You are on page 1of 10

pubs.acs.

org/cm Article

Origin of Defects and Positron Annihilation in Hybrid and All-


Inorganic Perovskites
Artem Musiienko,* Jakub Č ížek, Hassan Elhadidy, Petr Praus, Kate Higgins, Bogdan Dryzhakov,
Andrii Kanak, Franck Sureau, Jindrǐ ch Pipek, Eduard Belas, Marián Betušiak, Mykola Brynza, Eric Lukosi,
Bin Hu, and Mahshid Ahmadi*
Cite This: Chem. Mater. 2022, 34, 297−306 Read Online

ACCESS *
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Metrics & More Article Recommendations sı Supporting Information


Downloaded via CENTRAL SOUTH UNIV on February 9, 2023 at 03:26:43 (UTC).

ABSTRACT: Emerging metal-halide perovskites (MHPs) have shown advanced


charge transport properties suitable for application in solar cells, photodetectors,
and many more. While the past decade witnessed tremendous progress in MHPs,
very little is known about the origin of defects and their effect on carrier lifetime.
In this study, we compare hybrid and all-inorganic MHPs prepared by inverse
temperature solution and high-temperature melt growth to explore the influence
of material preparation on the formation of defects. The presence of a low
concentration of vacancies was shown in all MHPs regardless of their synthesis
method demonstrated by the interaction of positron particles with vacancies and
lattices of MHPs and explained by ab initio simulation of positron annihilation.
We combined the Raman, Fourier transform infrared (FTIR), and positron
annihilation spectroscopy methods to establish the nature of imperfections in MHPs grown using different methods. Our Raman and
FTIR results reveal that only the solution-grown crystals are prone to the incorporation of a solvent in bulk during synthesis. In vast
majority of studies, the charge carrier lifetime is explored using photoluminescence (PL) spectroscopy as it is a readily available
method. However, PL is very sensitive to both bulk and surface recombination phenomena. Combining current waveform time-of-
flight and time-resolved photoluminescence spectroscopy methods, the bulk recombination differs by a factor of 2 from crystals
grown by solution versus high-temperature melt. The results propose that solvent trapping matters, not intrinsic defects. The study
also suggests potential pathways for further improvement of hybrid and all-inorganic MHPs.

■ INTRODUCTION
Lead halide perovskites with the formula APbX3, where A+ is
method is desirable. Over the past few years, the solution-
based growth of CsPbBr3 has been demonstrated11,17 through
either cesium (Cs), methylammonium (MA), or formamidi- the utilization of growth methods such as the antisolvent
vapor-assisted12,18 and the inverse temperature crystallization
nium and X− is either chlorine (Cl), bromine (Br), or iodine
techniques.9,19,20 Despite progress in solution-based growth of
(I−), have emerged as promising candidates for ionizing
lead halide perovskite single crystals, the crystallinity,
radiation detectors because of their high attenuation
dimensions, and charge transport properties of melt-grown
coefficient, tunable bandgap,1−3 large mobility-lifetime (μτ)
single crystals are still superior to date.
product 1,2,4 and low charge trap density and defect
It is well established that defects have a critical effect on the
tolerance.2,5−8 In particular, the all-inorganic cesium lead
performance of semiconductor devices. The crucial charge
bromide (CsPbBr3) perovskite is advantageous over its
transport properties such as carrier mobility and lifetime are
organic−inorganic perovskite counterparts because of its
greatly affected by the concentration of defects.8 The effect of
long-term chemical and structural stability.8−12 Recently,
defects on charge transport properties of solution-grown
more researchers have shown that CsPbBr3 can achieve good
perovskites was broadly discussed in our recent studies.21,22
resolution detection of ionizing radiation13,14 in addition to
In this study, we explore the electrical structure, charge
their excellent potential for application in tandem solar cells
transport, and defect chemistry in solution- and melt-grown
and LED technology.15,16
Until recently, the growth of large-size, high-quality CsPbBr3
single crystals required high-temperature growth from melts Received: October 13, 2021
using the Bridgman melt growth method.13 As this melt Revised: December 2, 2021
growth method requires temperatures over 500 °C and Published: December 20, 2021
multiple purification steps to purify the raw materials, this
considerably increases the cost of device fabrication.14
Therefore, the need for an inexpensive solution-based growth

