exciton–polariton condensate175,176. Unfortunately, in lasing applications, challenges remain to realize the
such measurements have not yet been conducted on ultimate goal of electrically driven lasing. Developing
exciton–polaritons in perovskite nanostructures or new device architectures and fabrication strategies to
microstructures139,172,173,177. improve electrical injection, EL performance and device
stability are high priorities. Arrays of NWs with precisely
Future perspectives controlled orientations and positions could be a more
Many exciting aspects of perovskite nanostructures are practical platform for device integration and could
enabled by their facile crystal growth, soft and ionic enable a thin-​film type of device configuration. In this
crystal lattice, ion migration, surface chemistry, defect-​ regard, strategies need to be developed to grow52,63 or
tolerant nature and ferroelectric-​like large polarons. As assemble NWs into arrays186,187. Another major challenge
the field is still rapidly growing, many new opportunities is the thermal heating associated with high injected
and challenges are ahead of us. current densities, which may cause perovskite degra­
The library of perovskite nanostructures will continue dation. Given the ultralow thermal conductivity in per­
to expand. Replacing or partially substituting lead with ovskites188, strategies should be developed to effectively
non-​toxic metal ions will result in more environmentally dissipate heat from perovskites to substrates. In addition
friendly compounds with new properties and function­ to optoelectronic properties, owing to the soft lattices,
alities. Lead-​free perovskites such as ASnX3, Cs2AgBiBr6 it could also be interesting to explore the mechani­
and A3M2X9 (M = Bi or Sb) have been identified as prom­ cal and thermal properties162,188 of these perovskite
ising semiconductors with lower toxicity178. Another nanostructures.
notable example is germanium halide perovskites, which 2D perovskites are a promising alternative to organic
are a class of non-​centrosymmetric compounds that emitter and inorganic quantum dots for low-​c ost,
exhibit strong second-​harmonic generation179. CsGeI3 high-​p erformance light-​emitting devices. However,
is predicted to have an electro-​optic coefficient that owing to the structural complexity and lability of 2D
surpasses the highest value of LiNbO3 at the telecom­ perovskites, it is important to understand and control
munication wavelength of 1,550 nm (ref.180). Moreover, the composition and phase stacking in these solution-​
within the modular structural framework of perovskites, processed films to further increase device efficiency.
a large library of diverse organic cations offers promising Processing conditions for films with high PLQY and

Nature Reviews | Materials


Reviews

the underlying crystal-​growth mechanism have yet to could be studied using these nanostructures. Unlike
be determined. It was recently demonstrated that the the 3D perovskites with a limited number of possible
energy-​cascade structure can enhance gain properties A cations, the structural diversity of RP perovskites may
through fast energy transfer, which enables efficient enable us to understand the mechanistic role of the
optically pumped lasing from the smallest-​bandgap A cation on photophysical properties and to establish
phases189. Considering the rapid increase in EL perfor­ structure–property relationships. The variety of pure
mance, these 2D perovskite films promise the possibility n-​value RP perovskites, if their compositions could
of realizing electrically driven lasing, which has been an be controlled more precisely, offer a unique platform
elusive long-​standing goal in solution-​processed opto­ to better understand the effects of quantum confine­
electronics. Although 2D perovskites and their nanos­ ment. More in-​depth investigations on the fundamental
tructures in the strong quantum-​confinement regime physics of RP perovskites, such as polaron physics and
could be especially useful for LEDs in the blue region exciton–photon coupling191–193, might enable increased
of the spectrum147, little progress has been made so far solar or LED device performance and new applications,
because the synthesis methods tend to yield 3D-​like such as exciton–polariton lasing and polaritonics. As
perovskite phases. Creative synthesis with exquisite con­ exciton binding energies are higher in RP perovskites
trol over the RP phase compositions190 and potentially than in their 3D counterparts, stronger light–matter
over heterostructures, either in isolated nanostructures interactions and potentially lower lasing thresholds
or microstructures183 or in nanostructured thin films, may be expected in RP perovskites. The opportuni­
would accelerate the advances in efficient light-​emitting ties and applications of 1D and 2D perovskite nano­
devices. structures will be addressed in the years to come, and
Single-​crystal perovskite nanostructures will remain the future of metal halide perovskite nanostructures
interesting from the perspective of fundamental stud­ is bright.
ies. For example, there are several proposals on how a
polar A-​site cation contributes to charge screening that Published online xx xx xxxx

1. Hu, J., Odom, T. W. & Lieber, C. M. Chemistry and 18. Yina, W.-J., Shi, T. & Yan, Y. Unusual defect physics in 32. Fu, Y. et al. Solution growth of single crystal
physics in one dimension: synthesis and properties of CH3NH3PbI3 perovskite solar cell absorber. Appl. methylammonium lead halide perovskite
nanowires and nanotubes. Acc. Chem. Res. 32, Phys. Lett. 104, 063903 (2014). nanostructures for optoelectronic and photovoltaic
435–445 (1999). 19. Miyata, K., Atallah, T. L. & Zhu, X.-Y. Lead halide applications. J. Am. Chem. Soc. 137, 5810–5818
2. Xia, Y. et al. One-​dimensional nanostructures: perovskites: crystal-​liquid duality, phonon glass (2015).
synthesis, characterization, and applications. Adv. electron crystals, and large polaron formation. 33. Meng, F., Morin, S. A., Forticaux, A. & Jin, S.
