You are on page 1of 20

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Invited Feature Article

Cite This: Langmuir 2019, 35, 11609−11628 pubs.acs.org/Langmuir

Synthetic Evolution of Colloidal Metal Halide Perovskite


Nanocrystals
Chun Kiu Ng, Chujie Wang, and Jacek J. Jasieniak*
ARC Centre of Excellence in Exciton Science, Department of Materials Science and Engineering, Faculty of Engineering, Monash
University, Clayton, VIC 3800, Australia

ABSTRACT: Metal halide perovskite semiconductor nano-


crystals have emerged as a lucrative class of materials for many
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

optoelectronic applications. By leveraging the synthetic


toolboxes developed from decades of research into more
traditional semiconductor nanocrystals, remarkable progress
has been made across these materials in terms of their structural,
compositional, and optoelectronic control. Here, we review this
Downloaded via 45.239.80.30 on October 9, 2022 at 18:20:21 (UTC).

progress in terms of their underlying formation stages, synthetic


approaches, and postsynthetic treatment steps. This assessment
highlights the rapidly maturing nature of the perovskite
nanocrystal field, particularly with regard to their lead-based
derivatives. It further demonstrates that significant challenges
remain around precisely controlling their nucleation and growth
processes. In going forward, a deeper understanding of the role of precursors and ligands will significantly bolster the versatility
in the size, shape, composition, and functional properties of these exciting materials.

■ INTRODUCTION
The accelerated global reliance on optoelectronic devices
Cs+), B (e.g., Pb2+ and Sn2+), and X (Cl−, Br−, and I−) ions,
respectively.29 When t ≈ 1, the perovskite will possess a highly
requires the development of faster, more efficient. and cheaper symmetrical cubic phase (Figure 1g) phase, while for t > 1 and t
optoelectronic materials.1,2 Colloidal metal halide perovskite < 0.8, a hexagonal or tetragonal (Figure 1h)phase and a
nanocrystals (PNCs) have emerged as one of the most exciting rhombohedral or orthorhombic (Figure 1i) phase, respectively,
nanoscale optoelectronic material candidates.3 Since being first will be favored. This structural framework provides great
reported in 2014,4 the field of PNCs has advanced quickly by flexibility for tuning the composition of metal halide perovskites
leveraging the progress in the material design of their bulk and their resulting properties.
counterparts5,6 and synthetic developments in traditional The PNC field has closely followed the advances in
semiconductor nanocrystal (e.g., II−VI or III−V) systems.7−9 compositional control arising from bulk perovskite films. As a
This can be seen by the dramatic growth in the number of PNC result, most attention to date has been placed on lead-based
publications and the concomitant decline of those focused on candidates. These exhibit a distinct electronic structure
more traditional semiconductor nanocrystals (Figure 1a,b). As a compared to the more typical tetrahedrally coordinated
result of this activity, PNCs with reasonable synthetic control of semiconductors (e.g., CdSe or GaAs), with the lead halide
size, shape, and composition have been demonstrated in only a perovskite conduction and valence bands being dictated by
few years.10,11 This synthetic control has led to PNCs with antibonding Pb(6p) and hybridized X(p)−Pb(6s) orbitals.10
tunable absorption and photoluminescence (PL) across the This entails that they have rather benign defect states, which is a
entire visible spectrum,12,13 photoluminescence quantum yields key reason for their exceptional optoelectronic properties.
(PLQY) approaching unity,14,15 short radiative lifetimes of ∼10 Moreover, it indicates that the modification of the X and B sites
ns,16,17 narrow PL full-width at half maxima (fwhm) of 12 to 45 provides the most suitable strategies for modifying the band gap.
nm,4,18 and suppressed PL blinking statistics.19 These appealing Of these, the modification of the halide composition has
characteristics have further led to PNCs being successfully emerged as the most facile for tuning the optical properties of
harnessed within numerous optoelectronic applications, includ- PNCs across the visible spectrum,16 with the larger halides (i.e.,
ing light-emitting diodes (LEDs),20,21 lasers,22,23 photodetec- rCl− = 1.81 Å, rBr− = 1.96 Å, rI− = 2.20 Å) yielding higher band
tors,24,25 amplified spontaneous emitters,26 and photovoltaics gaps owing to their energetically deeper bonding p orbitals.10,31
(Figure 1c−f).27,28 In contrast, A-site modification has exhibited limited band gap
Metal halide perovskites have the generalized chemical tunability but has opened up hybrid and inorganic perovskite
formula of ABX3, where A and B are cations and X is a halide
anion. Their precise crystal phase can be gauged by the Received: April 24, 2019
Goldschmidt tolerance factor, t = (rA + rX)/[(√2)rB + (√2)rX], Revised: June 2, 2019
where rA, rB, and rX are the radii for the A (e.g., MA+, FA+, and Published: June 4, 2019

© 2019 American Chemical Society 11609 DOI: 10.1021/acs.langmuir.9b00855


Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

Figure 1. (a) Accumulative and (b) per annum bibliometrics of various semiconductor nanocrystal/colloidal/quantum dot systems from 1992 to 2018
(from the Web of Science database). (c) Scanning electron microscopy (SEM) cross section of a CsPbBr3 photovoltaic device, adapted with
permission from ref 28. Copyright 2016 Springer Nature. (d) Schematic of a colloidal CsPbBr3 green-emitting LED and (e) photographs of their 2 × 2
mm2 active area operating at 5 V. Reproduced from ref 20. Copyright 2018 American Chemical Society. (f) Ultrastable amplified spontaneous emission
of a CsPbBr3 PNC film with improved passivation. Reproduced from ref 26. Copyright 2015 American Chemical Society. Unit cell of general (g) cubic,
(h) tetragonal, and (i) orthorhombic ABX3 structures where A+ cations are green spheres and purple BX6 polyhedral comprise steel blue B2+ and purple
X− spheres. Reproduced with permission from ref 30. Copyright 2015 American Physical Society.

classes, which respectively use organic (e.g., MA+ or FA+) and the functional phase can be metastabilized under ambient
inorganic (e.g., Cs+) cations.12,32 Both of these classes exhibit conditions through a reduction in grain size and appropriate
promising optoelectronic properties; however, the all-inorganic surface passivation to modify its surface energy.44,45 Access to
perovskites have tended to exhibit improved thermal, chemical, such metastable structural regimes provides an important
and photostabilities.33−36 These appealing properties have led to avenue for expanding the structure−property landscape of
inorganic metal halide PNCs being the most studied to date, perovskites.
with CsPbX3 emerging as the archetypal system. These developments in the composition and phase engineer-
B-site modification has been motivated by its potential for ing of colloidal PNCs have been underpinned by synthetic
optoelectronic modification and the removal of lead, which is developments across both injection and noninjection techni-
inherently toxic.37 Lead’s closest analogue, Sn2+, is readily used ques.4,16,46 Injection methods involve the mixing of precursor
as a substituent in bulk perovskite counterparts to reduce the solutions to induce supersaturation and drive the formation of
band gap but has proven difficult to integrate within PNCs PNCs. Meanwhile, noninjection methods involve the controlled
because of its tendency to readily oxidize to Sn4+ within reactivity of precursors during a thermal heating profile to
nanocrystalline systems.38 As a result, the more redox-stable Bi3+ control the nucleation and growth dynamics. Across each of
and Sb3+ cations, which are also notably isoelectric with Pb2+ and these synthetic routes, the rapid nucleation and growth
Sn2+, respectively, have emerged as promising candidates for processes of PNCs have made their synthetic windows
achieving Pb-free PNCs.37 In addition to these direct band-edge extremely sensitive to factors such as the reaction temper-
recombination materials, new optoelectronic functionality has ature,47,48 reaction time,49 solvent properties,50,51 acid−base
been further introduced through B-site doping. The dual equilibria,52−54 reagent species,55−57 ligand species17,58 and
emission observed for Mn/CsPbCl339,40 and sensitized mixing rates.59 Studies have shown that adjustments to even one
lanthanide emission for La/CsPbX3 are the most common of these parameters can modify the structural properties of the
examples of these.41 resulting PNCs among quantum dots (QD), nanocube (NCu),
Beyond their enhanced optoelectronic properties, PNCs also nanoplatelet (NPL), nanosheet (NS), nanowire (NW), and
provide greater structural flexibility compared to their bulk nanorod (NR) morphologies.51,60,61 While this demonstrates
counterparts.16 This is most clearly demonstrated for CsPbI3, for versatility, it also highlights the challenges in achieving structural
which the bulk nonperovskite polymorph is a nonfunctional control within these emerging materials.
(nonluminescent, large band gap) orthorhombic γ-phase under The progress within the field of colloidal perovskite
ambient conditions42 and transforms into its desired functional nanocrystals has been rapid but is still maturing, with many
cubic α-phase above ∼310 °C.43 However, in its colloidal form, imminent research questions. In this timely review, we provide a
11610 DOI: 10.1021/acs.langmuir.9b00855
Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

detailed synopsis of the developments that have been made


toward synthesizing and postprocessing colloidal perovskites. As
a way of focus, here we do not go into detail regarding the
photophysics and applications of such materials but consider
only nontemplated synthetic methods because these have been
shown to be more suitable for the synthesis of homogeneous
PNC dispersions.

■ NANOCRYSTAL FORMATION
Nucleation. According to the classical LaMer model, the
formation of colloids involves the conversion of precursors to
reactive monomers that can form metastable clusters, some of
which ultimately exceed the critical radius and grow to form the
final nanocrystal.62 The conversion rates of the precursors
govern the dynamics of the system achieving supersaturation,
which are dictated by the specific precursor chemistry. Generally
speaking, the nucleation and growth events of PNCs can be
distilled down to a double-displacement or salt metathesis
reaction given by eq 1.
A+ + B2 + + 3X− → ABX3 (1)
For low-temperature PNC injection reactions (<90 °C), a Figure 2. (a) Evolution of CsPbI3 absorption spectra between 0.4 to 2.4
supersaturated state is achieved through solvent control, in s from 580 to 670 nm, corresponding to PNC sizes of 8−12.5 nm and
which the addition of an antisolvent drastically reduces the (b) the estimated particle size found over a range of different reaction
solubility of the precursors down to 10−7 g/mL.63 Within an temperatures. Reproduced from ref 18. Copyright 2016 American
appropriate synthetic window, PNCs precipitate and are Chemical Society. (c) TEM micrographs of CsPbI3 PNC as a function
concurrently stabilized by ligands during mixing, commonly in of reaction time and (d) concurrent evolution of the PL spectrum as it
the form of oleic acid (OA) and oleylamine (OLA) and their changes with the size distribution of the PNCs, as reproduced with
conjugates oleate (OA−) and oleylammonium (OLA+), permission from ref 74. Copyright 2016 Royal Society of Chemistry.
respectively.59 The stoichiometric reaction for a typical low-
temperature reaction is described by eq 2. involves heterogeneous seeding by Pb0 nanoparticles within
the reaction mixture. This is supported by the observation of
AX + BX 2 + OA + OLA → ABX3 + OA− + OLA+ (2) such seeds within Pb2+ precursor solutions prior to PNC
Meanwhile, the hot-injection (HI) method for PNCs uses nucleation50 and Pb-rich67 or even Pb0 50 “dark particles” within
reaction temperatures typically ranging from 90 to 240 °C. It the PNCs synthesized using low-temperature injection,32,52,58
harnesses a rapid injection process to effectively “decouple” the high-temperature injection,60,68,69 and multiple noninjection
nucleation and growth stages to result in more homogeneous methods.55,70,71 The existence of Pb0 seeds is consistent with
colloidal nanocrystal dispersions.64 Its success relies on the Pb2+ being easily reduced (E0 = −0.13 V)72 and the fact that
nucleation and growth dynamics being slower than the effective most synthetic methods use aliphatic amines for stabilization,
mixing times of the precursor solution. For PNCs, this has which are also mild reducing agents. However, further research
proven to be difficult because nucleation has been exper- is required to unambiguously validate this mechanism of
imentally verified to occur within 0.4 s (Figure 2a).18 Recent nucleation, particularly because PNCs have also been observed
theoretical works have further suggested that the nucleation to degrade under an electron beam during characterization,
processes themselves occur on submicrosecond time scales.65 yielding similar structural characteristics.73
These fast kinetics have made for detailed experimental Growth. A key synthetic challenge of PNCs is that their
investigations into the early formation stages of perovskites nucleation and growth stages are difficult to decouple. The
limited to date. Despite this, for a typical PNC HI synthesis, the depletion of available free monomers during growth is complete
generalized conversion of precursors into their reactive species within ∼5 s of reaction time for HI (Figure 2b).18 According to
follows eqs 3 and 4, with the overall formation reaction given by the classical LaMer model, under these reaction conditions the
eq 5. predominant growth mode is expected to rapidly transition from
that involving free monomer consumption on the very early time
Cs 2CO3 + 2OA → 2CsOA + H 2O + CO2 (3) scales to Ostwald ripening. Studies on CsPbI3 PNCs have shown
that size focusing of nucleated NCu occurred within the first few
PbX 2 + 2OA + 2OLA → PbOA 2 + 2OLAX (4) seconds of the reaction (Figure 2c,d).74 This was followed by
defocusing of the colloidal ensemble through diffusion-driven
CsOA + PbOA 2 + 3OLAX → CsPbX3 + 3OA− + 3OLA+ growth, which is consistent with the Ostwald ripening process.
(5) Similarly, the growth stage of PNCs in low-temperature
Across these reaction mechanisms, the intricacies of the injection syntheses has been estimated to be mostly complete
underlying nucleation mechanism remain unclear. One model of after 10−20 s.28 The lowered reaction kinetics compared to that
PNC formation involves the initial formation of corner-sharing of HI methods likely stems from the decreased monomer
PbX6 polyhedra as nuclei, which then have their interstitial voids diffusion rates and reactivities at lower temperatures. As-
gradually filled with A+ ions,66 akin to intramolecular exchange synthesized PNC dispersions from low-temperature synthetic
processes in bulk perovskite.34 An alternative mechanism methods can contain 10−30% by volume polar solvent, such as
11611 DOI: 10.1021/acs.langmuir.9b00855
Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

Figure 3. UV−vis absorption and PL spectra of CsPbI3 PNCs synthesized with (a, b) OA or (c, d) TMPPA paired with OLA, with the latter exhibiting
enhanced stability under ambient conditions. Reproduced with permission from ref 45. Copyright 2017 Royal Society of Chemistry. (e) DFT
modeling of oleylammonium ion pairs and sulfobetaine interactions with CsPbBr3 surface (Cs, gray; Pb, orange; Br, magenta). Reproduced from ref 81
Copyright 2018 American Chemical Society.