© 2021 American Chemical Society https://doi.org/10.1021/acs.chemmater.1c03540


297 Chem. Mater. 2022, 34, 297−306
Chemistry of Materials pubs.acs.org/cm Article

Figure 1. Positron lifetime spectra measured in (a) melt-grown CsPbBr3 and (b) solution-grown MAPbBr3, respectively. Experimental points are
plotted by open circles. The solid line shows the calculated model function. Residuals, that is, differences between the experimental spectrum and
model curve expressed in units of one standard deviation σ, are plotted in upper panels. For comparison, the panels (c,d) show the positron lifetime
spectrum measured in La-doped polycrystalline SrTiO3 perovskite Sr0.95La0.05TiO3 containing open volume point defects. As demonstrated in panel
(c), the positron lifetime spectrum of the Sr0.95La0.05TiO3 sample cannot be described by single component because there are two different positron
states in the sample. Two exponential components (red and green lines) are necessary for proper description of the positron lifetime spectrum of
the Sr0.95La0.05TiO3 sample as indicated in panel (d) by the dashed orange line.

lead halide perovskite single crystals to reveal pathways to lead halide perovskites. Dhar et al. demonstrated the utilization
efficient and cheap perovskite semiconductor devices. We of PAS in combination with Doppler broadening techniques to
identify the key factors limiting the operation efficiency of reveal the presence of MA defects in the MAPbI3 crystal
solution-based perovskites for radiation sensors by comparing lattice.28 Single-crystal MAPbI3 contains fewer defects than its
the electrical, charge transport, and defect properties of polycrystalline counterpart but has an equal probability of
CsPbBr3 and MAPbBr3 single crystals grown from the melt developing both cationic and anionic (halide) vacancies,
and solution processing. Special attention is given to the study whereas the polycrystalline sample contains mainly cationic
of defects’ nature that is crucial for further optimization of all- vacancies in its lattice. PAS is yet to be used to characterize the
inorganic and hybrid perovskite optoelectronic devices. defects of other lead halide perovskites including CsPbBr3 and
Positron annihilation spectroscopy (PAS) is commonly used MAPbBr3 single crystals.


to study defects and their chemical nature in various
semiconductors.23−26 In conventional PAS, the material is RESULTS
bombarded with positrons. These positrons are then
annihilated by the material’s electrons, producing two 511 Positron Interaction with Metal-Halide Perovskites
keV gamma rays. The positron annihilation rate is determined (MHPs). To find the possible origin of intrinsic point defects,
by the local electron density at the positron annihilation site.27 we study positron interaction with defects in melt-grown
The inverse of positron annihilation rate is called positron CsPbBr3 and solution-grown MAPbBr3 by positron lifetime
lifetime. Because the electron density of a defect site is spectroscopy on a picosecond (ps) timescale. In this method,
comparatively less than the electron density of the bulk the positron implanted into the sample is thermalized within a
material, the positron is trapped in a defect site for a longer few ps (i.e., its energy is quickly decreased down to thermal
time. Another positron annihilation technique, coincidence energy, kT). The thermalized positron is delocalized in the
Doppler broadening (CDB) spectroscopy of the annihilation lattice as a modulated Bloch-like wave. Open volume defects,
gamma rays, provides information about the electron such as vacancies, represent potential trapping sites for
momentum distribution in a material. Together, these positrons. Hence, if a sample contains defects, the delocalized
techniques are incredibly powerful in characterizing the positron is trapped in these defects. Because electron density in
chemical nature of defects in a material. To date, very few open volume defects is reduced, the positron trapped at defect
studies have used PAS to characterize the nature of defects in lives longer than the delocalized positron. Each positron state
298 https://doi.org/10.1021/acs.chemmater.1c03540
Chem. Mater. 2022, 34, 297−306
Chemistry of Materials pubs.acs.org/cm Article

Figure 2. Ab initio calculations of positron annihilation observables. Calculated electron density in the (001) plane for (a) CsPbBr3 lattice with no
defects, (b) Cs vacancy, (c) Br vacancy, and (d) Pb vacancy; (e−h) show corresponding positron density obtained from the ab initio calculations.
(i) Experimental CDB ratio curves (with respect to Al reference) for CsPbBr3 and MAPbBr3 samples grown by different methods. (j) Calculated
CDB ratio curves (with respect to pure Al reference) for free positrons in CsPbBr3 and MAPbBr3.