Mater. 15, 353–389 (2003). Sci. Adv. 3, e1701469 (2017). Screw dislocation driven growth of nanomaterials.
3. Zhang, A. & Lieber, C. M. Nano-​bioelectronics. Chem. 20. Rakita, Y., Cohen, S. R., Kedem, N. K., Hodes, G. Acc. Chem. Res. 46, 1616–1626 (2013).
Rev. 116, 215–257 (2016). & Cahen, D. Mechanical properties of APbX3 (A = Cs 34. Fu, Y. et al. Nanowire lasers of formamidinium lead
4. Yan, R., Gargas, D. & Yang, P. Nanowire photonics. or CH3NH3; X = I or Br) perovskite single crystals. halide perovskites and their stabilized alloys with
Nat. Photonics 3, 569–576 (2009). MRS Commun. 5, 623–629 (2015). improved stability. Nano Lett. 16, 1000–1008 (2016).
5. Kojima, A., Teshima, K., Shirai, Y. & Miyasaka, T. 21. Zhu, X. Y. & Podzorov, V. Charge carriers in hybrid 35. Fu, Y. et al. Broad wavelength tunable robust lasing
Organometal halide perovskites as visible-​light organic–inorganic lead halide perovskites might be from single-​crystal nanowires of cesium lead halide
sensitizers for photovoltaic cells. J. Am. Chem. Soc. protected as large polarons. J. Phys. Chem. Lett. 6, perovskites (CsPbX3, X = Cl, Br, I). ACS Nano 10,
131, 6050–6051 (2009). 4758–4761 (2015). 7963–7972 (2016).
6. Lee, M. M., Teuscher, J., Miyasaka, T., Murakami, T. N. 22. Yang, Y. et al. Low surface recombination velocity in 36. Eaton, S. W. et al. Lasing in robust cesium lead halide
& Snaith, H. J. Efficient hybrid solar cells based on solution-​grown CH3NH3PbBr3 perovskite single crystal. perovskite nanowires. Proc. Natl Acad. Sci. USA 113,
meso-​superstructured organometal halide perovskites. Nat. Commun. 6, 7961 (2015). 1993–1998 (2016).
Science 338, 643–647 (2012). 23. Fang, H.-H. et al. Ultrahigh sensitivity of 37. Dai, J. et al. Carrier decay properties of mixed cation
7. Xing, G. et al. Low-​temperature solution-​processed methylammonium lead tribromide perovskite single formamidinium–methylammonium lead iodide
wavelength-​tunable perovskites for lasing. Nat. Mater. crystals to environmental gases. Sci. Adv. 2, perovskite [HC(NH2)2]1−x[CH3NH3]xPbI3 nanorods.
13, 476–480 (2014). e1600534 (2016). J. Phys. Chem. Lett. 7, 5036–5043 (2016).
8. Deschler, F. et al. High photoluminescence efficiency 24. Brandt, R. E. et al. Searching for “defect-​tolerant” 38. Wong, A. B. et al. Growth and anion exchange
and optically pumped lasing in solution-​processed photovoltaic materials: combined theoretical and conversion of CH3NH3PbX3 nanorod arrays for
mixed halide perovskite semiconductors. J. Phys. experimental screening. Chem. Mater. 29, light-​emitting diodes. Nano Lett. 15, 5519–5524
Chem. Lett. 5, 1421–1426 (2014). 4667–4674 (2017). (2015).
9. Zhang, Q., Ha, S. T., Liu, X., Sum, T. C. & Xiong, Q. 25. Tian, W., Zhao, C., Leng, J., Cui, R. & Jin, S. Visualizing 39. Walsh, A., Scanlon, D. O., Chen, S., Gong, X. & Wei, S. H.
Room-​temperature near-​infrared high-​Q perovskite carrier diffusion in individual single-​crystal organolead Self-​regulation mechanism for charged point defects in
whispering-​gallery planar nanolasers. Nano Lett. 14, halide perovskite nanowires and nanoplates. J. Am. hybrid halide perovskites. Angew. Chem. Int. Ed. 54,
5995–6001 (2014). Chem. Soc. 137, 12458–12461 (2015). 1791–1794 (2015).
10. Zhu, H. et al. Lead halide perovskite nanowire lasers 26. Xiao, R. et al. Photocurrent mapping in single-​ 40. Eames, C. et al. Ionic transport in hybrid lead iodide
with low lasing thresholds and high quality factors. crystal methylammonium lead iodide perovskite perovskite solar cells. Nat. Commun. 6, 7497 (2015).
Nat. Mater. 14, 636–642 (2015). nanostructures. Nano Lett. 16, 7710–7717 41. Ha, S. T. et al. Synthesis of organic–inorganic lead
11. Correa-​Baena, J.-P. et al. Promises and challenges of (2016). halide perovskite nanoplatelets: towards high-​
perovskite solar cells. Science 358, 739–744 27. Darbandi, A. & Watkins, S. P. Measurement of performance perovskite solar cells and optoelectronic
(2017). minority carrier diffusion lengths in GaAs nanowires devices. Adv. Opt. Mater. 2, 838–844 (2014).