dimethylformamide (DMF), dimethyl sulfoxide (DMSO), or acid, have yielded higher PLQYs (>80%) than their shorter-
isopropanol (IPA).28,63 Owing to the solubility of the perovskite chained derivatives.58
lattice in such polar solvents, Ostwald ripening is enhanced Phosphorus-containing ligands have also been used in PNC
compared to syntheses where only nonpolar solvents are used.75 synthesis. Of these, trioctylphosphine (TOP) has been the most
It has been suggested that the presence of these solvent mixtures common because it forms highly soluble complexes with PbX2.16
makes for PNC formation subject to classic micellar formation Its use in mixed-ligand reactions involving OA and OLA has
theory, where the electrostatic and hydrophobic interactions enabled the synthesis of high-quality CsPbI3 quantum dots, with
dictated by the specific precursors, solvents, and ligands PLQYs approaching 100%.14 Importantly, nuclear magnetic
influence the final PNC morphology.76 resonance (NMR) spectroscopy measurements have revealed its
Following the initial, rapid growth stage of nanocrystal absence from the synthesized PNC surfaces, with only OLA+
formation, orientated attachment is the other noteworthy and OA− species being detected. A similar observation was
growth mechanism that is prevalent in PNC systems.77 This observed when CsPbI3 PNCs were synthesized with OA
growth mechanism is believed to occur when the extent of substituted with branched bis(2,4,4-trimethylpentyl)phosphinic
surface ligand passivation is poor or is destabilized to allow for acid (TMPPA) in the HI synthesis, only this time OLA+ was
coalescence. It is considered to be the dominant mechanism for observed as the sole ligand on the surface.45 Despite the
the formation of larger PNC structures, such as NWs, NPLs, and phosphinic acid not being an active surface binding species for
NSs.50,78 the final CsPbI3 PNC, the resulting OLA+ surface ligand shell
Ligand Interactions with Perovskite Nanocrystals. allowed for the successful metastabilization of the functional α-
Ligand species used for PNCs have often been adopted from phase under ambient conditions (Figure 3a−d). It remains
more traditional semiconductor colloids, with primary aliphatic unclear as to whether the absence of the phosphorus-containing
amines and carboxylic acid species being the most favored.48,57 ligands in the above NMR studies arises from weaker chemical
Surface investigations of synthesized PNCs using these ligand interactions at the PNC surface and/or from steric effects from
combinations have unambiguously confirmed that the aliphatic the branched ligand chains. The use of trioctylphosphine oxide
ammonium and carboxylate species are the dominant surface (TOPO) has similarly improved size uniformity and stability of
binding ligands.79,80 These form in solution when ligand pairs PNCs, although its presence on the surface confirms that it plays
undergo simple acid−base reactions, such OLA and OA reacting a passivating role rather than merely acting as a sacrificial
to form oleylammonium−oleate (OLA+−OA−).53 Density additive or spectator ligand.69,71
functional theory (DFT) modeling has further indicated high Investigations into less common ligands have included
ligand-to-surface interaction energies of these reacted species, nonprimary amines (i.e., secondary to quaternary amine
with that for OLA+ and OA− ranging from −1.73 to −1.95 eV for substitutes),58,85 bidentate ammonium species,51,85 aniline
OA−81 and to −1.14 eV for OLA+.15 In comparison, the binding derivatives,86 and zwitterions.81 Generally, these have yielded
energy of neutral Cs-oleate and Pb-oleate is approximately half lower-quality PNCs compared to the more traditional ligand
that of the oleate or an alkylammonium-alkanoic acid pair, thus combinations because of factors that include poorer steric
demonstrating that the charged species are more favorable stabilization, modified precursor/monomer solubilities, and
surface passivants.79 more prominent Ostwald ripening.51 However, there are a few
Investigations into alkylamine−alkanoic ligand pairs of exceptions, with the most notable being by Krieg et al.,81 who
different chain lengths have further shown that lengthier ligand used a long-chained sulfobetaine zwitterion as the sole ligand
species (octyl- or higher) yield better-stabilized and higher- (Figure 3e), which yielded improved chemical durability and
quality colloidal dispersions.82−84 This stems from the improved storage lifetimes compared to the standard OA and OLA ligand
steric stabilization provided by the longer-chained ligands,75 pair. This was attributed to the sulfobetaine molecule being
which also inhibit monomer uptake during PNC growth to unaffected by Bronsted acid−base equilibria and, because of a
typically result in smaller sizes and NCu formation.57,61 As a kinetically stabilized chelation mechanism, had a PNC surface
result, longer-chain-length acids, such as OA and tetradecanoic binding energy similar to that of OLA+.
11612 DOI: 10.1021/acs.langmuir.9b00855
Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

Table 1. Summary of Major Classes of Injection and Noninjection Synthetic Methods for Perovskite Nanocrystals

Table 2. Summary of the Major Synthetic Procedures for Perovskite Nanocrystalsa


method composition shape size (nm) PLQY reaction conditions ref
NT MAPbX3 NP 6 17% ambient, 80 °C 4
NT CsPbBr3 NCu, NPL 8.5, 41 × 8 31−78% ambient, RT 52
LARP MAPbX3 QD ∼3.3 50−70% ambient/N2, RT 88
LARP MA3Bi2X9 QD ∼3 4−12% ambient/N2, RT 89
LARP CsPbX3 QD 11 70−95% ambient, 0−30 °C 63
LARP Cs3Sb2X9 QD ∼3 20−46% ambient, RT 92
HI CsPbX3 NCu 4−15 50−90% N2, 140−200 °C 16
HI Mn:CsPbCl3 NCu 11 <27% N2, 185 °C 40
HI Lanthanide: CsPbCl3 NCu 6.3−7.6 28% N2, 160 °C 41
HI CsSnX3 NCu ∼10 0.14% N2, 170 °C 38
HI Cs3Sb2X9 NR, NW 20 × ≤3400 4% N2, 170−200 °C 106
HI Cs3AgBiCl6 NCu ∼8 ∼0.3% N2, 140 °C 110
convection CsPbX3 NCu 4.0−15 87% N2, 50−170 °C 46
solvothermal CsPbX3 NCu, NW 8.2−12.5 80% ambient, 160 °C 55
microwave CsPbX3 NCu 10−13 11−92% ambient, 800 W 115
ultrasonication CsPbX3 NP, NPL 8−15 10−92% ambient, RT 70
a
QD, quantum dot; NP, nanoparticle; NCu−nanocube; NPL, nanoplatelet; NS, nanosheet; NR, nanorod; NW, nanowire; RT, room temperature
∼25 °C; PLQY, photoluminescence quantum yield; N2, nitrogen gas; W, watts

■ SYNTHETIC METHODS
In this section, we will discuss the advances across the various synthesis
we have divided these sections to reflect the key advances across both
lead-based and lead-free perovskite classes.
Anti-Solvent-Assisted Low-Temperature Injection. Lead-
approaches that have been developed within both injection and Based Perovskites. The first reported low-temperature injection
noninjection approaches. A brief description and a summary of the synthesis of any PNC was by Schmidt et al., who prepared hybrid
advantages and limitations of each of these is presented in Table 1. MAPbBr3 nanoparticles.4 This synthesis was subsequently dubbed the
Furthermore, in Table 2, we provide an overview of the major synthetic non-template (NT) method. In their synthesis, a DMF solution that
procedures used across these methods. By far, the most investigated contained solubilized precursor salts (MABr, PbBr2) was mixed with a
synthetic approaches for PNCs are the injection methods. As such, here ligand solution containing OA and octylammonium bromide in 1-

11613 DOI: 10.1021/acs.langmuir.9b00855


Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

octadecene (ODE) at 80 °C. This was followed by the addition of an


antisolvent (acetone) to reduce the solubility of the salts and precipitate
the PNCs. The synthesized MAPbBr3 nanoparticles were found to have
an average size of 6.2 nm, with a typical emission at ∼525 nm and a
PLQY of 17%. Fine tuning of the molar ratios between the ammonium
species and PbBr2 to 4:1 was shown to yield ∼5.5 nm PNCs, with a
comparable emission wavelength but a significantly enhanced PLQYs
of up to 83%.87 This enhancement in PLQY was considered to arise
from ammonium cations binding to undercoordinated bromide ions.
The NT method has also been applied to all-inorganic CsPbX3 PNCs
by incorporating a Cs-oleate precursor solution, analogous to that used
in HI syntheses.52 Driven by the search for a more scalable and robust
processing method, Akkerman et al. explored the use of low-boiling-
point ligands and solvents within the NT method.28 By employing
propionic acid and butylamine as ligands and IPA and n-hexane as
solvents, the authors were able to scale the synthesis of crude CsPbBr3
PNCs to volumes of as high as 900 mL. The sizes of NT-synthesized
QDs could be controlled between 2.6 and 7.2 nm by simply changing
the relative polar and nonpolar solvent volumes.32 The PLQYs were
found to range between 80 and 92%.
As an analogue of the NT method, Zhang et al. introduced the
ligand-assisted re-precipitation (LARP) approach to synthesize
MAPbX3 QDs at room temperature.88 In this synthesis, all precursor
materials (MABr and PbBr2) and ligands (OLA and OA) were
dissolved in a polar solvent (DMF or DMSO). This solution was added
to a toluene antisolvent to drive the precipitation of 3.3 nm MAPbBr3
PNCs, which exhibited PLQYs of up to 70%. Through reaction
temperature control between 0 and 60 °C, the sizes of the MAPbBr3
PNCs were found to be tunable in the range of 1.8 and 3.6 nm, with
lower temperatures yielding lower particle sizes.47 Quantum confine-
ment effects could be observed at the smaller particles sizes, with the
enhanced exciton binding energies permitting enhanced PLQYs of as
high as 93%.
The synthesis of all-inorganic CsPbX3 PNCs has also been explored Figure 4. (a) Photograph of CsPbX3 (X = Cl, Br, I) PNC solutions
using LARP through the substitution of the A-site organic halide salt under UV illumination and (b) their associated PL spectra. Reproduced
with a CsX salt analogue.32,58,63 Such studies have demonstrated PNCs with permission from ref 58. Copyright 2016 Royal Society of
with controllable NCu sizes from 9 to 20 nm that exhibited PLQYs of as Chemistry. (c) Different chain length acid−base ligand combinations
high as 95% and narrow PL fwhm’s of 12 to 38 nm.32,63 Tuning the influencing the CsPbX3 morphology. Reproduced from ref 17.
halide composition between CsPbCl3 and CsPbI3 has further allowed Copyright 2016 American Chemical Society.
for spectral control across the 405 to 680 nm visible region (Figure
4a,b).58,63 Notably, the synthesis of functional cubic CsPbI3 has been
challenging using CsX salts because of the relatively large synthesized tune the emission between 370 and 560 nm. These all-inorganic PNCs
particle sizes, resulting in phase instability. It has been shown that exhibited enhanced thermal and chemical stabilities compared to their
through the use of Cs-oleate precursors, smaller ∼5 nm CsPbI3 QDs hybrid counterparts.90,91
could be prepared, thus enabling the cubic phase to be metastabilized.17 In addition to Cs+-containing Pb-free all-inorganic perovskites, the
The same authors showed that CsPbX3 PNCs could be synthesized synthesis of Rb+-containing zero-dimensional Rb3Bi3Cl16 has also
with a range of morphologies (QD, NCu, NPL, and NR) through recently been reported.93 The authors dissolved RbCl and BiCl3 in γ-
control of the acid−amine ligand chain lengths (Figure 4c). butyrolactone with small amounts of OA and octylamine as ligands.
Lead-Free Perovskites. Leng et al. was the first to report MA3Bi2Br9 This precursor solution was held at 80 °C for an hour before being
QDs synthesized through a modified LARP approach.89 The authors injected into ethanol. The resultant Rb3Bi3Cl16 PNCs displayed a PL
dissolved MABr and BiBr3 in DMF and ethyl acetate, using n- peak centered at 437 nm with a fwhm of 93 nm and a PLQY of 28%.
octylamine as a ligand, and then injected this solution into octane Their small size of ∼1.85 nm was attributed to the smaller atomic radii
containing OA at 80 °C. This resulted in ∼3 nm PNCs, which possessed of Rb+ and Bi3+, as well as the lower solubility of RbCl compared to that
a modest PLQYs of up to 12%. Postsynthetic halide exchange was able of CsCl, which would result in a more supersaturated solution and
to access MA3Bi2X9 (X = Cl, Br, I) compositions and tune the PL peaks therefore more rapid nucleation. Importantly, for their prospective use
from 360 to 540 nm. in real-world applications, these particles were found to retain 70% of
All-inorganic Cs3Bi2X9 (X = Cl, Br, I) has been synthesized by their fluorescence after a month of ambient storage and had a thermal
injecting a DMSO solution containing CsX and BiX3 precursors into stability of up to 400 °C.
IPA. The nanocrystals showed an emission tunability range of 400 to Hot Injection. Lead-Based Perovskites. The first report outlining
560 nm (Figure 5a−c).90 Without any aliphatic reaction ligands the HI synthesis of inorganic CsPbX3 PNCs was by Protesescu et al.16
present, ∼6 nm trigonal Cs3Bi2Br9 QDs were formed under these The authors reacted Cs2CO3 with OA in ODE under an inert gas (N2)
synthetic conditions, albeit with a low PQLY of 0.2%. This was environment at 120 °C to make a Cs-oleate precursor (eq 3).
enhanced to 4.5% when OA was introduced into the reaction. By using Meanwhile, PbX2 was separately dissolved in ODE with the assistance
octylamine and OA in the reaction and ethanol as the nonsolvent, of OLA and OA (eq 4). The Cs-oleate was injected into the PbX2
reactions carried out at 80 °C yielded 3.9 nm Cs3Bi2Br9 QDs with an solution at a temperature of 140 to 200 °C to form colloidal CsPbX3
improved PLQY of up to 19.4%.91 Using very similar synthetic PNCs (eq 5). Mixed-halide PNCs could be readily achieved by
approaches, antimony-substituted Cs3Sb2Br9 QDs have also been combining different PbX2 salts. Generally, Cs/Pb ratios of 1:2 to 1:4
successfully demonstrated.92 The 3.1 nm trigonal Cs3Sb2Br9 exhibited have yielded the most monodisperse PNCs.18 To preserve the narrow
an emission peak at 410 nm and a very promising PLQY of 46% (Figure size distribution of the PNCs, the reaction was terminated after 5 s by
5d−f). Postsynthetic halide exchange could be further performed to quenching in an ice bath. The resulting NCu’s were 4 to 15 nm in size,

11614 DOI: 10.1021/acs.langmuir.9b00855


Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

Figure 5. (a) Photographs and (b) XRD and (c) UV−vis absorption and PL spectra of Cs3Bi2X9 (X = Cl, Cl0.5Br0.5, Br, Br0.5I0.5, I) PNCs. Reproduced
with permission from ref 90. Copyright 2016 John Wiley and Sons. (d) UV−vis absorption and PL spectra of Cs3Sb3Br9 with an inset of its colloidal
solution under UV light. (e) Diagram of its atomic structure and (f) HR-TEM of ∼3 nm Cs3Sb3Br9 QDs. Reproduced from ref 92. Copyright 2017
American Chemical Society.