(e.g., positron delocalized in the lattice and positron trapped at properly described by two exponential components indicating
various kinds of vacancies) is represented by an exponential that there are two different positron states in the sample.
component in the positron lifetime spectrum characterized by Hence, in contrast to polycrystalline perovskites CsPbBr3
its lifetime and intensity. The lifetime spectra of positrons in and MAPbBr3 single crystalline perovskites exhibit single-
CsPbBr3 and MAPbBr3, perovskite single crystals are shown in component positron lifetime spectra. As will be shown in the
Figure 1. Both CsPbBr3 and MAPbBr3 perovskites exhibit a next section, the lifetime of this component agrees well with
single component, which deviates from multicomponent the calculated lifetime of free positrons in the perfect lattice.
spectra measured in polycrystalline perovskites.28 As the These results indicate that both CsPbBr3 and MAPbBr3 single
additional example of positron annihilation in other materials, crystals contain a very low concentration of open volume
Figure 1c,d shows the positron lifetime spectrum of La-doped defects, and virtually, all positrons are annihilated in the
polycrystalline SrTiO3 perovskite (Sr0.95La0.05TiO3) measured delocalized state. Hence, one can conclude that growth from
on the same PAS setup. From inspection of Figure 1c one can the melt as well as growth by the solution processing produces
realize that the positron lifetime spectrum of Sr0.95La0.05TiO3 crystals of comparable quality. Positron lifetimes of (349.6 ±
corresponds to multiexponential decay and cannot be 0.3) ps and (343.2 ± 0.3) ps were obtained from fitting of
described by a single component. As demonstrated in Figure spectra for melt-grown CsPbBr3 and solution-grown MAPbBr3
1d, the positron lifetime spectrum of Sr0.95La0.05TiO3 can be single crystals, respectively.
299 https://doi.org/10.1021/acs.chemmater.1c03540
Chem. Mater. 2022, 34, 297−306
Chemistry of Materials pubs.acs.org/cm Article

Table 1. Results of Ab Initio Calculations of Positron Lifetimesa


CsPbBr3 τ (ps) EB (eV) MAPbBr3 τ (ps) EB (eV)
experimental PAS lifetime 349.6 ± 0.3 experimental PAS lifetime 343.2 ± 0.3
bulk 355.1 bulk 333.4
Cs vacancy 364.8 0.329 MA-vacancy 357.9 0.055
Pb vacancy 371.9 0.023 Pb vacancy 347.0 0.010
Br vacancy 394.8 0.066 Br vacancy 342.9 0.011
a
Calculated positron lifetimes for a perfect lattice (bulk) and various types of vacancies. Estimated uncertainty of calculated lifetimes is 5 ps.
Calculated binding energy (EB) of positron to vacancies is shown in the table as well. Lifetimes measured in the experiment are listed in the first
row of the table.

Figure 3. Defect nature detected by Raman and FTIR. Raman spectra for solution-grown MAPbBr3, solution-grown CsPbBr3, melt-grown CsPbBr3
single crystals, and dimethyl sulfoxide (DMSO) solution for (a) low wavenumbers and (b) high wavenumbers. (c) FTIR results for the same
crystals.