12. Akkerman, Q. A., Rainò, G., Kovalenko, M. V. & by a nanoprobe technique. J. Appl. Phys. 120, 42. Meyers, J. K. et al. Self-​catalyzed vapor–liquid–solid
Manna, L. Genesis, challenges and opportunities for 014301 (2016). growth of lead halide nanowires and conversion to
colloidal lead halide perovskite nanocrystals. Nat. 28. Triplett, M. et al. Long minority carrier diffusion hybrid perovskites. Nano Lett. 17, 7561–7568
Mater. 17, 394–405 (2018). lengths in bridged silicon nanowires. Nano Lett. 15, (2017).
13. Brandt, R. E., Stevanovic, V., Ginley, D. S. & 523–529 (2015). 43. Chen, J. et al. Vapor-​phase epitaxial growth of aligned
Buonassisi, T. Identifying defect-​tolerant 29. Saparov, B. & Mitzi, D. B. Organic–inorganic nanowire networks of cesium lead halide perovskites
semiconductors with high minority-​carrier lifetimes: perovskites: structural versatility for functional (CsPbX3, X = Cl, Br, I). Nano Lett. 17, 460–466
beyond hybrid lead halide perovskites. MRS Commun. materials design. Chem. Rev. 116, 4558–4596 (2017).
5, 265–275 (2015). (2016). 44. Park, K. et al. Light–matter interactions in cesium
14. Katan, C., Mohite, A. D. & Even, J. Entropy in halide 30. Moore, D. T. et al. Crystallization kinetics of lead halide perovskite nanowire lasers. J. Phys. Chem.
perovskites. Nat. Mater. 17, 377–379 (2018). organic–inorganic trihalide perovskites and the role of Lett. 7, 3703–3710 (2016).
15. Miyata, K. & Zhu, X. Y. Ferroelectric large polarons. the lead anion in crystal growth. J. Am. Chem. Soc. 45. Zhou, H. et al. Vapor growth and tunable lasing of
Nat. Mater. 17, 379–381 (2018). 137, 2350–2358 (2015). band gap engineered cesium lead halide perovskite
16. Stranks, S. D. & Plochocka, P. The influence of the 31. Liang, K., Mitzi, D. B. & Prikas, M. T. Synthesis micro/nanorods with triangular cross section. ACS
Rashba effect. Nat. Mater. 17, 381–382 (2018). and characterization of organic−inorganic perovskite Nano 11, 1189–1195 (2017).
17. De Angelis, F. & Petrozza, A. Clues from defect thin films prepared using a versatile two-​step 46. Wang, Y. et al. Photon transport in one-​dimensional
photochemistry. Nat. Mater. 17, 383–384 dipping technique. Chem. Mater. 10, 403–411 incommensurately epitaxial CsPbX3 arrays. Nano Lett.
(2018). (1998). 16, 7974–7981 (2016).

www.nature.com/natrevmats
Reviews

47. Zhang, Q. et al. High-​quality whispering-​gallery-mode 73. Petrov, A. A. et al. New insight into the formation of their oriented assemblies. J. Am. Chem. Soc. 137,
lasing from cesium lead halide perovskite nanoplatelets. hybrid perovskite nanowires via structure directing 16008–16011 (2015).
Adv. Funct. Mater. 26, 6238–6245 (2016). adducts. Chem. Mater. 29, 587–594 (2017). 99. Bohn, B. J. et al. Boosting tunable blue luminescence
48. Oksenberg, E., Sanders, E., Popovitz-​Biro, R., 74. Gupta, S., Kershaw, S. V. & Rogach, A. L. 25th of halide perovskite nanoplatelets through
Houben, L. & Joselevich, E. Surface-​guided CsPbBr3 Anniversary article: ion exchange in colloidal postsynthetic surface trap repair. Nano Lett. 18,
perovskite nanowires on flat and faceted sapphire nanocrystals. Adv. Mater. 25, 6923–6944 (2013). 5231–5238 (2018).
with size-dependent photoluminescence and fast 75. Nedelcu, G. et al. Fast anion-​exchange in highly 100. Wu, Y. et al. In situ passivation of PbBr64− octahedra
photoconductive response. Nano Lett. 18, 424–433 luminescent nanocrystals of cesium lead halide toward blue luminescent CsPbBr3 nanoplatelets with
(2017). perovskites (CsPbX3, X = Cl, Br, I). Nano Lett. 15, near 100% absolute quantum yield. ACS Energy Lett.
49. Wang, Y. et al. High-​temperature ionic epitaxy of 5635–5640 (2015). 3, 2030–2037 (2018).
halide perovskite thin film and the hidden carrier 76. Dou, L. et al. Spatially resolved multicolor CsPbX3 101. Shamsi, J. et al. Colloidal synthesis of quantum
dynamics. Adv. Mater. 29, 1702643 (2017). nanowire heterojunctions via anion exchange. confined single crystal CsPbBr3 nanosheets with
50. Lili, W. et al. Unlocking the single-​domain epitaxy of Proc. Natl Acad. Sci. USA 114, 7216–7221 (2017). lateral size control up to the micrometer range. J. Am.
halide perovskites. Adv. Mater. Interfaces 4, 1701003 77. Pan, D. et al. Visualization and studies of ion-​diffusion Chem. Soc. 138, 7240–7243 (2016).