Figure 6. Colloidal dispersions of all-inorganic CsPbX3 perovskite nanocrystals (a) under UV light (λ = 365 nm), (b) their photoluminescent emission
profiles, (c) and UV−vis absorption spectra. Reproduced from ref 16. Copyright 2015 American Chemical Society.

with lower temperatures yielding the smaller particle sizes. Through length dependence, with the thinnest NPLs of 1.8 nm in thickness being
temperature and halide composition engineering, the resulting CsPbX3 grown using hexylamine. Alternative metal precursors, including lead
PNCs exhibited tunable absorption and PL across the entire visible oxide,95 lead acetate,68 cesium acetate,57 and cesium hydroxide,58 have
spectrum, with PLQYs of 50 to 90% and narrow PL fwhm of 12 to 42 also been explored as a way to gain finer control of the reaction
nm (Figure 6). conditions.56,83 In these syntheses, the use of ammonium halides or
Further investigations into the effects of reaction temperature on the benign metal halides as halide sources allowed for better stoichiometric
HI synthesis of CsPbX3 showed that lower reaction temperatures of 90 control over the reaction mixture.96 Typically, X/Pb ratios ranging from
to 130 °C strongly favored asymmetric growth through oriented 2:1 to 4:1 have been shown to yield the highest-quality PNCs, with
attachment, resulting in 2D NPLs or NSs (Figure 7a).78,94 The PLQYs improved PLQYs, narrower fwhm’s, and higher PL lifetimes.95 The
of the NPLs ranged between 10 and 84%, with the thinner NPLs having optoelectronic improvements under these halide-rich syntheses were
lower values likely due to enhanced surface trap contributions. considered to arise from reduced surface halide vacancies on the PNCs,
Lowering the reaction temperature to 70 °C resulted in a transparent which can act as traps.
suspension of amorphous, micrometer-sized sheets composed of Tuning the halide composition readily modifies the spectral nature of
unreacted precursors. Meanwhile, within the standard 150 to 200 °C PNCs: A- and B-site engineering can provide extended opportunities to
synthetic window, extended reaction times of more than 10 min were modify their structure−property relations. Protesescu has been one of a
shown to transform the NCu PNCs, initially into NWs through an few authors to demonstrate A-site modification using HI, incorporating
oriented attachment mechanism and then into coarsened micrometer- FA into CsPbI3 PNCs by simply injecting the FA-oleate precursor
sized crystals at prolonged reaction times of >180 min (Figure 7b).49 together with Cs-oleate during synthesis.12 Using a molar ratio of 2:1
Beyond the reaction temperature and time, ligand and precursor FA/Cs in the precursor solution, it was found that only 10% of the FA
engineering are the other major handles used to modify the resulting was incorporated into the PNC structure. The synthesized
structural properties of PNCs. Pan et al. investigated a variety of FA0.1Cs0.9PbI3 PNCs exhibited PL emission at 685 nm with a PLQY
aliphatic carboxylate and amine ligand pairs, ranging from acetic to oleic of >70%, and their cubic phase was maintained for months under
acid and hexyl- to oleylamine, respectively, generally observing an ambient conditions (Figure 8a). The latter is consistent with CsPbI3
increased NCu size for decreasing acid chain lengths across these.57 PNCs alloyed with a larger secondary A-site cation (e.g., FA+) being
Meanwhile, using shorter amines than OLA resulted in PNCs adopting more stable in a cubic phase because of their Goldschmidt tolerance
NPL morphologies. The NPL thickness exhibited an amine chain- factor being closer to 1.97,98

11615 DOI: 10.1021/acs.langmuir.9b00855


Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

Figure 7. Conditions that can influence the morphology of HI-synthesized CsPbBr3 PNCs with inset photographs of solutions under UV light. (a)
Reaction temperature showing NPLs forming at lower temperatures as opposed to NCu at 150 °C, scale bar 50 nm. Reproduced from ref 78. Copyright
2015 American Chemical Society. (b) Reaction times of >180 min showing initial NCu transitioning into NWs and NSs with a scale bar of 100 nm.
Reproduced from ref 49. Copyright 2015 American Chemical Society.

Rubidium has also been shown to be an alternative to cesium for all- specific lanthanide to be observed. By changing the lanthanide species,
inorganic PNCs. This was first demonstrated with the synthesis of tunable photoluminescence ranging from 500 nm to the near-infrared
orthorhombic RbPbI3 NWs by following the standard HI procedure has been observed under laser illumination, with the overall PLQY
and replacing Cs2CO3 with Rb2CO3.99 A reaction time of 30 min at 150 being around 15 to 30% (Figure 8c). Importantly, Zhou et al. reported
°C allowed the initially formed NRs to grow into NWs with a diameter that cerium and ytterbium codoped CsPbCl1.5Br1.5 PNCs possessed a
of ∼32 nm and lengths of <100 μm. More extensive studies have found QY of 146% at 988 nm.103 This demonstration of quantum cutting
that the low tolerance factor of ∼0.8 for RbPbX3 (X = Br, Cl) is close to further highlights that lanthanide-doped PNCs provide a scope of
the lower formability range of perovskites and is not readily application as security and biological markers, high-efficiency photo-
synthesized.100 This problem could be alleviated by using mixed A- voltaics, and optical sensors.
site RbxCs1−xPbX3 (X = Br, Cl) compositions, yielding PLQYs Lead-Free Perovskites. Efforts toward replacing Pb with Sn within
comparable to CsPbX3 PNCs. HI synthesis have shown that by simply substituting PbX2 with SnX2,
B-site cation doping has been shown to be more compatible with lead-free CsSnX3 PNCs with an ∼10 nm size can be readily produced.38
current HI protocols. Mn2+ has emerged as one of the more successful These RT orthorhombic (X = Br, I) or cubic (X = Cl) CsSnX3 NCu’s
dopants because it has the same valence as the native Pb2+. Liu et al. first exhibited a perovskite structure with a tunable band gap based on their
reported that by simply dissolving MnX2 precursor simultaneously with halide composition (Figure 9a−c). However, because of the distortion
PbX2 within a typical PNC HI protocol, Mn/CsPbX3 PNCs could be of the halide sites and higher defect densities compared to those of their
produced.40 At higher MnX2 precursor concentrations, the Mn2+ lead analogues, they exhibited a maximum PLQY of only 0.14%.
dopant amount in the PNCs would increase, achieving a saturated Interestingly, by using SnI4 as a precursor within the same HI
substitution ratio of 46% at a Mn/Pb reaction ratio of 10:1 (Figure procedure, more air-stable Cs2SnI6 PNCs resulted, although the exact
8d,e).39 As a result of the doping process, these PNCs exhibited dual mechanism for this remains unclear.104 By extending the reaction time
emission properties, with a higher-energy intrinsic perovskite emission from 1 min to 1 h, the initial Cs2SnI6 PNCs were observed to transition
competing with a lower-energy radiative recombination process to NRs and NWs before adopting NPL morphologies. Their highest
between the 4T1 and 6A1 states of Mn2+ at ∼580 nm.101 Mn2+ doping PLQY remained low at 0.48%, indicating a significant scope for further
has been shown to further improve the thermal and ambient structural refinement.
atmospheric stabilities of CsPbBr3 PNCs.102 Bismuth substitution has been another common approach to
Pan et al. has demonstrated that lanthanide ions (Ce3+, Sm3+, Eu3+, achieving lead-free PNCs. By using BiX3 as the precursor in the HI
Tb3+, Dy3+, Er3+, and Yb3+) can also be introduced into the CsPbCl3 protocol, ∼18 nm hexagonally shaped Cs3Bi2X9 PNCs were produced,
PNCs by simply adding lanthanide chloride salts to the Pb precursor which also possessed a hexagonal crystalline phase (Figure 9e).105
solution during the HI process.41 Because of the 3+ valence of the Owing to this large size and different phase, the PLQY of these Cs3Bi2X9
lanthanide ions and their significantly smaller ionic radius versus that of PNCs was very low (0.017%) compared to that of those synthesized
lead, the doped CsPbCl3 exhibited a distorted perovskite structure with using low-temperature methods. Notably, the hexagonal Cs3Bi2X9
a smaller lattice constant (e.g., 3.96 Å for undoped CsPbCl3 and 3.87 Å PNCs possessed an interesting direct−indirect band structure, which
for Yb/CsPbCl3, Figure 8b). In these systems, the perovskite acts as a is composed of a direct exciton transition and also a phonon-assisted
sensitizer for the dopant, which enables secondary emission from the indirect transition. Such a unique band structure endows the Cs3Bi2X9

11616 DOI: 10.1021/acs.langmuir.9b00855


Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

Figure 8. (a) Optical absorption and PL spectra of FA0.1Cs0.9PbI3 and FAPbI3 PNCs freshly synthesized and 6 months later. Insets show photographs
of their respective solutions under sunlight and λ = 365 nm UV light. Reproduced from ref 12. Copyright 2017 American Chemical Society. (b) XRD
and (c) PL characterization of lanthanide-alloyed CsPbCl3 PNCs. Reproduced from ref 41. Copyright 2017 American Chemical Society. (d) TEM
micrographs of CsPbxMn1−xCl3 NCu prepared at 170 °C with various Pb/Mn molar ratios in its reaction mixture and (e) the corresponding PL spectra
with emission peaks at ∼410 nm and ∼580 nm (higher Pb/Mn ratios result in higher Mn levels of doping determined via EDX). Reproduced from ref
39. Copyright 2017 American Chemical Society.

Figure 9. (a) XRD and (b) UV−vis absorption spectra of CsSnX3 (X = Cl, Cl0.5Br0.5, Br, Br0.5I0.5, I) PNCs. (c) TEM micrograph of CsSnI3 nanocrystals
(scale 100 nm) with the inset showing NCu ≈ 9.9 nm in size. Reproduced from ref 38. Copyright 2016 American Chemical Society. (d) UV−vis
absorption and PL spectra of Cs3Bi2I9 PNCs showing direct (580 nm) and indirect (605 nm) emission. (e) TEM micrographs showing particle sizes of
∼18 nm. Reproduced from ref 105. Copyright 2017 American Chemical Society.

PNCs with a dual peak emission, which is centered at 580 and 605 nm and NWs with diameters of 20 nm have been synthesized with tunable
for X = I (Figure 9f). This dual emission was not observed in LARP- lengths of up to 290 nm and 3.4 μm, respectively.106 This size tunability
synthesized Cs3Bi2X9 PNCs, which may be due to their different was achieved through controlling the amount of the SbCl3 precursor
morphology and/or crystalline phase. and OA/hexadecylammonium ligand ratios in their reactions. The
To a lesser extent, antimony-substituted perovskites have also been emission peak maximum of NW samples at 436 nm exhibited a PLQY
investigated. By using antimony chloride (SbCl3) within HI protocols of 4%. Cs3Sb2I9 NPLs and NRs have been similarly synthesized using
comparable to those mentioned above, mostly trigonal Cs3Sb2Cl9 NRs reaction temperatures of 180 and 230 °C, respectively.107 The NPLs

11617 DOI: 10.1021/acs.langmuir.9b00855


Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

Figure 10. (a−d) TEM micrographs of convection-synthesized CsPbI3 PNCs with sizes of 3.3, 4.5, 6.5, and 10.4 nm. (e) PL spectra for different halide
compositions. (f) Photographs of CsPbBr3 and CsPbI3 PNC solutions. (g) Absorption and PL spectra of CsPbI3 PNCs grown at 50, 80, and 110 °C,
adapted with permission from ref 46. Copyright 2016 Springer Nature.

Figure 11. TEM micrographs of solvothermally synthesized PNCs with different morphologies. (a) CsPbBr3 NCu, inset: HR-TEM revealing 0.58 nm
lattice spacing; (b) CsPbBr3; (c) CsPb(Br/I)3; (d) CsPb(Cl/Br)3 NWs; and (e) their corresponding absorption and PL spectra. Reproduced with
permission from ref 55. Copyright 2017 John Wiley and Sons. (f) CsPbBr3 NPLs; (g) hexagonal Cs4PbBr6 nanocrystals; and (h) transformation of the
unit cell and optical properties between the two phases with Cs-oleate and PbBr2 additions. Reproduced from ref 113. Copyright 2018 American
Chemical Society.

had lateral lengths ∼14−27 nm with thicknesses of ∼1.5 nm and NRs synthesized from a modified HI method in which BiBr3 and AgNO3
with diameters of ∼46 nm with lengths of ∼655 nm. These PNCs were first dissolved in ODE at 200 °C using OA, OLA, and
exhibited low PLQYs of ∼5% due to deep defects originating from the hydrobromic acid solubilizers. Then a solution of Cs-oleate in ODE
Sb 5p orbital splitting. was injected into this reaction mixture to form 9.5 nm NCu of
Beyond single B-site perovskites, double perovskites have also been Cs2AgBiBr6 after 5 s of reaction time.109 Alternative syntheses of
demonstrated using monovalent Ag+ and trivalent Bi3+ B-site ion pairs Cs2AgBiX6 (X = Cl, Br) have explored dissolving the acetate salts of Cs,
to synthesize Cs2AgBiX6.108 Colloidal Cs2AgBiBr6 PNCs were Ag, and Bi in ODE, OA, and OLA solutions and then injecting

11618 DOI: 10.1021/acs.langmuir.9b00855


Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

Figure 12. (a) TEM micrographs, (b) XRD patterns, and (c) absorption and PL spectra of different PNC halide compositions. Reproduced with
permission from ref 115. Copyright 2017 Royal Society of Chemistry.