Ab Initio Calculation of Positron Annihilation Note that positron binding energy to a vacancy is the
Parameters. To explain the positron lifetime decay, we difference between the ground-state energy of the delocalized
performed ab initio calculations of lifetimes of positrons (free) positron and the ground-state energy of the positron
annihilated in various states based on local density trapped in the vacancy. The calculated bulk lifetimes, that is,
approximation.29 More details of this calculation can be lifetimes of free positrons in a perfect lattice, are in reasonable
found in the Supporting Information. The calculated electron agreement with lifetime measured in the experiment. On the
density found in the perfect CsPbBr3 lattice and at different other hand, calculated lifetimes of positrons trapped in
kinds of vacancies is shown in (Figure 2a−d). The calculated vacancies are significantly higher than that observed in the
positron density in the (001) plane for the perfect CsPbBr3 experiment.
lattice and various kinds of vacancies is shown in Figure 2e−h. We further considered the annihilation of the positron in a
Results of ab initio calculations of positron lifetimes for various free state, which is a lattice without defects. The calculated
positron states in CsPbBr3 and MAPbBr3 (according to eq S1 bulk positron lifetimes according to eq S1 and S2 and Figure
and S2 and Figure 2) are summarized in Table 1. Estimated 2a−h for a perfect lattice of CsPbBr3 and MAPbBr3 are around
uncertainty of calculated lifetimes is around 5 ps. A cation 355.1 and 333.4 ps, respectively. The difference in positron
vacancy (Cs vacancy in CsPbBr3 and MA-vacancy in lifetime (several tens of ps) in CsPbBr3 and MAPbBr3 crystals
MAPbBr3) represents a deep positron trap in both compounds. can be explained by the interaction of positrons with different
Pb and Br vacancies are capable of positron trapping as well, lattices. Because of different cations (MA and Cs), the lattice
but the positron binding energy to these vacancies is smaller. has slightly different cation radii and lattice constant. As a
300 https://doi.org/10.1021/acs.chemmater.1c03540
Chem. Mater. 2022, 34, 297−306
Chemistry of Materials pubs.acs.org/cm Article