(2017). kinetics in cesium lead bromide perovskite nanowires. 102. Liu, J. et al. Two-​dimensional CH3NH3PbI3 perovskite:
51. Chen, J. et al. Single-​crystal thin films of cesium lead Nano Lett. 18, 1807–1813 (2018). synthesis and optoelectronic application. ACS Nano
bromide perovskite epitaxially grown on metal oxide 78. Tian, W., Leng, J., Zhao, C. & Jin, S. Long-​distance 10, 3536–3542 (2016).
perovskite (SrTiO3). J. Am. Chem. Soc. 139, charge carrier funneling in perovskite nanowires 103. Wang, Y., Shi, Y., Xin, G., Lian, J. & Shi, J.
13525–13532 (2017). enabled by built-​in halide gradient. J. Am. Chem. Soc. Two-​dimensional van der Waals epitaxy kinetics in a
52. Shoaib, M. et al. Directional growth of ultralong 139, 579–582 (2017). three-​dimensional perovskite halide. Cryst. Growth
CsPbBr3 perovskite nanowires for high-​performance 79. Wang, Y. et al. Epitaxial halide perovskite lateral Des. 15, 4741–4749 (2015).
photodetectors. J. Am. Chem. Soc. 139, double heterostructure. ACS Nano 11, 3355–3364 104. Ma, C. et al. Heterostructured WS2/CH3NH3PbI3
15592–15595 (2017). (2017). photoconductors with suppressed dark current
53. Schmidt, L. C. et al. Nontemplate synthesis of 80. Kong, Q. et al. Phase-​transition–induced p-​n junction and enhanced photodetectivity. Adv. Mater. 28,
CH3NH3PbBr3 perovskite nanoparticles. J. Am. Chem. in single halide perovskite nanowire. Proc. Natl Acad. 3683–3689 (2016).
Soc. 136, 850–853 (2014). Sci. USA 115, 8889–8894 (2018). 105. Cheng, H.-C. et al. van der Waals heterojunction
54. Zhu, F. et al. Shape evolution and single particle 81. Nasilowski, M., Mahler, B., Lhuillier, E., Ithurria, S. devices based on organohalide perovskites and
luminescence of organometal halide perovskite & Dubertret, B. Two-​dimensional colloidal two-​dimensional materials. Nano Lett. 16, 367–373
nanocrystals. ACS Nano 9, 2948–2959 (2015). nanocrystals. Chem. Rev. 116, 10934–10982 (2016).
55. Zhang, D., Eaton, S. W., Yu, Y., Dou, L. & Yang, P. (2016). 106. Leguy, A. M. A. et al. Reversible hydration of
Solution-​phase synthesis of cesium lead halide 82. Weidman, M. C., Goodman, A. J. & Tisdale, W. A. CH3NH3PbI3 in films, single crystals, and solar cells.
perovskite nanowires. J. Am. Chem. Soc. 137, Colloidal halide perovskite nanoplatelets: an exciting Chem. Mater. 27, 3397–3407 (2015).
9230–9233 (2015). new class of semiconductor nanomaterials. Chem. 107. Smith, I. C., Hoke, E. T., Solis-​Ibarra, D., McGehee, M. D.
56. Zhang, D. et al. Synthesis of composition tunable Mater. 29, 5019–5030 (2017). & Karunadasa, H. I. A layered hybrid perovskite
and highly luminescent cesium lead halide nanowires 83. Stoumpos, C. C. et al. Ruddlesden–Popper hybrid lead solar-​cell absorber with enhanced moisture stability.
through anion-​exchange reactions. J. Am. Chem. Soc. iodide perovskite 2D homologous semiconductors. Angew. Chem. Int. Ed. 126, 11414–11417 (2014).
138, 7236–7239 (2016). Chem. Mater. 28, 2852–2867 (2016). 108. Ling, Y. et al. Bright light-​emitting diodes based on
57. Zhang, D. et al. Ultrathin colloidal cesium lead halide 84. Stoumpos, C. C. et al. High members of the 2D organometal halide perovskite nanoplatelets. Adv.
perovskite nanowires. J. Am. Chem. Soc. 138, Ruddlesden-​Popper halide perovskites: synthesis, Mater. 28, 305–311 (2016).
13155–13158 (2016). optical properties, and solar cells of 109. Yang, S. et al. Functionalization of perovskite thin films
58. Tong, Y. et al. From precursor powders to CsPbX3 (CH3(CH2)3NH3)2(CH3NH3)4Pb5I16. Chem 2, 427–440 with moisture-​tolerant molecules. Nat. Energy 1,
perovskite nanowires: one-​pot synthesis, growth (2017). 15016 (2016).
mechanism, and oriented self-​assembly. Angew. Chem. 85. Mao, L. et al. Hybrid Dion–Jacobson 2D lead iodide 110. Ma, Z.-Q., Shao, Y., Wong, P. K., Shi, X. & Pan, H.
Int. Ed. 56, 13887–13892 (2017). perovskites. J. Am. Chem. Soc. 140, 3775–3783 Structural and electronic properties of two-​
59. Imran, M. et al. Colloidal synthesis of strongly (2018). dimensional organic–inorganic halide perovskites and
fluorescent CsPbBr3 nanowires with width tunable 86. Gong, X. et al. Electron–phonon interaction in efficient their stability against moisture. J. Phys. Chem. C 122,
down to the quantum confinement regime. Chem. perovskite blue emitters. Nat. Mater. 17, 550–556 5844–5853 (2018).