trimethylsilyl halide at 140 °C into grown ∼8.1 nm NCu over 10 to 60 be tuned across the visible region while maintaining high PLQYs of up
s.110 The decoupling of the metallic salts and the halide precursors in to 80% and narrow fwhm’s of 12 to 36 nm. It was additionally shown
this approach provided greater control over the stoichiometry. These that if the precursors were predissolved prior to heating the reaction
cubic double perovskites exhibited low PLQY (0.3%),110 albeit with mixture, then the modified reaction kinetics results in NWs with a
improved thermal, light, and moisture stability compared to that of thickness of only ∼3.2 nm and lengths of several hundred nanometers
other Pb-free perovskites.109 Notably, detailed structural investigations could be produced in almost 100% morphological yield (Figure 11b−
have indicated that the major degradation pathway in these materials e).
arises from silver diffusion and its chemical reduction and coalescence The solvothermal synthesis of CsPbBr3 was further modified by
to form Ag and trigonal-phase Cs3Bi2X9 and Cs3BiX6.110,111 Over- tuning the Cs/Pb precursor ratio and the reaction time.113 Using
coming these structural limitations remains one of the existing predissolved precursors, at a Cs/Pb molar ratio of ∼1:10 and a reaction
challenges in effectively harnessing double perovskite structures. time of at least 30 min at 100 °C, ∼4.2-nm-thick NPLs with lateral
Convection. Convection synthesis (often referred to generally as lengths of ∼100 nm and widths of ∼15 nm formed (Figure 11f). As the
“heat-up”) is among the simplest noninjection syntheses, relying on thickness of the NPLs remained unchanged in the quantum
controlled heating to induce nucleation and growth. It is also the least confinement regime, the PL emission could be fine-tuned to between
explored, owing to the still nascent understanding of precursor 452 to 465 nm by controlling the lateral NPL sizes to between 101 × 15
evolution for PNCs, resulting in challenges in controlling their nm and 44 × 42 nm, respectively. At progressively longer reaction
nucleation and growth dynamics. Nonetheless, Chen et al. has used times, the NPLs exhibited coarsening behavior, demonstrated by a
this approach to successfully synthesize CsPbX3 PNCs.46 Cs2CO3 and shortening of their lengths and an increase in their widths. Similar
PbX2 precursors were dissolved in ODE using OA and OLA ligands at trends in the shape evolution were observed as the Cs/Pb molar ratio
room temperature, and then the reaction mixture was heated to was increased, although at a nearly equimolar ratio rhombohedral-
temperatures ranging from between 50 and 170 °C for 10 min to phased Cs4PbBr6 PNCs with a hexagonal shape were observed (Figure
nucleate and grow 1 to 4 nm CsPb(Cl/Br)3 and 1.6 to 10.4 nm CsPbI3
11g,h).
QDs (Figure 10a−d). While the size tunability demonstrated in this
The synthesis of Pb-free perovskites using a solvothermal method
approach is superior to those for injection methods, it is evident that it
has also been demonstrated with CsSnX3 (X = Cl, Br, I) NRs.114 In this
also yielded PNCs with higher polydispersity and irregular shapes.
approach, SnX2 was mixed with ODE, OA, and OLA. A secondary
Despite this, the control of size and composition enabled tunable PL
mixture of Cs2CO3, ODE, OA, hydrochloric acid, TOPO, and
across the entire visible region with PLQYs of >80% (Figure 10e−g).
The scalability of the reaction was demonstrated at gram levels. diethylenetriamine was then combined with the first mixture under
Tsiwah et al. further showed that by using reaction chemistries stirring at 80 °C for an hour. The combined mixture was further diluted
similar to that above, bromide PNCs exhibited a rhombohedral with diethylenetriamine and transferred to an autoclave at 180 °C for 6
Cs4PbBr6 to orthorhombic CsPbBr3 phase transformation upon h. This resulted in NRs of 5 nm diameter and 10 nm length, with
increased reaction temperature or time.112 By replacing the OLA absorption onsets and photoluminescence emission peaks that could be
with shorter-chained octylamine, NSs with a diameter of several tuned via their halide composition from 588 to 688 nm and 625 to 709
micrometers were produced. This indicated that the shorter amine nm, respectively.
similarly induced anisotropic growth in both injection and convection Microwave. The use of a microwave as the heat source requires a
syntheses. polar solvent and/or ligands within a reaction medium that can absorb
Solvothermal. The solvothermal route is a simple, scalable, and the microwave radiation to ensure rapid and uniform heating. Long et
versatile method for nanocrystal production. It is performed in an al. were the first to report the microwave synthesis of inorganic PNCs
isochoric environment, with the reaction being conducted in a (Figure 12).115 The authors reacted Cs2CO3, PbX2, OA, and OLA in
pressurized vessel at elevated temperature. Chen et al. was the first to ODE at a Cs/Pb molar ratio of 1:3 within a 800 W household
use the solvothermal method to prepare CsPbX3 PNCs.55 Cesium microwave oven for 4 min. Despite the rather rudimentary nature of the
acetate (CsAc), PbX2, OA, and OLA in ODE were placed inside an synthesis, 10 to 13 nm NCu CsPbX3 PNCs were achieved. The
autoclave, with TOP and TOPO being used to enhance the final microwave irradiation time was correlated to the size of the PNCs, with
solubility of the different lead salts. Upon heating the reaction mixture times of between 2 and 7 min yielding tunable sizes of between 5 and 14
to 160 °C for 30 min in a rolling oven, high-quality perovskite NCu with nm. By decreasing the Cs/Pb precursor ratio to 1:6 in the above
a size of 8 to 12 nm was produced (Figure 11a). By using different lead synthesis, NWs with a length of >100 nm were preferentially
halides or their mixtures, the emission properties of these PNCs could formed.115,116

11619 DOI: 10.1021/acs.langmuir.9b00855


Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

Figure 13. (a) Schematic illustration of ultrasonication PNC synthesis. (b) Photographs under visible and UV light for PNCs with different halide
compositions (c) and their respective absorption and PL spectra. Reproduced with permission from ref 70. Copyright 2016 John Wiley and Sons. (d)
TEM micrograph for a CsPbBr3 supercrystal. (e) PL comparison for normal and supercrystal CsPbBr3 films. Reproduced with permission from ref 119.
Copyright 2016 John Wiley and Sons.

Pan et al. further investigated the synthesis of CsPbBr3 NCs using a Analogous to the solvothermal method, decreasing the Cs/Pb ratio
microwave reactor, studying factors that included temperature, the favored the formation of NPLs.
predissolution of precursors, and heating rates.71 It was found that a By increasing the concentration 10-fold (1 mmol Cs2CO3, 3 mmol
synthesis temperature of ∼80 °C drove anisotropic growth, yielding PbBr2) and extending the reaction time to 30 min, it has been shown
NPLs with a thickness of ∼3.3 nm and edge lengths of ∼23.4 nm. that the ultrasonication approach can produce CsPbBr3 supercrystals of
Increasing the temperature to 140 °C produced homogeneous ∼10.2 200 to 400 nm diameter that are composed of closely assembled 10
nm NCu PNCs. A further temperature rise to 180 °C yielded sized NCu (Figure 13d).119 Such supercrystals exhibit PLQYs
inhomogeneous PNC dispersions. Predissolving the precursors at 60 comparable to those of their isolated NC counterparts but are red-
°C and then reacting them at 160 °C under otherwise identical shifted by ∼20 nm due to strong electronic coupling (Figure 13e).
conditions resulted in the formation of NRs at morphological yields of
>90%, with lengths of >60 nm and diameters of only ∼2.7 nm. It was
noted that the presence of TOPO in the reaction mixture was critical to
obtaining high-quality NRs. The authors suggested that the TOPO
■ POSTSYNTHETIC TREATMENTS
Purification. Postsynthetic purification of PNCs involves
aided the dissolution of the precursors. Finally, a minimum heating rate the removal of residual solvent, ligand, and precursor impurities
of 18 °C/min was found to be required to generate homogeneous through multiple precipitation and redispersion steps. Owing to
PNCs. A slower heating rate of 8 °C/min significantly broadened the the highly ionic perovskite structure, antisolvents with larger
PNC size distribution and produced particles >150 nm in size. The dipole moments of ∼4.0 D (DMF, DMSO) or protic solvents
latter is consistent with the fundamental limitation of heat-up syntheses (methanol, ethanol) in which the underlying perovskite lattice
in that slow heating rates prolong the nucleation event to yield exhibits a high solubility have been found not to preserve the
inhomogeneous dispersions.117 structural properties of the nanocrystals.79 As a result, modest-
Ultrasonication. The formation of hybrid MAPbX3 (X = Br, I) polarity aprotic antisolvents have been favored. The inability of
PNCs has been demonstrated using ultrasonication, albeit in an indirect
such solvents to undergo proton-mediated X-type ligand
manner. Huang et al. dispersed PbX2 and MAX in a mixture of OA and
OLA that acted as a coordinating solvent.118 This mixture was exchange further enables a better preservation of the surface
ultrasonicated to form bulk MAPbX3 within 5 min. After several further chemistry during purification.75
hours of ultrasonication, these bulk perovskite crystals were trans- Because of the modest binding constants of the ligands at the
formed into nanocrystals. It is not clear whether these were generated surface of PNCs, the required multiple purification cycles
through a dissolution−recrystallization process and/or progressive decrease the ligand density at the PNC surface.120 To mitigate
fragmentation caused by the cavitation process. Nonetheless, an ligand loss and the deterioration of PNC quality during this
analysis of the MAPbBr3 samples indicated that the PNCs were ∼4 nm process, it has been shown that small amounts of additional
in size and possessed PLQYs of up 72%. organic acid and amine ligands are required to maintain the
In a more direct synthesis, Tong et al. was the first to use acid−base equilibria required for colloidal stability and the
ultrasonication to synthesize CsPbX3 nanocrystals (Figure 13a).70
Cs2CO3, PbX2, OA, and OLA were mixed in ODE under ambient air,
retention of PL properties.75,121 Despite this, under a large
and then a 30 W sonication tip was applied to the mixture for 10 min. excess of amine,80,122 the irreversible degradation of PNCs was
This induced the in situ formation of solubilized Cs+ and Pb2+ precursor observed, as governed by eq 6. Excess acid was also found to
complexes, which concurrently reacted to produce CsPbX3 (X = Cl, Br, destabilize the PNCs.122
I, Cl/Br, and Br/I) PNCs. The sizes of the CsPbBr3 and CsPbI3 PNCs amine
were determined to be 10 to 15 nm and 8 to 12 nm, respectively, which 4CsPbBr3 ⎯⎯⎯⎯⎯→ Cs4PbBr6 + 3PbBr2 (6)
are slightly higher than those obtained via HI. The produced PNCs 123
exhibited visibly tunable optical absorption and emission properties, Postsynthetic additives in the form of thiocyanate- and
with the emission possessing QYs of up to 90% for the Br and I tetrafluoroborate-based124 salts have also been successfully used
derivatives and narrow fwhm’s of 12 to 40 nm (Figure 13b,c). to improve PNCs, achieving PLQYs approaching unity. It is
11620 DOI: 10.1021/acs.langmuir.9b00855
Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

Figure 14. Postsynthesis anion exchange for mixed PNC halide compositions. (a) Schematic of possible halide exchanges, (b) XRD spectra, (c)
combined UV−vis absorption (solid lines) and PL emission spectra (dashed lines), and TEM micrograph of (d) CsPb(Br1.5Cl1.5) and (e)
CsPb(Br1.5I1.5). Reproduced from ref 129. Copyright 2015 American Chemical Society.

Figure 15. Postsynthesis cation exchange for mixed PNC B-site compositions. Colloidal CsPbBr3 PNCs under UV illumination with increasing
additions of (a) SnBr2, (b) CdBr2, and (c) ZnBr2. Combined UV−vis absorption (dashed lines) and PL emission (solid lines) spectra of CsPbBr3
dispersions with (d) Sn2+ and (e) Cd2+ and Zn2+ additions resulting in spectral blue shifts. Reproduced from ref 130. Copyright 2017 American
Chemical Society.

believed that these anions salts remove excess surface lead atoms wash cycles.126 Further exploration of such additives during
from the surface, thereby removing shallow traps that purification processes will be necessary to achieve the optimized
detrimentally impact the optical properties. Similar effects structural and optoelectronic properties of PNCs.
have been observed with the addition of ammonium salts to Anion Exchange. The highly ionic nature of PNCs can
enhance surface passivation and restrict PNC growth to improve readily undergo postsynthetic anion exchange reactions to
their overall colloidal and PL stability.125 Moreover, the use of enable tuning of their optical properties (Figure 14).16 The
mixed treatment solutions, in the form of didodecyldimethyl- relatively rigid cationic sublattice and the high diffusivity of
ammonium bromide (DDAB) and PbBr2 in toluene has resulted halide vacancies31 facilitate rapid Cl− ↔ Br− and Br− ↔ I− halide
in PNCs with PLQYs of >90% that were amenable to multiple exchange across all PNCs upon the addition of the introduced
11621 DOI: 10.1021/acs.langmuir.9b00855
Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