result, positron trapping and positron lifetimes are slightly cations,33,34 respectively. Noticeably, we also see a shoulder
different for inorganic and hybrid lead halide perovskites, peak around 70 cm−1 for both the solution-grown and melt-
which was confirmed by ab initio calculations. grown CsPbBr3. The peak has also been observed in the
Ab Initio Calculation of Momentum Distribution of Raman spectrum of melt-grown Cs4PbBr6 single crystals,35
Annihilating Electron−Positron Pairs. The Doppler broad- indicating a possible second phase on the surface of these
ening PAS method can bring additional information about the single crystals.
local chemical environment of defects. The momentum We proceeded further by obtaining the Raman spectra at
distribution of an annihilating electron−positron pair can be higher wavenumbers to explore molecular modes, as shown in
calculated from first principles and can be directly compared Figure 3b. Overall, the Raman spectra for the solution-grown
with experimental results of CDB spectroscopy. The present MAPbBr3 single crystal are consistent with those reported in
approach well describes the high momentum range (p > 10−3 the literature, where the higher-frequency peaks are related to
m0c) of the momentum distribution of the annihilating pair the different MA+ cation vibrational modes, such as those at
with dominating contribution of core electrons. For the sake of 918 and 968 cm−1 being attributed to the CH3NH3+ rocking
comparison with the experiment, the calculated momentum and C−N stretching, respectively.32,36 Next, we compare the
distributions are presented as ratio curves with respect to Raman spectra for the solution-grown and melt-grown
aluminum (Al), where the calculated momentum distributions CsPbBr3 single crystals. Noticeably, the spectra for the
were divided by the momentum distribution calculated for a solution-grown CsPbBr3 single crystal contain well-defined
perfect Al crystal with an FCC structure and lattice parameter peaks that are absent for the melt-grown CsPbBr3 single
of a = 4.0495 Å. crystal. We attribute these peaks to that commonly observed in
Experimental CDB ratio curves (related to pure Al the Raman spectrum for dimethyl sulfoxide (DMSO or
reference) are plotted in Figure 2i. Calculated ratio curves (CH3)2SO),37 which is the solvent used in our growth process.
for a perfect (defect-free) CsPbBr3 and MAPbBr3 lattice are In particular, the Raman transitions at 678 and 713 cm−1 are
shown in Figure 2j. The ab initio calculations of CDB curves uncharacteristic of pure crystalline CsPbBr3 and are at
are not accurate enough to allow for quantitative comparison wavenumbers higher than what is expected for lattice phonon
with the experimental results. Nevertheless, there is clearly a modes. We attribute these peaks to C−S symmetric and
good qualitative agreement between CDB curves calculated for antisymmetric stretches, and the peak at 1000 cm−1 is assigned
perfect (defect-free) CsPbBr3 and the MAPbBr3 lattice and the to the SO stretch. Therefore, the origin of these peaks could
curves measured in the experiment. A local maximum at p ≈ 8 originate from an impurity induced by solution growth, for
× 10−3m0c followed by a minimum at p ≈ 13 × 10−3m0c and a example, solvated DMSO during growth. We confirm this
broad peak located at higher momenta observed in the through a Raman measurement of pure DMSO, which clearly
experimental curves are well reproduced in the calculated shows both peaks, indicating the presence of DMSO in the
curves. MAPbBr3 exhibits a lower peak at p ≈ 8 × 10−3m0c, and solution-grown CsPbBr3 single crystal. Individual peak assign-
the position of the local minimum and the following broad ments for both the solution-grown MAPbBr3 and solution-
peak are shifted to a slightly higher momentum compared to grown CsPbBr3 single crystals are provided in Table 2.
CsPbBr3. This is reproduced in the calculated curves as well. FTIR characterization was used to confirm that solvent
Hence, it can be concluded that positrons in both samples trapping within solution-grown single crystals is indeed taking
(CsPbBr3 and MAPbBr3) are annihilated predominantly in the place. Our measured FTIR spectrum of the solution-grown
free state, that is, not trapped at defects.
Because of the absence of positron trapping in defects in the Table 2. Chemical Origin of Defectsa
lattice of lead bromide perovskite crystals and considering the
lower sensitivity limit of PAS,30 we can estimate that the Raman solution-grown solution-grown FTIR solution-grown
concentration of negatively charged vacancies does not exceed shift MAPbBr3 peak CsPbBr3 peak Peak CsPbBr3 peak
(cm−1) assignment assignment (cm−1) assignment
1015 cm−3.30 It should be noted that if the sample contains
678 C−S symmetric 1028 SO
vacancies, an additional exponential component with the stretch
lifetime corresponding to positrons trapped in the vacancy 713 C−S asymmetric 1308 −CH3
appears in the spectrum. When the concentration of vacancies stretch symmetric
decreases, the intensity of this component decreases as well. If 887 1399 −CH3
the intensity drops below ∼10%, the vacancy component asymmetric
cannot be resolved in the spectrum anymore. It happens for a 918 CH3NH3+ 1482 −CH3
(rocking) asymmetric
vacancy concentration of ∼1015 cm−3. 968 C−N stretching 1600 environment
Unraveling the Defect Nature in Hybrid and All- 1000 SO stretch 3660 environment
Inorganic Perovskites. Next, we utilized Raman spectros- 1252 CH3−NH3+ 2909 −CH3 vibrations
copy and Fourier transform infrared (FTIR) spectroscopy to rocking
study the origin of defects. First, the Raman spectra at low 1345 C−H symmetric 2992 −CH3 vibrations
wavenumbers provide information regarding the vibrational deformation
modes of the metal-halide sublattice,31 as shown in Figure 3a. 1411 CH3 degenerate
deformation
The observed Raman peaks below 200 cm−1 in MAPbBr3 are
1426
consistent with the literature and are attributed to lattice
1477 asymmetric
vibrations.32 Next, we note that the Raman spectra for both the NH3+ bending
solution-grown and melt-grown CsPbBr3 are similar. As shown 1587 NH3+ twisting
in Figure 3a with arrows, the broad peaks around 73 and 127
cm−1 have previously been reported as the vibration mode of a
Individual peak assignments for both the solution-grown MAPbBr3
the [PbBr 6] 4− octahedron and the motion of the Cs + and solution-grown CsPbBr3 single crystals at high wavenumbers.