Mater. 28, 6450–6454 (2016). (2018). 111. Liu, J., Leng, J., Wu, K., Zhang, J. & Jin, S.
60. Ashley, M. J. et al. Templated synthesis of uniform 87. Niu, W., Eiden, A., Prakash, G. V. & Baumberg, J. J. Observation of internal photoinduced electron
perovskite nanowire arrays. J. Am. Chem. Soc. 138, Exfoliation of self-​assembled 2D organic-​inorganic and hole separation in hybrid two-​dimentional
10096–10099 (2016). perovskite semiconductors. Appl. Phys. Lett. 104, perovskite films. J. Am. Chem. Soc. 139,
61. Waleed, A. et al. All inorganic cesium lead iodide 171111 (2014). 1432–1435 (2017).
perovskite nanowires with stabilized cubic phase at 88. Yaffe, O. et al. Excitons in ultrathin organic-​inorganic 112. Shang, Q. et al. Unveiling structurally engineered
room temperature and nanowire array-​based perovskite crystals. Phys. Rev. B 92, 045414 (2015). carrier dynamics in hybrid quasi-​two-dimensional
photodetectors. Nano Lett. 17, 4951–4957 (2017). 89. Leng, K. et al. Molecularly thin two-​dimensional perovskite thin films toward controllable emission.
62. Oener, S. Z. et al. Perovskite nanowire extrusion. hybrid perovskites with tunable optoelectronic J. Phys. Chem. Lett. 8, 4431–4438 (2017).
Nano Lett. 17, 6557–6563 (2017). properties due to reversible surface relaxation. 113. Quan, L. N. et al. Ligand-​stabilized reduced-​
63. Liu, P. et al. Organic–inorganic hybrid perovskite Nat. Mater. 17, 908–914 (2018). dimensionality perovskites. J. Am. Chem. Soc. 138,
nanowire laser arrays. ACS Nano 11, 5766–5773 90. Dou, L. et al. Atomically thin two-​dimensional 2649–2655 (2016).
(2017). organic-​inorganic hybrid perovskites. Science 349, 114. Tsai, H. et al. High-​efficiency two-​dimensional
64. Jeong, B. et al. Solvent-​assisted gel printing for 1518–1521 (2015). Ruddlesden–Popper perovskite solar cells. Nature
micropatterning thin organic–inorganic hybrid 91. Ma, D. et al. Single-​crystal microplates of two-​ 536, 312–316 (2016).
perovskite films. ACS Nano 10, 9026–9035 (2016). dimensional organic–inorganic lead halide layered 115. McHale, J. M., Auroux, A., Perrotta, A. J. &
65. Kamminga, M. E. et al. Micropatterned 2D hybrid perovskites for optoelectronics. Nano Res. 10, Navrotsky, A. Surface energies and thermodynamic
perovskite thin films with enhanced 2117–2129 (2017). phase stability in nanocrystalline aluminas. Science
photoluminescence lifetimes. ACS Appl. Mater. 92. Tyagi, P., Arveson, S. M. & Tisdale, W. A. Colloidal 277, 788–791 (1997).
Interfaces 10, 12878–12885 (2018). organohalide perovskite nanoplatelets exhibiting 116. Protesescu, L. et al. Nanocrystals of cesium lead
66. He, X. et al. Patterning multicolored microdisk laser quantum confinement. J. Phys. Chem. Lett. 6, halide perovskites (CsPbX3, X = Cl, Br, and I): novel
arrays of cesium lead halide perovskite. Adv. Mater. 1911–1916 (2015). optoelectronic materials showing bright emission with
29, 1604510 (2017). 93. Sichert, J. A. et al. Quantum size effect in organometal wide color gamut. Nano Lett. 15, 3692–3696
67. Chen, Y.-X. et al. General space-​confined on-​substrate halide perovskite nanoplatelets. Nano Lett. 15, (2015).
fabrication of thickness-​adjustable hybrid perovskite 6521–6527 (2015). 117. Swarnkar, A. et al. Quantum dot–induced phase
single-​crystalline thin films. J. Am. Chem. Soc. 138, 94. Yang, S. et al. Ultrathin two-​dimensional organic– stabilization of α-​CsPbI3 perovskite for high-​efficiency
16196–16199 (2016). inorganic hybrid perovskite nanosheets with bright, photovoltaics. Science 354, 92–95 (2016).
68. Horváth, E. et al. Nanowires of methylammonium lead tunable photoluminescence and high stability. Angew. 118. Zhao, B. et al. Thermodynamically stable
iodide (CH3NH3PbI3) prepared by low temperature Chem. Int. Ed. 56, 4252–4255 (2017). orthorhombic γ-​CsPbI3 thin films for high-​performance
solution-​mediated crystallization. Nano Lett. 14, 95. Weidman, M. C., Seitz, M., Stranks, S. D. & Tisdale, W. A. photovoltaics. J. Am. Chem. Soc. 140, 11716–11725
6761–6766 (2014). Highly tunable colloidal perovskite nanoplatelets (2018).