halide source (e.g., PbX2 solution or an ammonium halide toluene mixture to directly dope CsPbBr3 NPLs and NCu.134 In
solution). The extent of anion exchange can be fully controlled this case, sensitized emission from the Mn was observed only at
by the ratio between the original and introduced halides.127 The room temperature for the NPLs, with back energy transfer from
exchange process retains monodisperse colloidal PNCs with the Mn to the host PNCs causing quenching for the NCu.
similar optical properties compared to directly synthesized Notably, the same authors showed that a similar cosolvent
PNCs of the same composition.31,115 Furthermore, this process approach, albeit using toluene and methyl acetate, could
is morphologically independent, as has been demonstrated on successful dope CsPbX3 with Yb3+, as evidenced by its
NW- and NPL-shaped PNCs.52,128 However, the exchange characteristic 2F5/2 → 2F7/2 IR emission being observed at
process is not universal across halide-based perovskites. Because ∼990 nm. Importantly, the doping of both Mn and Yb into the
of the large difference in ionic radii of Cl− (1.81 Å) and I− (2.20 CsPbX3 nanocrystals narrowed their Urbach tail, indicating a
Å), attempts to perform a Cl− ↔ I− halide exchange process favorable reduction in their intrinsic defect states.
have resulted in only the slow conversion to a full trihalide Attempts to substituting the A-site Cs+ cations with Rb+, Ag+,
perovskite material.31 To circumvent this limitation, it has been Cu+, or Ba2+ or the B-site Pb2+ cations with Ge2+ or Bi3+ have
shown that Cl− ↔ I− exchanges can be achieved through resulted in the decomposition of the original CsPbX3 PNCs.31
intermediate exchange processes using Br− (1.96 Å). This This incompatibility could stem from oxidative instabilities,
unique property of PNC systems to undergo anion exchange is incompatible ion sizes, or competition with more thermody-
the key reason for its ability to easily tune its emission across the namically stable phases.10
visible spectrum. In comparison, traditional metal chalcogenide
nanocrystal systems undergo only cation exchanges and are
limited to tuning their emissions via the quantum confinement
■ SUMMARY
Here we have assessed the progress toward the synthesis of
effect through particle size control.3 colloidal metal halide perovskite nanocrystals. These materials
Cation Exchange. Analogous to anionic exchange pro- have leveraged advances from traditional semiconductor
cesses, cation exchange at both the A and B sites can similarly colloids to achieve modest size and morphological control
tune the optoelectronic properties of PNCs. Akkerman et al. across various hybrid and inorganic compositions within a span
observed that for CsPbX3, the A-site Cs+ (1.67 Å) could be of only 5 years. This structural and compositional control has
exchanged for the larger MA+ (2.17 Å) cation.129 Similarly, Cs+ enabled the exceptional optoelectronic properties of perovskites
and FA+ (2.53 Å) cations could be interchanged to form in their nanocrystalline forms to be showcased, with visibly
(FA,Cs)PbI3 PNCs.12 tunable absorption and photoluminescence properties, narrow
Partial exchange of the B-site Pb2+ cation in CsPbBr3 was first emission bands, and near-unity photoluminescence quantum
demonstrated by van der Stam et al. by mixing a PNCs solution yields being demonstrated. These properties make perovskite
and M2+ cation precursor solutions (MBr2, M = Sn2+, Zn2+, and nanocrystals a remarkable candidate for many optoelectronic
Cd 2+ ) in toluene. 130 The emission of the resultant applications, where high efficiency and low-temperature
CsPb1−xMxBr3 PNCs exhibited a blue shift proportional to the processing are advantageous.
M2+ precursor concentration (Figure 15). The blue shift of the The hot-injection method has emerged as the most common
emission was attributed to the contraction of the PbBr6 for synthesizing high-quality perovskite nanocrystals. Relying on
polyhedra, after partial substitution of Pb2+ cations by the a high-temperature mixing process under aliphatic chemical
smaller M2+ cations, which caused a closer and stronger Pb−Br conditions, control of the rapid nucleation and growth stages
orbital interaction that raised the conduction band minimum. through ligand, precursor, solvent, and temperature profile
Despite M2+ being present in large excess, there was only partial tuning has yielded reasonable size- and shape-controlled
(≤10%) substitution of the M2+ cation into the PNC. It was perovskite nanocrystals. Lower-temperature injection syntheses
reasoned that the perovskite crystal structure is primarily that mostly use aliphatic and polar solvent mixtures have been
stabilized by the rigid cationic sublattice and that cation concurrently developed. These have taken advantage of the
substitution, especially at the B site, required higher activation slower kinetics to scale the reactions from tens of mL up to >1 L.
energies than for anion exchange.31 This is consistent with the A number of noninjection methods have also emerged as
findings from studies on the solid state ion exchanges in single- suitable synthetic alternatives that are more suitable for scale up.
crystal perovskite NWs.131 By adopting similar reaction mixtures to the injection methods,
To overcome this B-site stability, Mondal et al. showed that perovskite nanocrystals of varying size and shape have been
ultrasonication facilitated enhanced alloying of CsPbCl3 PNCs synthesized through these, although with lesser structural
with Cd2+ using a saturated solution of CdCl2 in ethanol.132 tunability and quality compared to those of the injection
Doping levels of up to 40% were claimed, with the PLQY approaches.
increasing from 3% to near unity. The significant PLQY increase In reviewing all of the reported synthetic approaches to date, it
was attributed to the removal of nonradiative defect states is apparent that a greater understanding of the nucleation and
caused by chloride vacancies and a reduction in the PbCl6 growth processes, the precursor evolution, and the interactions
polyhedral tilting through the incorporation of the smaller Cd2+ at the perovskite nanocrystal surface is still required for each
ions. This also gave Cd/CsPbCl3 PNCs improved photo- method. The fast dynamics in these systems have largely masked
stability, with PLQYs of up to 90% being retained after several the intricate details around the nucleation process and early
months. growth stages. Further uncertainties around specific solvent and
Postsynthetic Mn2+ doping of PNCs has also been explored. ligand interactions at the perovskite nanocrystal surface, which
Huang et al. used a concurrent halide exchange approach include the role that stereochemistry, polarity, conjugation, and
through the addition of MnCl2 in DMF to CsPbBr3 PNCs in chemical functionality play in mediating ligand−surface
toluene to yield PNCs with characteristic emission peaks of the interactions, provide an incomplete picture of what is occurring
sensitized Mn and CsPb(Cl,Br)3.133 In a more direct approach, at the surface and how the specific interactions can be tuned.
Nag’s group simply used MnBr2 dissolved in an acetone and Progress in these areas will require the expansion of the currently
11622 DOI: 10.1021/acs.langmuir.9b00855
Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

limited library of precursors and ligands as well as more detailed


theoretical studies to understand the synthetic evolution of the
reaction mixtures. These insights will concurrently assist in
further developing postsynthetic ligand and ion-exchange
processes to provide greater structural and optical versatility of
these materials.
In going forward, the overarching goal in the field of
perovskite nanocrystals is the development of stable and
nontoxic perovskite derivatives that maintain the lucrative
optoelectronic properties of existing lead-based derivatives.
Preliminary works on hybrid and inorganic systems, such as
MA3Bi2X9 and Cs3Sb2X9, respectively, have shown great promise
toward achieving this goal. However, the low-quantum yields,
high intrinsic defects, and structural instabilities that are
currently observed within such materials remain major hurdles. Chujie Wang received his BE (Honors) in materials science and
To overcome these, tremendous effort in tailored ligand and engineering from Monash University and Central South University in
precursor design and compositional engineering will be required 2015. During his bachelor’s degree, he worked on different projects in
to achieve the degree of synthetic and structural control several research groups: mass production of reduced graphene (Prof.
necessary to yield high-quality perovskite nanomaterials with the Dan Li) and tissue engineering of artificial skin (Prof. Neil Cameron
desirable optoelectronic properties. Through these continuing and Alfred Hospital). In 2016, he continued at Monash University with
advances, this exciting field of perovskite nanocrystals will reach a Ph.D. in the Department of Materials Science and Engineering under
a state of maturity, one that will deepen our understanding of the supervision of Assoc. Prof. Jacek Jasieniak. His current research
structure−property relations at the nanoscale and foster a interests are focused on the phase transformation and surface chemistry
paradigm shift in the application of nanocrystals across many of inorganic perovskite nanocrystals.
fields.

■ AUTHOR INFORMATION
Corresponding Author
*E-mail: jacek.jasieniak@monash.edu.
ORCID
Chun Kiu Ng: 0000-0001-5456-1020
Jacek J. Jasieniak: 0000-0002-1608-6860
Author Contributions
The manuscript was conceived by J.J. and written through the
contributions of all authors.
Notes
The authors declare no competing financial interest.
Biographies
Jacek Jasieniak completed a Bachelor of Science (First Class Honours)
from Flinders University (2003) and a Ph.D. from the University of
Melbourne (2008) under the supervision of Prof. Paul Mulvaney. He
then undertook postdoctoral work at the Commonwealth Scientific and
Industrial Research Organisation (CSIRO) with Dr. Scott Watkins and
Dr. Ezio Rizzardo (2008−2011), and was a Fulbrigh Scholar with Prof.
Alan Heeger at the University of CaliforniaSanta Barbara (2011 to
2012). In 2012, he returned to CSIRO, progressing to a senior research
scientist and then group leader. In 2015, he moved to Monash
University as an associate professor, where he was also appointed as the
director of the cross-disciplinary Monash Energy Materials and Systems
Institute. His research interests include the development of nanoma-
terials and their use in various next-generation energy technologies.

Chun Kiu received his BE(Hons)/BSc at Monash University in 2015.


During his bachelor’s degree, he worked on numerous projects under
■ ACKNOWLEDGMENTS
The authors thank the ARC Centre of Excellence in Exciton
Science (CE170100026) for financial support.


several academics: cobalt electrolytes for dye-sensitized solar cells
(Prof. Udo Bach), perovskite photovoltaic-driven water splitting (late
Prof. Leonne Spiccia), and organic photovoltaics (Prof. Chris McNeill). ABBREVIATIONS
In 2016, he continued at Monash University with a Ph.D. in the °C, degrees Celsius; μm, micrometer; Ag, silver; Ba, barium; Bi,
Department of Materials Science and Engineering under the super- bismuth; Br−, bromide; Cd, cadmium; CdBr2, cadmium
vision of Assoc. Prof. Jacek Jasieniak in the area of perovskite bromide; CdSe, cadmium selenide; Ce, cerium; Sm, samarium;
nanocrystal synthesis and its incorporation into solar cells. Cl−, chloride; cm, centimeter; Cs, cesium; Cs2CO3, cesium
11623 DOI: 10.1021/acs.langmuir.9b00855
Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

carbonate; Cs4PbBr6, zero-dimensional cesium lead bromide; Efficient Perovskite Solar Cells. Energy Environ. Sci. 2017, 10 (3), 710−
CsAc, cesium acetate; CsBr, cesium bromide; CsOA, cesium 727.
oleate; CsPbBr3, cesium lead bromide; CsPbCl3, cesium lead (7) Yin, Y.; Alivisatos, A. P. Colloidal Nanocrystal Synthesis and the
chloride; CsPbI3, cesium lead iodide; CsPbX3, cesium lead Organic-Inorganic Interface. Nature 2005, 437 (7059), 664−670.
halide; CsX, cesium halide; Cu, copper; DDAB, didodecyldi- (8) Reiss, P.; Protière, M.; Li, L. Core/Shell Semiconductor
Nanocrystals. Small 2009, 5 (2), 154−168.
methylammonium bromide; D, dipole moment in debye; DFT, (9) Norris, D. J.; Efros, A. L.; Erwin, S. C. Doped Nanocrystals. Science
density functional theory; DMF, dimethylformamide; DMSO, (Washington, DC, U. S.) 2008, 319 (5871), 1776−1779.
dimethyl sulfoxide; Dy, dysprosium; E0, standard reduction (10) Kovalenko, M. V.; Protesescu, L.; Bodnarchuk, M. I. Properties
potential; EDX, energy-dispersive X-ray; Er, erbium; EtOH, and Potential Optoelectronic Applications of Lead Halide Perovskite
ethanol; QDs, quantum dots; Eu, europium; eV, electronvolt; Nanocrystals. Science (Washington, DC, U. S.) 2017, 358 (6364), 745−
FA+, formamidinium; FACsI3, formamidinium cesium iodide; 750.
FAPbBr3, formamidinium lead bromide; FAPbI3, formamidi- (11) Swarnkar, A.; Ravi, V. K.; Nag, A. Beyond Colloidal Cesium Lead
nium lead iodide; FTIR, Fourier transform infrared spectrosco- Halide Perovskite Nanocrystals: Analogous Metal Halides and Doping.
py; fwhm, full width at half-maximum; min, minutes; g, gram; ACS Energy Lett. 2017, 2 (5), 1089−1098.
GaAs, gallium arsenide; Ge, germanium; HI, hot injection; I−, (12) Protesescu, L.; Yakunin, S.; Kumar, S.; Bär, J.; Bertolotti, F.;
iodide; IPA, isopropanol; LARP, ligand-assisted reprecipitation; Masciocchi, N.; Guagliardi, A.; Grotevent, M.; Shorubalko, I.;
LEDs, light-emitting diodes; M, moles per liter; MA+, Bodnarchuk, M. I.; et al. Dismantling the “Red Wall” of Colloidal
methylammonium; MABr, methylammonium bromide; Perovskites: Highly Luminescent Formamidinium and Formamidi-
nium−Cesium Lead Iodide Nanocrystals. ACS Nano 2017, 11 (3),
MAPbBr3, methylammonium lead bromide; MAPbCl3, methyl-
3119−3134.
ammonium lead chloride; MAPbI3, methylammonium lead (13) Gonzalez-Carrero, S.; Espallargas, G. M.; Galian, R. E.; Pérez-
iodide; MAPbX3, methylammonium lead halide; MASnCl3, Prieto, J. Blue-Luminescent Organic Lead Bromide Perovskites: Highly
methylammonium tin chloride; MASnI3, methylammonium tin Dispersible and Photostable Materials. J. Mater. Chem. A 2015, 3 (26),
iodide; MAX, methylammonium halide; mL, millimeters; mm, 14039−14045.
millimeter; Mn, manganese; MnCl2, manganese chloride; (14) Liu, F.; Zhang, Y.; Ding, C.; Kobayashi, S.; Izuishi, T.; Nakazawa,
MnX2, manganese halide; N2, nitrogen gas; NCu, nanocubes; N.; Toyoda, T.; Ohta, T.; Hayase, S.; Minemoto, T.; et al. Highly
NH3+, ammonium; NH4X, ammonium halide; nm, nanometer; Luminescent Phase-Stable CsPbI3 Perovskite Quantum Dots Achiev-
NMR, nuclear magnetic resonance; NPL, nanplatelet; NR, ing Near 100% Absolute Photoluminescence Quantum Yield. ACS
nanorod; ns, nanoseconds; NS, nanosheet; NT, nontemplate; Nano 2017, 11 (10), 10373−10383.
NW, nanowire; OA−, oleate; OA, oleic acid; ODE, 1- (15) Pan, J.; Shang, Y.; Yin, J.; De Bastiani, M.; Peng, W.; Dursun, I.;
octadecene; OLA, oleylamine; OLA + , oleylammonium; Sinatra, L.; El-Zohry, A. M.; Hedhili, M. N.; Emwas, A. H.; et al.
OLAX, oleylammonium halide; Pb, lead; PbBr2, lead bromide; Bidentate Ligand-Passivated CsPbI3 Perovskite Nanocrystals for Stable
PbCl2, lead chloride; PbI2, lead iodide; PbO, lead oxide; PbOA2, Near-Unity Photoluminescence Quantum Yield and Efficient Red
Light-Emitting Diodes. J. Am. Chem. Soc. 2018, 140 (2), 562−565.
lead oleate; PbS, lead sulfide; PbSe, lead selenide; PbX2, lead (16) Protesescu, L.; Yakunin, S.; Bodnarchuk, M. I.; Krieg, F.; Caputo,
halide; PL, photoluminescence; PLQY, photoluminescence R.; Hendon, C. H.; Yang, R. X.; Walsh, A.; Kovalenko, M. V.
quantum yield; PNC, perovskite nanocrystal; ref, reference; Nanocrystals of Cesium Lead Halide Perovskites (CsPbX3, X = Cl, Br,
Rb, rubidium; rpm, revolutions per minute; s, seconds; SAXS, and I): Novel Optoelectronic Materials Showing Bright Emission with
small-angle X-ray scattering; Sb, antimony; SbCl3, antimony Wide Color Gamut. Nano Lett. 2015, 15 (6), 3692−3696.
chloride; SEM, scanning electron microscopy; Sn, tin; SnBr2, tin (17) Sun, S.; Yuan, D.; Xu, Y.; Wang, A.; Deng, Z. Ligand-Mediated
bromide; t, Goldschmidt tolerance factor; Tb, terbium; TEM, Synthesis of Shape-Controlled Cesium Lead Halide Perovskite
transmission electron microscopy; TMPPA, bis(2,4,4-trimethyl- Nanocrystals via Reprecipitation Process at Room Temperature. ACS
pentyl)phosphinic acid; TOP, trioctylphosphine; TOPO, tri- Nano 2016, 10 (3), 3648−3657.
octylphosphine oxide; UV−vis, ultraviolet−visible; V, volt; W, (18) Lignos, I.; Stavrakis, S.; Nedelcu, G.; Protesescu, L.; Demello, A.
watts; mmol, millimole; X, halide; XPS, X-ray photoelectron; J.; Kovalenko, M. V. Synthesis of Cesium Lead Halide Perovskite
XRD, X-ray diffraction; Yb, ytterbium; Zn, zinc; ZnBr2, zinc Nanocrystals in a Droplet-Based Microfluidic Platform: Fast Parametric
bromide; ZnX2, zinc halide. Space Mapping. Nano Lett. 2016, 16 (3), 1869−1877.