301 https://doi.org/10.1021/acs.chemmater.1c03540
Chem. Mater. 2022, 34, 297−306
Chemistry of Materials pubs.acs.org/cm Article

Figure 4. Surface and bulk charge transport in MHPs. (a) Schematic of PL and ToF-CW methods. (b) PL spectra of CsPbBr3 and MAPbBr3 single
crystals grown by melt and solution methods (The images on the right show the surface emission after the excitation with a 445 nm laser). More
details can be found in Supporting Information, Figure S1 and Experimental Section. (c) TRPL decay of as-grown CsPbBr3 and MAPbBr3 single
crystals grown by melt and solution methods. The analysis of TRPL and the relative contribution of τ1 and τ2 can be found in Supporting
Information, Table S1. (d) ToF-CW decay in melt-grown CsPbBr3 and solution-grown MAPbBr3 samples.

MAPbBr3 single crystal is consistent with that reported in the vibrational modes, as these measurements were performed
literature38,39 (see Figure 3c). Prominent peaks are at 969, under ambient conditions.40
1477, 1584, and 3165 cm−1, which can be attributed to the C− Charge Transport and Recombination in MHPs. We
N stretch, symmetric NH3+ bend, asymmetric NH3+ bend, and showed the presence of trapped solvents in solution-grown
symmetric NH3+ stretch, respectively. As seen with the Raman perovskites. Now we probe surface and bulk charge transport
characterization, there are noticeably defined peaks, such as properties to connect the chemical origin of defects with the
951, 1028, and 1482 cm−1, for the solution-grown CsPbBr3 transport properties of MHP devices. The free charge carrier
single crystal that are not observed for the melt-grown CsPbBr3 losses at the surface, bulk, and interfaces are the main
single crystal. We attribute these peaks to the solvent trapped recombination pathways in semiconductor optoelectronic
devices. Here, we first compare free charge recombination
within the single crystal as most of these peaks correspond to
dynamics at the surface and in the bulk of hybrid and all-
peaks observed in the FTIR spectrum for DMSO as described
inorganic lead bromide perovskites using photoluminescence
in Table 2. In Figure 3c, we compare the FTIR spectrum of (PL) and time-of-flight current waveform (ToF-CW) spec-
DMSO to that of the solution-grown CsPbBr3 to further troscopy (Figure 4a). As can be seen in Figure 4b, there is a
illustrate this point. The trapped solvent can form point defects small deviation in the PL spectra of melt-grown CsPbBr3 and
and defect complexes responsible for the shorter carrier solution-grown CsPbBr3 and MAPbBr3 single crystals with
lifetime in solution-grown single crystals. In addition, the bandgap energies of 2.34, 2.33, and 2.31 eV, respectively. A
trapped solvent can deteriorate the stability of the perovskite second PL peak at ∼562 nm is observed in the melt-grown
devices. For both the solution-grown and melt-grown CsPbBr3 CsPbBr3 single crystals. The origin of double peak emission in
single crystals, we attribute a few low-intensity peaks at 1600 MHPs was attributed to the self-absorption in several previous
and 3500 cm−1 to the environment, specifically water studies.41−43 It was demonstrated that the often-seen double
302 https://doi.org/10.1021/acs.chemmater.1c03540
Chem. Mater. 2022, 34, 297−306
Chemistry of Materials pubs.acs.org/cm Article

PL peaks are a general phenomenon which occurs because of and melt- and solution-grown MAPbBr3 show a similar
the high internal reflection probability of PL in MHP samples. mechanism of positron annihilation in the lattice. Thus, both
The intensity of the additional peak can be controlled by the inorganic and hybrid perovskites do not contain defects
reflection and therefore could be different from sample to capable of positron trapping, that is, the concentration of
sample. In addition, the excitation of the samples with 445 nm vacancies is lower than ≈1015 cm−3.
light induces bright luminescence and dark spots emitted from This study clarifies the impact of growth conditions on the
the near-surface region of these crystals, likely because of the performance of hybrid and all-inorganic radiation sensors. The
presence of surface defects.44 Time-resolved photolumines- comparison of fundamental properties of single crystals grown
cence (TRPL) measurements reveal a nanosecond decay of the from the melt and solution gives valuable information needed
PL signal in all studied samples (Figure 4c). Using a double for further development of perovskite materials. The
exponential fit, we found that the fast and slow PL decay time demonstrated defect nature and recombination channels reveal
constants for MAPbBr3 are around 40 and 140 ns, respectively. further pathways that can lead to more efficient, stable, and
For the melt- and solution-grown CsPbBr3 crystals, the fast/ cheap perovskite devices.