69. Deng, W. et al. Aligned single-​crystalline perovskite through variable cation, metal, and halide 119. Levchuk, I. et al. Brightly luminescent and color-​
microwire arrays for high-​performance flexible image composition. ACS Nano 10, 7830–7839 (2016). tunable formamidinium lead halide perovskite FAPbX3
sensors with long-​term stability. Adv. Mater. 28, 96. Akkerman, Q. A. et al. Solution synthesis approach to (X = Cl, Br, I) colloidal nanocrystals. Nano Lett. 17,
2201–2208 (2016). colloidal cesium lead halide perovskite nanoplatelets 2765–2770 (2017).
70. Gao, L. et al. Passivated single-​crystalline CH3NH3PbI3 with monolayer-​level thickness control. J. Am. Chem. 120. Protesescu, L. et al. Dismantling the “red wall”
nanowire photodetector with high detectivity and Soc. 138, 1010–1016 (2016). of colloidal perovskites: highly luminescent
polarization sensitivity. Nano Lett. 16, 7446–7454 97. Wheeler, L. M., Anderson, N. C., Bliss, T. S., formamidinium and formamidinium–cesium lead
(2016). Hautzinger, M. P. & Neale, N. R. Dynamic evolution of iodide nanocrystals. ACS Nano 11, 3119–3134
71. Im, J.-H. et al. Nanowire perovskite solar cell. Nano 2D layers within perovskite nanocrystals via salt pair (2017).
Lett. 15, 2120–2126 (2015). extraction and reinsertion. J. Phys. Chem. C 122, 121. Fu, Y. et al. Stabilization of the metastable lead iodide
72. Zhu, P. et al. Direct conversion of perovskite thin films 14029–14038 (2018). perovskite phase via surface functionalization. Nano
into nanowires with kinetic control for flexible 98. Bekenstein, Y., Koscher, B. A., Eaton, S. W., Yang, P. Lett. 17, 4405–4414 (2017).
optoelectronic devices. Nano Lett. 16, 871–876 & Alivisatos, A. P. Highly luminescent colloidal 122. Wang, C., Chesman, A. S. R. & Jasieniak, J. J.
(2016). nanoplates of perovskite cesium lead halide and Stabilizing the cubic perovskite phase of CsPbI3

Nature Reviews | Materials


Reviews

nanocrystals by using an alkyl phosphinic acid. Chem. 148. Yuan, M. et al. Perovskite energy funnels for efficient perovskite nanowires. Nano Lett. 18, 3335–3343
Commun. 53, 232–235 (2017). light-​emitting diodes. Nat. Nanotechnol. 11, 872–877 (2018).
123. Fu, Y. et al. Selective stabilization and photophysical (2016). 175. Deng, H., Weihs, G., Santori, C., Bloch, J. &
properties of metastable perovskite polymorphs of 149. Wang, N. et al. Perovskite light-​emitting diodes based Yamamoto, Y. Condensation of semiconductor
CsPbI3 in thin films. Chem. Mater. 29, 8385–8394 on solution-​processed self-​organized multiple quantum microcavity exciton polaritons. Science 298,
(2017). wells. Nat. Photonics 10, 699–704 (2016). 199–202 (2002).
124. Zhang, T. et al. Bication lead iodide 2D perovskite 150. Byun, J. et al. Efficient visible quasi-2D perovskite 176. Klaas, M. et al. Evolution of temporal coherence in
component to stabilize inorganic α-​CsPbI3 perovskite light-​emitting diodes. Adv. Mater. 28, 7515–7520 confined exciton-​polariton condensates. Phys. Rev.
phase for high-​efficiency solar cells. Sci. Adv. 3, (2016). Lett. 120, 017401 (2018).
e1700841 (2017). 151. Zhang, S. et al. Efficient red perovskite light-​emitting 177. Su, R. et al. Room-​temperature polariton lasing in
125. Wang, Q. et al. Stabilizing the α-​phase of CsPbI3 diodes based on solution-​processed multiple quantum all-​inorganic perovskite nanoplatelets. Nano Lett. 17,
perovskite by sulfobetaine zwitterions in one-​step wells. Adv. Mater. 29, 1606600 (2017). 3982–3988 (2017).
spin-​coating films. Joule 1, 371–382 (2017). 152. Yang, X. et al. Efficient green light-​emitting diodes 178. Giustino, F. & Snaith, H. J. Toward lead-​free perovskite
126. Li, B. et al. Surface passivation engineering strategy based on quasi-​two-dimensional composition and solar cells. ACS Energy Lett. 1, 1233–1240 (2016).
to fully-​inorganic cubic CsPbI3 perovskites for high-​ phase engineered perovskite with surface passivation. 179. Stoumpos, C. C. et al. Hybrid germanium iodide
performance solar cells. Nat. Commun. 9, 1076 Nat. Commun. 9, 570 (2018). perovskite semiconductors: active lone pairs,
(2018). 153. Zou, W. et al. Minimising efficiency roll-​off in high-​ structural distortions, direct and indirect energy gaps,
127. Fang, H.-H. et al. Unravelling light-​induced brightness perovskite light-​emitting diodes. Nat. and strong nonlinear optical properties. J. Am. Chem.
degradation of layered perovskite crystals and Commun. 9, 608 (2018). Soc. 137, 6804–6819 (2015).
design of efficient encapsulation for improved 154. Chen, Z. et al. High-​performance color-​tunable 180. Walters, G. & Sargent, E. H. Electro-​optic response in
photostability. Adv. Funct. Mater. 28, 1800305 perovskite light emitting devices through structural germanium halide perovskites. J. Phys. Chem. Lett. 9,
(2018). modulation from bulk to layered film. Adv. Mater. 29, 1018–1027 (2018).