(19) Swarnkar, A.; Chulliyil, R.; Ravi, V. K.; Irfanullah, M.;
Chowdhury, A.; Nag, A. Colloidal CsPbBr3 Perovskite Nanocrystals:
REFERENCES Luminescence beyond Traditional Quantum Dots. Angew. Chem., Int.
(1) IndustryARC. Optoelectronics Market: By Components, by Devices, Ed. 2015, 54 (51), 15424−15428.
by End User Industry and by Geography; Forecast 2016−2021, 2015. (20) Shi, Z.; Li, S.; Li, Y.; Ji, H.; Li, X.; Wu, D.; Xu, T.; Chen, Y.; Tian,
(2) Srivastava, H.; Jayamon, J. Compound Semiconductor Market: Y.; Zhang, Y.; et al. Strategy of Solution-Processed All-Inorganic
Global Opportunity Analysis and Industry Forecast, 2017−2023, 2017. Heterostructure for Humidity/Temperature-Stable Perovskite Quan-
(3) Akkerman, Q. A.; Rainò, G.; Kovalenko, M. V.; Manna, L. Genesis, tum Dot Light-Emitting Diodes. ACS Nano 2018, 12 (2), 1462−1472.
(21) Song, J.; Li, J.; Li, X.; Xu, L.; Dong, Y.; Zeng, H. Quantum Dot
Challenges and Opportunities for Colloidal Lead Halide Perovskite
Light-Emitting Diodes Based on Inorganic Perovskite Cesium Lead
Nanocrystals. Nat. Mater. 2018, 17 (5), 394−405.
Halides (CsPbX3). Adv. Mater. 2015, 27 (44), 7162−7167.
(4) Schmidt, L. C.; Pertegás, A.; González-Carrero, S.; Malinkiewicz,
(22) Yakunin, S.; Protesescu, L.; Krieg, F.; Bodnarchuk, M. I.;
O.; Agouram, S.; Mínguez Espallargas, G.; Bolink, H. J.; Galian, R. E.; Nedelcu, G.; Humer, M.; De Luca, G.; Fiebig, M.; Heiss, W.;
Pérez-Prieto, J. Nontemplate Synthesis of CH3NH3PbBr3 Perovskite Kovalenko, M. V. Low-Threshold Amplified Spontaneous Emission
Nanoparticles. J. Am. Chem. Soc. 2014, 136 (3), 850−853. and Lasing from Colloidal Nanocrystals of Caesium Lead Halide
(5) Li, W.; Wang, Z.; Deschler, F.; Gao, S.; Friend, R. H.; Cheetham, Perovskites. Nat. Commun. 2015, 6, 8056.
A. K. Chemically Diverse and Multifunctional Hybrid Organic− (23) Fu, Y.; Zhu, H.; Stoumpos, C. C.; Ding, Q.; Wang, J.; Kanatzidis,
Inorganic Perovskites. Nat. Rev. Mater. 2017, 2 (3), 16099. M. G.; Zhu, X.; Jin, S. Broad Wavelength Tunable Robust Lasing from
(6) Correa-Baena, J.-P.; Abate, A.; Saliba, M.; Tress, W.; Jesper Single-Crystal Nanowires of Cesium Lead Halide Perovskites
Jacobsson, T.; Grätzel, M.; Hagfeldt, A. The Rapid Evolution of Highly (CsPbX3, X = Cl, Br, I). ACS Nano 2016, 10 (8), 7963−7972.

11624 DOI: 10.1021/acs.langmuir.9b00855


Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

(24) Lv, L.; Xu, Y.; Fang, H.; Luo, W.; Xu, F.; Liu, L.; Wang, B.; Zhang, Nanocrystals: Highly Improved and Expanded Optical Properties.
X.; Yang, D.; Hu, W.; et al. Generalized Colloidal Synthesis of High- Nano Lett. 2017, 17 (12), 8005−8011.
Quality, Two-Dimensional Cesium Lead Halide Perovskite Nanosheets (42) Dastidar, S.; Egger, D. A.; Tan, L. Z.; Cromer, S. B.; Dillon, A. D.;
and Their Applications in Photodetectors. Nanoscale 2016, 8 (28), Liu, S.; Kronik, L.; Rappe, A. M.; Fafarman, A. T. High Chloride
13589−13596. Doping Levels Stabilize the Perovskite Phase of Cesium Lead Iodide.
(25) Ramasamy, P.; Lim, D.-H.; Kim, B.; Lee, S.-H.; Lee, M.-S.; Lee, Nano Lett. 2016, 16 (6), 3563−3570.
J.-S. All-Inorganic Cesium Lead Halide Perovskite Nanocrystals for (43) Eperon, G. E.; Paterno, G. M.; Sutton, R. J.; Zampetti, A.;
Photodetector Applications. Chem. Commun. 2016, 52 (10), 2067− Haghighirad, A. A.; Cacialli, F.; Snaith, H. J. Inorganic Caesium Lead
2070. Iodide Perovskite Solar Cells. J. Mater. Chem. A 2015, 3 (39), 19688−
(26) Pan, J.; Sarmah, S. P.; Murali, B.; Dursun, I.; Peng, W.; Parida, M. 19695.
R.; Liu, J.; Sinatra, L.; Alyami, N.; Zhao, C.; et al. Air-Stable Surface- (44) Fu, Y.; Rea, M. T.; Chen, J.; Morrow, D. J.; Hautzinger, M. P.;
Passivated Perovskite Quantum Dots for Ultra-Robust, Single- and Zhao, Y.; Pan, D.; Manger, L. H.; Wright, J. C.; Goldsmith, R. H.; et al.
Two-Photon-Induced Amplified Spontaneous Emission. J. Phys. Chem. Selective Stabilization and Photophysical Properties of Metastable
Lett. 2015, 6 (24), 5027−5033. Perovskite Polymorphs of CsPbI3 in Thin Films. Chem. Mater. 2017, 29
(27) Sanehira, E. M.; Marshall, A. R.; Christians, J. A.; Harvey, S. P.; (19), 8385−8394.
Ciesielski, P. N.; Wheeler, L. M.; Schulz, P.; Lin, L. Y.; Beard, M. C.; (45) Wang, C.; Chesman, A. S. R.; Jasieniak, J. J. Stabilizing the Cubic
Luther, J. M. Enhanced Mobility CsPbI3 Quantum Dot Arrays for Perovskite Phase of CsPbI3 Nanocrystals by Using an Alkyl Phosphinic
Record-Efficiency, High-Voltage Photovoltaic Cells. Sci. Adv. 2017, 3 Acid. Chem. Commun. 2017, 53 (1), 232−235.
(10), eaao4204. (46) Chen, X.; Peng, L.; Huang, K.; Shi, Z.; Xie, R.; Yang, W. Non-
(28) Akkerman, Q. A.; Gandini, M.; Di Stasio, F.; Rastogi, P.; Palazon, Injection Gram-Scale Synthesis of Cesium Lead Halide Perovskite
F.; Bertoni, G.; Ball, J. M.; Prato, M.; Petrozza, A.; Manna, L. Strongly Quantum Dots with Controllable Size and Composition. Nano Res.
Emissive Perovskite Nanocrystal Inks for High-Voltage Solar Cells. Nat. 2016, 9 (7), 1994−2006.
Energy 2017, 2 (2), 16194. (47) Huang, H.; Susha, A. S.; Kershaw, S. V.; Hung, T. F.; Rogach, A.
(29) Goldschmidt, V. M. Laws of Crystal Chemistry. Naturwissen- L. Control of Emission Color of High Quantum Yield CH3NH3PbBr3
schaften 1926, 14 (21), 477−485. Perovskite Quantum Dots by Precipitation Temperature. Adv. Sci.
(30) Da Silva, E. L.; Skelton, J. M.; Parker, S. C.; Walsh, A. Phase 2015, 2 (9), 1500194.
Stability and Transformations in the Halide Perovskite CsSnI3. Phys. (48) Yuan, Y.; Liu, Z.; Liu, Z.; Peng, L.; Li, Y.; Tang, A.
Rev. B: Condens. Matter Mater. Phys. 2015, 91 (14), 144107. Photoluminescence and Self-Assembly of Cesium Lead Halide
(31) Nedelcu, G.; Protesescu, L.; Yakunin, S.; Bodnarchuk, M. I.; Perovskite Nanocrystals: Effects of Chain Length of Organic Amines
Grotevent, M. J.; Kovalenko, M. V. Fast Anion-Exchange in Highly and Reaction Temperature. Appl. Surf. Sci. 2017, 405, 280−288.
Luminescent Nanocrystals of Cesium Lead Halide Perovskites (49) Zhang, D.; Eaton, S. W.; Yu, Y.; Dou, L.; Yang, P. Solution-Phase
(CsPbX3, X = Cl, Br, I). Nano Lett. 2015, 15 (8), 5635−5640.
Synthesis of Cesium Lead Halide Perovskite Nanowires. J. Am. Chem.
(32) Huang, H.; Zhao, F.; Liu, L.; Zhang, F.; Wu, X. G.; Shi, L.; Zou,
Soc. 2015, 137 (29), 9230−9233.
B.; Pei, Q.; Zhong, H. Emulsion Synthesis of Size-Tunable
(50) Udayabhaskararao, T.; Kazes, M.; Houben, L.; Lin, H.; Oron, D.
CH3NH3PbBr3 Quantum Dots: An Alternative Route toward Efficient
Nucleation, Growth, and Structural Transformations of Perovskite
Light-Emitting Diodes. ACS Appl. Mater. Interfaces 2015, 7 (51),
Nanocrystals. Chem. Mater. 2017, 29 (3), 1302−1308.
28128−28133.
(51) Zhu, F.; Men, L.; Guo, Y.; Zhu, Q.; Bhattacharjee, U.; Goodwin,
(33) Kulbak, M.; Cahen, D.; Hodes, G. How Important Is the Organic
P. M.; Petrich, J. W.; Smith, E. A.; Vela, J. Shape Evolution and Single
Part of Lead Halide Perovskite Photovoltaic Cells? Efficient CsPbBr3
Particle Luminescence of Organometal Halide Perovskite Nanocryst-
Cells. J. Phys. Chem. Lett. 2015, 6 (13), 2452−2456.
(34) Manser, J. S.; Saidaminov, M. I.; Christians, J. A.; Bakr, O. M.; als. ACS Nano 2015, 9 (3), 2948−2959.
Kamat, P. V. Making and Breaking of Lead Halide Perovskites. Acc. (52) Akkerman, Q. A.; Motti, S. G.; Srimath Kandada, A. R.; Mosconi,
Chem. Res. 2016, 49 (2), 330−338. E.; D’Innocenzo, V.; Bertoni, G.; Marras, S.; Kamino, B. A.; Miranda,
(35) Kulbak, M.; Gupta, S.; Kedem, N.; Levine, I.; Bendikov, T.; L.; De Angelis, F.; et al. Solution Synthesis Approach to Colloidal
Hodes, G.; Cahen, D. Cesium Enhances Long-Term Stability of Lead Cesium Lead Halide Perovskite Nanoplatelets with Monolayer-Level
Bromide Perovskite-Based Solar Cells. J. Phys. Chem. Lett. 2016, 7 (1), Thickness Control. J. Am. Chem. Soc. 2016, 138 (3), 1010−1016.
167−172. (53) Almeida, G.; Goldoni, L.; Akkerman, Q.; Dang, Z.; Khan, A. H.;
(36) Gualdrón-Reyes, A. F.; Yoon, S. J.; Barea, E. M.; Agouram, S.; Marras, S.; Moreels, I.; Manna, L. Role of Acid-Base Equilibria in the
Muñoz-Sanjosé, V.; Meléndez, Á . M.; Niño-Gómez, M. E.; Mora-Seró, Size, Shape, and Phase Control of Cesium Lead Bromide Nanocrystals.
I. Controlling the Phase Segregation in Mixed Halide Perovskites ACS Nano 2018, 12 (2), 1704−1711.
through Nanocrystal Size. ACS Energy Lett. 2019, 4 (1), 54−62. (54) Amgar, D.; Stern, A.; Rotem, D.; Porath, D.; Etgar, L. Tunable
(37) Shi, Z.; Guo, J.; Chen, Y.; Li, Q.; Pan, Y.; Zhang, H.; Xia, Y.; Length and Optical Properties of CsPbX3 (X = Cl, Br, I) Nanowires
Huang, W. Lead-Free Organic−Inorganic Hybrid Perovskites for with a Few Unit Cells. Nano Lett. 2017, 17 (2), 1007−1013.
Photovoltaic Applications: Recent Advances and Perspectives. Adv. (55) Chen, M.; Zou, Y.; Wu, L.; Pan, Q.; Yang, D.; Hu, H.; Tan, Y.;
Mater. 2017, 29 (16), 1605005. Zhong, Q.; Xu, Y.; Liu, H.; et al. Solvothermal Synthesis of High-
(38) Jellicoe, T. C.; Richter, J. M.; Glass, H. F. J.; Tabachnyk, M.; Quality All-Inorganic Cesium Lead Halide Perovskite Nanocrystals:
Brady, R.; Dutton, S. E.; Rao, A.; Friend, R. H.; Credgington, D.; From Nanocube to Ultrathin Nanowire. Adv. Funct. Mater. 2017, 27
Greenham, N. C.; et al. Synthesis and Optical Properties of Lead-Free (23), 1701121.
Cesium Tin Halide Perovskite Nanocrystals. J. Am. Chem. Soc. 2016, (56) Yassitepe, E.; Yang, Z.; Voznyy, O.; Kim, Y.; Walters, G.;
138 (9), 2941−2944. Castañeda, J. A.; Kanjanaboos, P.; Yuan, M.; Gong, X.; Fan, F.; et al.
(39) Liu, H.; Wu, Z.; Shao, J.; Yao, D.; Gao, H.; Liu, Y.; Yu, W.; Zhang, Amine-Free Synthesis of Cesium Lead Halide Perovskite Quantum
H.; Yang, B. CsPbxMn1-XCl3 Perovskite Quantum Dots with High Mn Dots for Efficient Light-Emitting Diodes. Adv. Funct. Mater. 2016, 26
Substitution Ratio. ACS Nano 2017, 11 (2), 2239−2247. (47), 8757−8763.
(40) Liu, W.; Lin, Q.; Li, H.; Wu, K.; Robel, I.; Pietryga, J. M.; Klimov, (57) Pan, A.; He, B.; Fan, X.; Liu, Z.; Urban, J. J.; Alivisatos, A. P.; He,
V. I. Mn2+-Doped Lead Halide Perovskite Nanocrystals with Dual- L.; Liu, Y. Insight into the Ligand-Mediated Synthesis of Colloidal
Color Emission Controlled by Halide Content. J. Am. Chem. Soc. 2016, CsPbBr3 Perovskite Nanocrystals: The Role of Organic Acid, Base, and
138 (45), 14954−14961. Cesium Precursors. ACS Nano 2016, 10 (8), 7943−7954.
(41) Pan, G.; Bai, X.; Yang, D.; Chen, X.; Jing, P.; Qu, S.; Zhang, L.; (58) Wei, S.; Yang, Y.; Kang, X.; Wang, L.; Huang, L.; Pan, D. Room-
Zhou, D.; Zhu, J.; Xu, W.; et al. Doping Lanthanide into Perovskite Temperature and Gram-Scale Synthesis of CsPbX3(X = Cl, Br, I)