slow decay time constants were found to be around 7/100 and
2/37 ns, respectively. EXPERIMENTAL SECTION
To further explore charge carrier lifetime in these crystals,
Single-Crystal Solution Growth. The MAPbBr3 and CsPbBr3
ToF-CW spectroscopy was performed. The holes (generated single crystals were grown using the inverse temperature crystal-
by a short 1 μs laser pulse with a wavelength of 450 nm and 1 lization method.1 MABr (99.99% from Greatcell Solar Materials),
μW cm−2 intensity) drift in the bulk of lead halide perovskite CsBr (99.999% from Sigma-Aldrich), and PbBr2 (99.999% from
crystals and induce a current waveform. Thus, the ToF-CW Sigma-Aldrich) were used without further purification. For MAPbBr3,
reveals bulk recombination dynamics and lifetime of the holes. a molar ratio of 1.2 to 1 of MABr to PbBr2 was dissolved in N,N-
Recently, we showed that charge trapping by bulk defects is Dimethylformamide (99.8% anhydrous from Sigma-Aldrich) for a 1
responsible for ToF-CW decay.21 Here, the ToF-CW spec- M solution. For CsPbBr3, a molar ratio of 1 to 2 of CsBr to PbBr2 was
troscopy reveals microsecond dynamics of drifting free holes dissolved in dimethyl sulfoxide (99.9% anhydrous from Sigma-
Aldrich) for a 1 M solution.2 Both solutions were filtered using a 0.2
(Figure 4d) under an applied bias of 60 V. The lifetime of free
μm PTFE membrane syringe filter. Crystals precipitated during
holes in the bulk was estimated by conducting an exponential gradual heating at a rate of 5 °C/day from room temperature to ∼70
fit to the ToF transients. Interestingly, the melt-grown CsPbBr3 °C for MAPbBr3 and ∼120 °C for CsPbBr3.
crystal shows a hole lifetime of around τCS = 210 μs and the Single-Crystal Bridgman Technique. CsPbBr3 crystals were
MAPbBr3 crystal shows a hole lifetime of around τMA = 95 μs. grown by the Bridgman technique. CsBr (6 N) and PbBr2 (5 N)
The solution-grown CsPbBr3 crystal did not exhibit a reliable (manufactured by Alfa Aesar) in the stoichiometric ratio were loaded
ToF transient probably because of significant hole trapping by in a quartz ampoule, evacuated to 10−5 mbar, and sealed. Synthesis of
the deep and shallow defects. It should be noted that the PL perovskite occurred by the fusion of starting materials in a three-zone
lifetime affected by surface properties differs from bulk lifetime vertical furnace. The ampoule with reagents was slowly heated to a
temperature of 645 °C to exceed the CsBr melting point. At this
measured by ToF-CW. Other studies also highlighted the
temperature, an isothermal holding was carried out for several hours
critical effect of surface recombination on TRPL measure- to achieve melt homogenization and complete the reaction between
ments.45 These results elucidate a notable difference in charge CsBr and PbBr2. Afterward, the furnace was cooled down to room
recombination dynamics on the surface and within the bulk of temperature for 20 h. The following stage was growing a CsPbBr3
the lead halide perovskites. single crystal. The ampoule with the synthesized perovskite was