128. Fan, Z. et al. Layer-​by-layer degradation of 1603157 (2017). 181. Lin, Y. et al. Suppressed ion migration in low-​
methylammonium lead tri-​iodide perovskite 155. Congreve, D. N. et al. Tunable light-​emitting diodes dimensional perovskites. ACS Energy Lett. 2,
microplates. Joule 1, 548–562 (2017). utilizing quantum-​confined layered perovskite 1571–1572 (2017).
129. Sutherland, B. R. & Sargent, E. H. Perovskite photonic emitters. ACS Photonics 4, 476–481 (2017). 182. Xiao, X. et al. Suppressed ion migration along the
sources. Nat. Photonics 10, 295–302 (2016). 156. Sadhanala, A. et al. Blue-​green color tunable solution in-​plane direction in layered perovskites. ACS Energy
130. Huang, M. H. et al. Room-​temperature ultraviolet processable organolead chloride–bromide mixed Lett. 3, 684–688 (2018).
nanowire nanolasers. Science 292, 1897–1899 halide perovskites for optoelectronic applications. 183. Fu, Y. et al. Multicolor heterostructures of two-​
(2001). Nano Lett. 15, 6095–6101 (2015). dimensional layered halide perovskites that show
131. Duan, X., Huang, Y., Agarwal, R. & Lieber, C. M. 157. Nah, S. et al. Spatially segregated free-​carrier and interlayer energy transfer. J. Am. Chem. Soc. 140,
Single-​nanowire electrically driven lasers. Nature 421, exciton populations in individual lead halide perovskite 15675–15683 (2018).
241–245 (2003). grains. Nat. Photonics 11, 285–288 (2017). 184. Tan, Z. et al. Two-​dimensional (C4H9NH3)2PbBr4
132. Röder, R. et al. Continuous wave nanowire lasing. 158. deQuilettes, D. W. et al. Impact of microstructure on perovskite crystals for high-​performance
Nano Lett. 13, 3602–3606 (2013). local carrier lifetime in perovskite solar cells. Science photodetector. J. Am. Chem. Soc. 138,
133. Mayer, B. et al. Continuous wave lasing from individual 348, 683–686 (2015). 16612–16615 (2016).
GaAs-​AlGaAs core-​shell nanowires. Appl. Phys. Lett. 159. Zhu, H. et al. Screening in crystalline liquids protects 185. Shi, E. et al. Two-​dimensional halide perovskite
108, 071107 (2016). energetic carriers in hybrid perovskites. Science 353, nanomaterials and heterostructures. Chem. Soc. Rev.
134. Sutherland, B. R., Hoogland, S., Adachi, M. M., 1409–1413 (2016). 47, 6046–6072 (2018).
Wong, C. T. O. & Sargent, E. H. Conformal organohalide 160. Miyata, K. et al. Large polarons in lead halide 186. Jin, S. et al. Scalable interconnection and integration
perovskites enable lasing on spherical resonators. perovskites. Sci. Adv. 3, e1701217 (2017). of nanowire devices without registration. Nano Lett. 4,
ACS Nano 8, 10947–10952 (2014). 161. Frost, J. M., Whalley, L. D. & Walsh, A. Slow cooling 915–919 (2004).
135. Mayer, B. et al. Lasing from individual GaAs-​AlGaAs of hot polarons in halide perovskite solar cells. 187. Whang, D., Jin, S. & Lieber, C. M. Nanolithography
core-​shell nanowires up to room temperature. Nat. ACS Energy Lett. 2, 2647–2652 (2017). using hierarchically assembled nanowire masks. Nano
Commun. 4, 2931 (2013). 162. Wang, Y. et al. Cation dynamics governed thermal Lett. 3, 951–954 (2003).
136. Saxena, D. et al. Optically pumped room-​temperature properties of lead halide perovskite nanowires. 188. Lee, W. et al. Ultralow thermal conductivity in all-​
GaAs nanowire lasers. Nat. Photonics 7, 963–968 Nano Lett. 18, 2772–2779 (2018). inorganic halide perovskites. Proc. Natl Acad. Sci. USA
(2013). 163. Zhu, H. et al. Organic cations might not be essential to 114, 8693–8697 (2017).