11625 DOI: 10.1021/acs.langmuir.9b00855


Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

Perovskite Nanocrystals with 50−85% Photoluminescence Quantum (76) Pileni, M. P. The Role of Soft Colloidal Templates in Controlling
Yields. Chem. Commun. 2016, 52 (45), 7265−7268. the Size and Shape of Inorganic Nanocrystals. Nat. Mater. 2003, 2 (3),
(59) Huang, H.; Raith, J.; Kershaw, S. V.; Kalytchuk, S.; Tomanec, O.; 145−150.
Jing, L.; Susha, A. S.; Zboril, R.; Rogach, A. L. Growth Mechanism of (77) Zhang, J.; Huang, F.; Lin, Z. Progress of Nanocrystalline Growth
Strongly Emitting CH3NH3PbBr3 Perovskite Nanocrystals with a Kinetics Based on Oriented Attachment. Nanoscale 2010, 2 (1), 18−34.
Tunable Bandgap. Nat. Commun. 2017, 8 (1), 996. (78) Bekenstein, Y.; Koscher, B. A.; Eaton, S. W.; Yang, P.; Alivisatos,
(60) Vybornyi, O.; Yakunin, S.; Kovalenko, M. V. Polar-Solvent-Free A. P. Highly Luminescent Colloidal Nanoplates of Perovskite Cesium
Colloidal Synthesis of Highly Luminescent Alkylammonium Lead Lead Halide and Their Oriented Assemblies. J. Am. Chem. Soc. 2015,
Halide Perovskite Nanocrystals. Nanoscale 2016, 8 (12), 6278−6283. 137 (51), 16008−16011.
(61) Shamsi, J.; Dang, Z.; Bianchini, P.; Canale, C.; Stasio, F.; Di; (79) Kim, Y.; Yassitepe, E.; Voznyy, O.; Comin, R.; Walters, G.; Gong,
Brescia, R.; Prato, M.; Manna, L. Colloidal Synthesis of Quantum X.; Kanjanaboos, P.; Nogueira, A. F.; Sargent, E. H. Efficient
Confined Single Crystal CsPbBr3 Nanosheets with Lateral Size Control Luminescence from Perovskite Quantum Dot Solids. ACS Appl.
up to the Micrometer Range. J. Am. Chem. Soc. 2016, 138 (23), 7240− Mater. Interfaces 2015, 7 (45), 25007−25013.
7243. (80) De Roo, J.; Ibáñez, M.; Geiregat, P.; Nedelcu, G.; Walravens, W.;
(62) Shaw, D. J. Introduction to Colloid and Surface Chemistry, 4th ed.; Maes, J.; Martins, J. C.; Van Driessche, I.; Kovalenko, M. V.; Hens, Z.
Butterworth-Heinemann, 1992. Highly Dynamic Ligand Binding and Light Absorption Coefficient of
(63) Li, X.; Wu, Y.; Zhang, S.; Cai, B.; Gu, Y.; Song, J.; Zeng, H. Cesium Lead Bromide Perovskite Nanocrystals. ACS Nano 2016, 10
CsPbX3 Quantum Dots for Lighting and Displays: Room-Temperature (2), 2071−2081.
Synthesis, Photoluminescence Superiorities, Underlying Origins and (81) Krieg, F.; Ochsenbein, S. T.; Yakunin, S.; ten Brinck, S.; Aellen,
White Light-Emitting Diodes. Adv. Funct. Mater. 2016, 26 (15), 2435− P.; Süess, A.; Clerc, B.; Guggisberg, D.; Nazarenko, O.; Shynkarenko,
2445. Y.; et al. Colloidal CsPbX 3 (X = Cl, Br, I) Nanocrystals 2.0:
(64) Murray, C. B.; Norris, D. J.; Bawendi, M. G. Synthesis and Zwitterionic Capping Ligands for Improved Durability and Stability.
Characterization of Nearly Monodisperse CdE (E = S, Se, Te) ACS Energy Lett. 2018, 3 (3), 641−646.
Semiconductor Nanocrystallites. J. Am. Chem. Soc. 1993, 115 (19), (82) Cho, J.; Jin, H.; Sellers, D. G.; Watson, D. F.; Son, D. H.;
8706−8715. Banerjee, S. Influence of Ligand Shell Ordering on Dimensional
(65) Ahlawat, P.; Piaggi, P.; Grätzel, M.; Parrinello, M.; Röthlisberger, Confinement of Cesium Lead Bromide (CsPbBr3) Perovskite
U. Atomistic Mechanism of the Nucleation of Methylammonium Lead Nanoplatelets. J. Mater. Chem. C 2017, 5 (34), 8810−8818.
Iodide Perovskite from Solution. Nat. Commun. 2019. (83) Di Stasio, F.; Christodoulou, S.; Huo, N.; Konstantatos, G. Near-
(66) Li, Y.; Huang, H.; Xiong, Y.; Kershaw, S. V.; Rogach, A. L. Unity Photoluminescence Quantum Yield in CsPbBr3 Nanocrystal
Revealing the Formation Mechanism of CsPbBr3 Perovskite Nano- Solid-State Films via Postsynthesis Treatment with Lead Bromide.
crystals Produced via a Slowed-Down Microwave-Assisted Synthesis. Chem. Mater. 2017, 29 (18), 7663−7667.
Angew. Chem., Int. Ed. 2018, 57 (20), 5833−5837. (84) Aharon, S.; Etgar, L. Two Dimensional Organometal Halide
(67) Ko, Y. H.; Prabhakaran, P.; Jalalah, M.; Lee, S. J.; Lee, K. S.; Park, Perovskite Nanorods with Tunable Optical Properties. Nano Lett.
J. G. Correlating Nano Black Spots and Optical Stability in Mixed 2016, 16 (5), 3230−3235.
Halide Perovskite Quantum Dots. J. Mater. Chem. C 2018, 6 (29), (85) Cho, J.; Banerjee, S. Ligand-Directed Stabilization of Ternary
7803−7813. Phases: Synthetic Control of Structural Dimensionality in Solution-
(68) Protesescu, L.; Yakunin, S.; Bodnarchuk, M. I.; Bertolotti, F.; Grown Cesium Lead Bromide Nanocrystals. Chem. Mater. 2018, 30
Masciocchi, N.; Guagliardi, A.; Kovalenko, M. V. Monodisperse (17), 6144−6155.
Formamidinium Lead Bromide Nanocrystals with Bright and Stable (86) Aamir, M.; Khan, M. D.; Sher, M.; Malik, M. A.; Akhtar, J.;
Green Photoluminescence. J. Am. Chem. Soc. 2016, 138 (43), 14202− Revaprasadu, N. Synthesis of Hybrid to Inorganic Quasi 2D-Layered
14205. Perovskite Nanoparticles. ChemistrySelect 2017, 2 (20), 5595−5599.
(69) Wu, L.; Zhong, Q.; Yang, D.; Chen, M.; Hu, H.; Pan, Q.; Liu, H.; (87) Gonzalez-Carrero, S.; Galian, R. E.; Pérez-Prieto, J. Maximizing
Cao, M.; Xu, Y.; Sun, B.; et al. Improving the Stability and Size the Emissive Properties of CH3NH3PbBr3 Perovskite Nanoparticles. J.
Tunability of Cesium Lead Halide Perovskite Nanocrystals Using Mater. Chem. A 2015, 3 (17), 9187−9193.
Trioctylphosphine Oxide as the Capping Ligand. Langmuir 2017, 33 (88) Zhang, F.; Zhong, H.; Chen, C.; Wu, X.; Hu, X.; Huang, H.; Han,
(44), 12689−12696. J.; Zou, B.; Dong, Y. Brightly Luminescent and Color-Tunable
(70) Tong, Y.; Bladt, E.; Aygüler, M. F.; Manzi, A.; Milowska, K. Z.; Colloidal CH3NH3PbX3 (X = Br, I, Cl) Quantum Dots: Potential
Hintermayr, V. A.; Docampo, P.; Bals, S.; Urban, A. S.; Polavarapu, L.; Alternatives for Display Technology. ACS Nano 2015, 9 (4), 4533−
et al. Highly Luminescent Cesium Lead Halide Perovskite Nanocrystals 4542.
with Tunable Composition and Thickness by Ultrasonication. Angew. (89) Leng, M.; Chen, Z.; Yang, Y.; Li, Z.; Zeng, K.; Li, K.; Niu, G.; He,
Chem., Int. Ed. 2016, 55 (44), 13887−13892. Y.; Zhou, Q.; Tang, J. Lead-Free, Blue Emitting Bismuth Halide
(71) Pan, Q.; Hu, H.; Zou, Y.; Chen, M.; Wu, L.; Yang, D.; Yuan, X.; Perovskite Quantum Dots. Angew. Chem., Int. Ed. 2016, 55 (48),
Fan, J.; Sun, B.; Zhang, Q. Microwave-Assisted Synthesis of High- 15012−15016.
Quality All-Inorganic CsPbX3 (X = Cl, Br, I) Perovskite Nanocrystals (90) Yang, B.; Chen, J.; Hong, F.; Mao, X.; Zheng, K.; Yang, S.; Li, Y.;
and the Application in Light Emitting Diode. J. Mater. Chem. C 2017, 5 Pullerits, T.; Deng, W.; Han, K. Lead-Free, Air-Stable All-Inorganic
(42), 10947−10954. Cesium Bismuth Halide Perovskite Nanocrystals. Angew. Chem., Int. Ed.
(72) Bratsch, S. G. Standard Electrode Potentials and Temperature 2017, 56 (41), 12471−12475.
Coefficients in Water at 298.15 K. J. Phys. Chem. Ref. Data 1989, 18 (1), (91) Leng, M.; Yang, Y.; Zeng, K.; Chen, Z.; Tan, Z.; Li, S.; Li, J.; Xu,
1−21. B.; Li, D.; Hautzinger, M. P.; et al. All-Inorganic Bismuth-Based
(73) Sichert, J. A.; Tong, Y.; Mutz, N.; Vollmer, M.; Fischer, S.; Perovskite Quantum Dots with Bright Blue Photoluminescence and
Milowska, K. Z.; García Cortadella, R.; Nickel, B.; Cardenas-Daw, C.; Excellent Stability. Adv. Funct. Mater. 2018, 28 (1), 1704446.
Stolarczyk, J. K.; et al. Quantum Size Effect in Organometal Halide (92) Zhang, J.; Yang, Y.; Deng, H.; Farooq, U.; Yang, X.; Khan, J.;
Perovskite Nanoplatelets. Nano Lett. 2015, 15 (10), 6521−6527. Tang, J.; Song, H. High Quantum Yield Blue Emission from Lead-Free
(74) Koolyk, M.; Amgar, D.; Aharon, S.; Etgar, L. Kinetics of Cesium Inorganic Antimony Halide Perovskite Colloidal Quantum Dots. ACS
Lead Halide Perovskite Nanoparticle Growth; Focusing and de- Nano 2017, 11 (9), 9294−9302.
Focusing of Size Distribution. Nanoscale 2016, 8 (12), 6403−6409. (93) Xie, J. L.; Huang, Z. Q.; Wang, B.; Chen, W. J.; Lu, W. X.; Liu, X.;
(75) Kirakosyan, A.; Yun, S.; Yoon, S.-G.; Choi, J. Surface Engineering Song, J. L. New Lead-Free Perovskite Rb7Bi3Cl16 Nanocrystals with
for Improved Stability of CH3NH3PbBr3 Perovskite Nanocrystals. Blue Luminescence and Excellent Moisture-Stability. Nanoscale 2019,
Nanoscale 2018, 10 (4), 1885−1891. 11 (14), 6719−6726.

11626 DOI: 10.1021/acs.langmuir.9b00855


Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

(94) Yang, S.; Zheng, Y. C.; Hou, Y.; Chen, X.; Chen, Y.; Wang, Y.; Silver-Bismuth Halide Compositions. Nano Lett. 2018, 18 (6), 3502−
Zhao, H.; Yang, H. G. Formation Mechanism of Freestanding 3508.
CH3NH3PbI3 Functional Crystals: In Situ Transformation vs (112) Tsiwah, E. A.; Ding, Y.; Li, Z.; Zhao, Z.; Wang, M.; Hu, C.; Liu,
Dissolution−Crystallization. Chem. Mater. 2014, 26 (23), 6705−6710. X.; Sun, C.; Zhao, X.; Xie, Y. One-Pot Scalable Synthesis of All-
(95) Liu, P.; Chen, W.; Wang, W.; Xu, B.; Wu, D.; Hao, J.; Cao, W.; Inorganic Perovskite Nanocrystals with Tunable Morphology,
Fang, F.; Zeng, Y.; Pan, R.; et al. Halide-Rich Synthesized Cesium Lead Composition and Photoluminescence. CrystEngComm 2017, 19 (46),
Bromide Perovskite Nanocrystals for Light-Emitting Diodes with 7041−7049.
Improved Performance. Chem. Mater. 2017, 29 (12), 5168−5173. (113) Zhai, W.; Lin, J.; Li, Q.; Zheng, K.; Huang, Y.; Yao, Y.; He, X.;
(96) Woo, J. Y.; Kim, Y.; Bae, J.; Kim, T. G.; Kim, J. W.; Lee, D. C.; Li, L.; Yu, C.; Liu, C.; et al. Solvothermal Synthesis of Ultrathin Cesium
Jeong, S. Highly Stable Cesium Lead Halide Perovskite Nanocrystals Lead Halide Perovskite Nanoplatelets with Tunable Lateral Sizes and
through in Situ Lead Halide Inorganic Passivation. Chem. Mater. 2017, Their Reversible Transformation into Cs4PbBr6 Nanocrystals. Chem.
29 (17), 7088−7092. Mater. 2018, 30 (11), 3714−3721.
(97) Linaburg, M. R.; McClure, E. T.; Majher, J. D.; Woodward, P. M. (114) Chen, L.-J.; Lee, C.-R.; Chuang, Y.-J.; Wu, Z.-H.; Chen, C.
Cs1-XRbxPbCl3 and Cs1-XRbxPbBr3 Solid Solutions: Understanding Synthesis and Optical Properties of Lead-Free Cesium Tin Halide
Octahedral Tilting in Lead Halide Perovskites. Chem. Mater. 2017, 29 Perovskite Quantum Rods with High-Performance Solar Cell
(8), 3507−3514. Application. J. Phys. Chem. Lett. 2016, 7 (24), 5028−5035.
(98) Wang, C.; Zhang, Y.; Wang, A.; Wang, Q.; Tang, H.; Shen, W.; Li, (115) Long, Z.; Ren, H.; Sun, J.; Ouyang, J.; Na, N. High-Throughput
Z.; Deng, Z. Controlled Synthesis of Composition Tunable and Tunable Synthesis of Colloidal CsPbX3 Perovskite Nanocrystals in
Formamidinium Cesium Double Cation Lead Halide Perovskite a Heterogeneous System by Microwave Irradiation. Chem. Commun.
Nanowires and Nanosheets with Improved Stability. Chem. Mater. 2017, 53 (71), 9914−9917.
2017, 29 (5), 2157−2166. (116) Liu, W.; Zheng, J.; Cao, S.; Wang, L.; Gao, F.; Chou, K. C.; Hou,
(99) Lim, D. H.; Ramasamy, P.; Kwak, D. H.; Lee, J. S. Solution-Phase X.; Yang, W. General Strategy for Rapid Production of Low-
Synthesis of Rubidium Lead Iodide Orthorhombic Perovskite Nano- Dimensional All-Inorganic CsPbBr3 Perovskite Nanocrystals with
wires. Nanotechnology 2017, 28 (25), 255601. Controlled Dimensionalities and Sizes. Inorg. Chem. 2018, 57 (3),
(100) Amgar, D.; Binyamin, T.; Uvarov, V.; Etgar, L. Near Ultra- 1598−1603.
Violet to Mid-Visible Band Gap Tuning of Mixed Cation RbXCs1- (117) van Embden, J.; Chesman, A. S. R.; Jasieniak, J. J. The Heat-Up
XPbX3 (X = Cl or Br) Perovskite Nanoparticles. Nanoscale 2018, 10 Synthesis of Colloidal Nanocrystals. Chem. Mater. 2015, 27 (7), 2246−
(13), 6060−6068. 2285.
(101) Parobek, D.; Roman, B. J.; Dong, Y.; Jin, H.; Lee, E.; Sheldon, (118) Huang, H.; Xue, Q.; Chen, B.; Xiong, Y.; Schneider, J.; Zhi, C.;
M.; Son, D. H. Exciton-to-Dopant Energy Transfer in Mn-Doped Zhong, H.; Rogach, A. L. Top-Down Fabrication of Stable
Cesium Lead Halide Perovskite Nanocrystals. Nano Lett. 2016, 16 Methylammonium Lead Halide Perovskite Nanocrystals by Employing
(12), 7376−7380. a Mixture of Ligands as Coordinating Solvents. Angew. Chem., Int. Ed.
(102) Zou, S.; Liu, Y.; Li, J.; Liu, C.; Feng, R.; Jiang, F.; Li, Y.; Song, J.; 2017, 56 (32), 9571−9576.
Zeng, H.; Hong, M.; et al. Stabilizing Cesium Lead Halide Perovskite (119) Tong, Y.; Yao, E.; Manzi, A.; Bladt, E.; Wang, K.; Döblinger, M.;
Lattice through Mn(II) Substitution for Air-Stable Light-Emitting Bals, S.; Mü ller-buschbaum, P.; Urban, A. S.; Polavarapu, L.
Diodes. J. Am. Chem. Soc. 2017, 139 (33), 11443−11450. Spontaneous Self-Assembly of Perovskite Nanocrystals into Electroni-
(103) Zhou, D.; Liu, D.; Pan, G.; Chen, X.; Li, D.; Xu, W.; Bai, X.; cally Coupled Supercrystals: Toward Filling the Green Gap. Adv. Mater.
Song, H. Cerium and Ytterbium Codoped Halide Perovskite Quantum 2018, 30, 1801117.
Dots: A Novel and Efficient Downconverter for Improving the (120) Li, J.; Xu, L.; Wang, T.; Song, J.; Chen, J.; Xue, J.; Dong, Y.; Cai,
Performance of Silicon Solar Cells. Adv. Mater. 2017, 29 (42), 1704149. B.; Shan, Q.; Han, B.; et al. 50-Fold EQE Improvement up to 6.27% of
(104) Shen, W.; Wang, A.; Yang, M.; Deng, Z.; Sun, S.; Yan, X.; Wang, Solution-Processed All-Inorganic Perovskite CsPbBr3 QLEDs via
P.; Pan, X.; Zhang, M. Controlled Synthesis of Lead-Free and Stable Surface Ligand Density Control. Adv. Mater. 2017, 29 (5), 1603885.
Perovskite Derivative Cs2SnI6 Nanocrystals via a Facile Hot-Injection (121) Ripka, E. G.; Deschene, C. R.; Franck, J. M.; Bae, I. T.; Maye, M.
Process. Chem. Mater. 2016, 28 (22), 8132−8140. M. Understanding the Surface Properties of Halide Exchanged Cesium
(105) Zhang, Y.; Yin, J.; Parida, M. R.; Ahmed, G. H.; Pan, J.; Bakr, O. Lead Halide Nanoparticles. Langmuir 2018, 34 (37), 11139−11146.
M.; Bredas, J.-L.; Mohammed, O. F. Direct-Indirect Nature of the (122) Ravi, V. K.; Santra, P. K.; Joshi, N.; Chugh, J.; Singh, S. K.;
Bandgap in Lead-Free Perovskite Nanocrystals. J. Phys. Chem. Lett. Rensmo, H.; Ghosh, P.; Nag, A. Origin of the Substitution Mechanism
2017, 8, 3173−3177. for the Binding of Organic Ligands on the Surface of CsPbBr3
(106) Acharya, S.; Sain, S.; Kumar, G. S.; Pradhan, B.; Dalui, A.; Perovskite Nanocubes. J. Phys. Chem. Lett. 2017, 8 (20), 4988−4994.
Pradhan, S. K.; Ghorai, U. K. Size Tunable Cesium Antimony Chloride (123) Koscher, B. A.; Swabeck, J. K.; Bronstein, N. D.; Alivisatos, A. P.
Perovskite Nanowires and Nanorods. Chem. Mater. 2018, 30 (6), Essentially Trap-Free CsPbBr3 Colloidal Nanocrystals by Postsyn-
2135−2142. thetic Thiocyanate Surface Treatment. J. Am. Chem. Soc. 2017, 139
(107) Pal, J.; Manna, S.; Mondal, A.; Das, S.; Adarsh, K. V.; Nag, A. (19), 6566−6569.
Colloidal Synthesis and Photophysics of M3Sb2I9 (M = Cs and Rb) (124) Ahmed, T.; Seth, S.; Samanta, A. Boosting the Photo-
Nanocrystals: Lead-Free Perovskites. Angew. Chem., Int. Ed. 2017, 56 luminescence of CsPbX3 (X = Cl, Br, I) Perovskite Nanocrystals
(45), 14187−14191. Covering a Wide Wavelength Range by Post-Synthetic Treatment with
(108) Ravi, V. K.; Singhal, N.; Nag, A. Initiation and Future Prospects Tetrafluoroborate Salts. Chem. Mater. 2018, 30 (11), 3633−3637.
of Colloidal Metal Halide Double-Perovskite Nanocrystals: Cs2Ag- (125) Mir, W. J.; Swarnkar, A.; Nag, A. Postsynthesis Mn-Doping in
BiX6 (X = Cl, Br, I). J. Mater. Chem. A 2018, 6 (44), 21666−21675. CsPbI3 Nanocrystals to Stabilize the Black Perovskite Phase. Nanoscale
(109) Zhou, L.; Xu, Y. F.; Chen, B. X.; Kuang, D.-B.; Su, C. Y. 2019, 11 (10), 4278−4286.
Synthesis and Photocatalytic Application of Stable Lead-Free (126) Bodnarchuk, M. I.; Boehme, S. C.; Ten Brinck, S.; Bernasconi,
Cs2AgBiBr6 Perovskite Nanocrystals. Small 2018, 14 (11), 1703762. C.; Shynkarenko, Y.; Krieg, F.; Widmer, R.; Aeschlimann, B.; Günther,
(110) Creutz, S. E.; Crites, E. N.; De Siena, M. C.; Gamelin, D. R. D.; Kovalenko, M. V.; et al. Rationalizing and Controlling the Surface
Colloidal Nanocrystals of Lead-Free Double-Perovskite (Elpasolite) Structure and Electronic Passivation of Cesium Lead Halide Nano-
Semiconductors: Synthesis and Anion Exchange to Access New crystals. ACS Energy Lett. 2019, 4 (1), 63−74.
Materials. Nano Lett. 2018, 18 (2), 1118−1123. (127) Doane, T. L.; Ryan, K. L.; Pathade, L.; Cruz, K. J.; Zang, H.;
(111) Bekenstein, Y.; Dahl, J. C.; Huang, J.; Osowiecki, W. T.; Cotlet, M.; Maye, M. M. Using Perovskite Nanoparticles as Halide
Swabeck, J. K.; Chan, E. M.; Yang, P.; Alivisatos, A. P. The Making and Reservoirs in Catalysis and as Spectrochemical Probes of Ions in
Breaking of Lead-Free Double Perovskite Nanocrystals of Cesium Solution. ACS Nano 2016, 10 (6), 5864−5872.

11627 DOI: 10.1021/acs.langmuir.9b00855


Langmuir 2019, 35, 11609−11628
Langmuir Invited Feature Article

(128) Zhang, D.; Yang, Y.; Bekenstein, Y.; Yu, Y.; Gibson, N. A.;
Wong, A. B.; Eaton, S. W.; Kornienko, N.; Kong, Q.; Lai, M.; et al.
Synthesis of Composition Tunable and Highly Luminescent Cesium
Lead Halide Nanowires through Anion-Exchange Reactions. J. Am.
Chem. Soc. 2016, 138 (23), 7236−7239.
(129) Akkerman, Q. A.; D’Innocenzo, V.; Accornero, S.; Scarpellini,
A.; Petrozza, A.; Prato, M.; Manna, L. Tuning the Optical Properties of
Cesium Lead Halide Perovskite Nanocrystals by Anion Exchange
Reactions. J. Am. Chem. Soc. 2015, 137 (32), 10276−10281.
(130) van der Stam, W.; Geuchies, J. J.; Altantzis, T.; Van Den Bos, K.
H. W.; Meeldijk, J. D.; Van Aert, S.; Bals, S.; Vanmaekelbergh, D.; De
Mello Donega, C. Highly Emissive Divalent-Ion-Doped Colloidal
CsPb1-XMxBr3 Perovskite Nanocrystals through Cation Exchange. J.
Am. Chem. Soc. 2017, 139 (11), 4087−4097.
(131) Pan, D.; Fu, Y.; Chen, J.; Czech, K. J.; Wright, J. C.; Jin, S.
Visualization and Studies of Ion-Diffusion Kinetics in Cesium Lead
Bromide Perovskite Nanowires. Nano Lett. 2018, 18 (3), 1807−1813.
(132) Mondal, N.; De, A.; Samanta, A. Achieving Near-Unity
Photoluminescence Efficiency for Blue-Violet-Emitting Perovskite
Nanocrystals. ACS Energy Lett. 2019, 4 (1), 32−39.
(133) Huang, G.; Wang, C.; Xu, S.; Zong, S.; Lu, J.; Wang, Z.; Lu, C.;
Cui, Y. Postsynthetic Doping of MnCl2Molecules into Preformed
CsPbBr3 Perovskite Nanocrystals via a Halide Exchange-Driven Cation
Exchange. Adv. Mater. 2017, 29 (29), 1700095.
(134) Mir, W. J.; Mahor, Y.; Lohar, A.; Jagadeeswararao, M.; Das, S.;
Mahamuni, S.; Nag, A. Postsynthesis Doping of Mn and Yb into
CsPbX3 (X = Cl, Br, or I) Perovskite Nanocrystals for Downconversion
Emission. Chem. Mater. 2018, 30 (22), 8170−8178.

11628 DOI: 10.1021/acs.langmuir.9b00855


Langmuir 2019, 35, 11609−11628

You might also like