137. Xing, J. et al. Vapor phase synthesis of organometal the remarkable properties of band edge carriers in 189. Zhang, H. et al. 2D Ruddlesden–Popper perovskites
halide perovskite nanowires for tunable room-​ lead halide perovskites. Adv. Mater. 29, 1603072 microring laser array. Adv. Mater. 30, 1706186 (2018).
temperature nanolasers. Nano Lett. 15, 4571–4577 (2017). 190. Xing, J. et al. Color-​stable highly luminescent sky-​blue
(2015). 164. Xing, G. et al. Long-​range balanced electron- perovskite light-​emitting diodes. Nat. Commun. 9,
138. Zhang, Q. et al. Advances in small perovskite-​based and hole-​transport lengths in organic-​inorganic 3541 (2018).
lasers. Small Methods 1, 1700163 (2017). CH3NH3PbI3. Science 342, 344–347 (2013). 191. Fujita, T., Sato, Y., Kuitani, T. & Ishihara, T. Tunable
139. Evans, T. J. et al. Continuous-​wave lasing in cesium 165. Stranks, S. D. et al. Electron-​hole diffusion lengths polariton absorption of distributed feedback
lead bromide perovskite nanowires. Adv. Opt. Mater. exceeding 1 micrometer in an organometal trihalide microcavities at room temperature. Phys. Rev. B 57,
6, 1700982 (2018). perovskite absorber. Science 342, 341–344 (2013). 12428–12434 (1998).
140. Gu, Z. et al. Two-​photon pumped CH3NH3PbBr3 166. Shi, D. et al. Low trap-​state density and long carrier 192. Lanty, G., Lauret, J. S., Deleporte, E., Bouchoule, S.
perovskite microwire lasers. Adv. Opt. Mater. 4, diffusion in organolead trihalide perovskite single & Lafosse, X. UV polaritonic emission from a
472–479 (2016). crystals. Science 347, 519–522 (2015). perovskite-​based microcavity. Appl. Phys. Lett. 93,
141. Zhang, W. et al. Controlling the cavity structures of 167. Dong, Q. et al. Electron-​hole diffusion lengths 081101 (2008).
two-​photon pumped perovskite microlasers. Adv. >175 μm in solution-​grown CH3NH3PbI3 single crystals. 193. Wang, J. et al. Room temperature coherently coupled
Mater. 28, 4040–4046 (2016). Science 347, 967–970 (2015). exciton–polaritons in two-​dimensional organic–
142. Mao, W. et al. Controlled growth of monocrystalline 168. Dobrovolsky, A., Merdasa, A., Unger, E. L., Yartsev, A. inorganic perovskite. ACS Nano 12, 8382–8389
organo-​lead halide perovskite and its application in & Scheblykin, I. G. Defect-​induced local variation of (2018).
photonic devices. Angew. Chem. Int. Ed. 56, crystal phase transition temperature in metal-​halide
12486–12491 (2017). perovskites. Nat. Commun. 8, 34 (2017). Acknowledgements
143. Deng, W. et al. Ultrahigh-​responsivity photodetectors 169. Gillin, W. P., Dunstan, D. J., Homewood, K. P., S.J., Y.F. and M.P.H. are supported by the US Department of
from perovskite nanowire arrays for sequentially Howard, L. K. & Sealy, B. J. Interdiffusion in InGaAs/ Energy, Office of Science, Basic Energy Sciences, Division
tunable spectral measurement. Nano Lett. 17, GaAs quantum well structures as a function of depth. of Materials Sciences and Engineering (award no. DE-​FG02-
2482–2489 (2017). J. Appl. Phys. 73, 3782–3786 (1993). 09ER46664). X.-Y.Z. is supported by the US Department of
144. Xing, G. et al. Transcending the slow bimolecular 170. Sanvitto, D. & Kéna-​Cohen, S. The road towards Energy, Office of Science, Basic Energy Sciences (grant no.
recombination in lead-​halide perovskites for polaritonic devices. Nat. Mater. 15, 1061–1073 SE-0000234931). J.C. thanks the China Scholarship Council
electroluminescence. Nat. Commun. 8, 14558 (2016). for support.
(2017). 171. Fraser, M. D., Höfling, S. & Yamamoto, Y. Physics
145. Kim, Y.-H., Cho, H. & Lee, T.-W. Metal halide perovskite and applications of exciton–polariton lasers. Nat. Mater. Author contributions
light emitters. Proc. Natl Acad. Sci. USA 113, 15, 1049–1052 (2016). Y.F. and S.J. discussed the content of the article and Y.F.
11694–11702 (2016). 172. Zhang, S. et al. Strong exciton–photon coupling in researched data for the article. All authors contributed to the
146. Era, M., Morimoto, S., Tsutsui, T. & Saito, S. hybrid inorganic–organic perovskite micro/nanowires. writing and editing of the article prior to submission.
Organic-​inorganic heterostructure electroluminescent Adv. Opt. Mater. 6, 1701032 (2018).
device using a layered perovskite semiconductor 173. Du, W. et al. Strong exciton–photon coupling and Competing interests
(C6H5C2H4NH3)2PbI4. Appl. Phys. Lett. 65, 676–678 lasing behavior in all-​inorganic CsPbBr3 micro/ The authors declare no competing interests.
(1994). nanowire Fabry-​Pérot cavity. ACS Photonics 5,
147. Liang, D. et al. Color-​pure violet-​light-emitting diodes 2051–2059 (2018). Publisher’s note
based on layered lead halide perovskite nanoplates. 174. Shang, Q. et al. Surface plasmon enhanced strong Springer Nature remains neutral with regard to jurisdictional
ACS Nano 10, 6897–6904 (2016). exciton–photon coupling in hybrid inorganic–organic claims in published maps and institutional affiliations.

www.nature.com/natrevmats

You might also like