■ CONCLUSIONS
In summary, comparing the charge decay by TRPL and ToF,
gradually heated up to 585 °C and lowered at a rate of 3 mm/h into
the “cold” zone of the furnace. After completing the directional
crystallization, lowering was stopped, and the grown crystal was slowly
cooled down to room temperature for 30 h. The paper focuses on
we can conclude that these two methods show different decay bulk defects and the effect of the trapped solvent on the charge
dynamics. Free carriers drifting in the ToF experiment are properties. The study of different polish treatments is out of focus in
mainly affected by defects in the bulk of perovskites. this paper.
Meanwhile, surface quality has a substantial effect on PL PL Spectroscopy. PL and TRPL were performed using a Horiba
decay. TRPL does not reflect the lifetime of charge carriers in Fluorolog3 spectrometer. A 405 nm CW laser was used for excitation
the bulk of MHPs. The dominant charge losses occur at the during PL measurements. A 343 nm pulsed laser (280 fs, 200 kHz)
surface defects with a recombination lifetime of several to was used for TRPL measurements at the peak emission intensity.
hundreds of nanoseconds. The bulk charge transport shows a Luminescence Measurements. A confocal microspectrofluorim-
eter adapted for time-resolved fluorescence measurements using a
much longer lifetime of hundreds of microseconds. Among the phase−modulation principle with homodyne data acquisition was
studied hybrid and inorganic lead bromide perovskites, the employed to obtain fluorescence spectra of perovskite samples and to
CsPbBr3 crystal grown from melt shows a longer bulk charge determine the fluorescence lifetimes. The laser diode module
carrier lifetime of 210 μs. We attribute the difference in charge (Omicron LDM 442.50.A350, Germany) with sinusoidal intensity
transport properties between melt- and solution-grown crystals modulation (50 mW peak output, attenuated to one to tens of μW at
to solvent trapping within solution-grown single crystals. the sample) was used for the excitation at 445 nm wavelength. A
DMSO peaks were detected by Raman and FTIR spectroscopy confocal epifluorescence upright microscope (Zeiss UMSP−80,
methods in solution-grown single crystals. By analyzing the Germany) with a 10× objective was used to collect the fluorescence
PAS results, we conclude that positrons annihilate in the free emission signal from the excited volume of the sample. It is focused
on the entrance slit of the Jobin−Yvon HR640 spectrograph equipped
state (i.e., not trapped at defects) in both all-inorganic and with a 100 line/mm grating. The spectral detection window (375 nm
hybrid perovskite semiconductors grown by different methods. wide) covered both the excitation wavelength (elastic scattering
From a comparison of measured positron lifetimes and provided us a lifetime reference) and the perovskite emission
momentum distributions of annihilating electron−positron spectrum. The fluorescence lifetime was calculated from the
pairs with ab initio calculations, CsPbBr3 grown from solution frequency-dependent phase shift and the intensity demodulation for

303 https://doi.org/10.1021/acs.chemmater.1c03540
Chem. Mater. 2022, 34, 297−306
Chemistry of Materials pubs.acs.org/cm Article

modulation frequencies covering evenly 5−165 MHz interval. Mahshid Ahmadi − Joint Institute for Advanced Materials,
Calculated phase differences and modulation ratios are used as an Department of Materials Science and Engineering, University
input table for the GLOBAL fitting program (Laboratory for of Tennessee, Knoxville, Tennessee 37996, United States;
Fluorescence Dynamics, Irvine, USA) to calculate single fluorescence orcid.org/0000-0002-3268-7957; Email: mahmadi3@
lifetime components and their percentage spectral representation.
Raman Spectroscopy and FTIR Measurements. Raman
utk.edu
spectroscopy was carried out on a HORIBA XploRA PLUS Raman Authors
Jakub Č ížek − Faculty of Mathematics and Physics,
system with a 785 nm laser at 1 mW through a confocal microscope
with a 100× objective lens. Multiple scans were averaged for each
measurement between 600 and 2500 cm−1. FTIR spectroscopy on Department of Low-temperature Physics, Charles University,
single crystals and DMSO was performed on a Thermo-Scientific Prague 8 CZ-18000, Czech Republic
Nicolet iS50 FTIR. Hassan Elhadidy − Faculty of Mathematics and Physics,
Positron Annihilation Spectroscopy. A 22Na radioisotope with Institute of Physics, Charles University, Prague 2 CZ-121 16,
an activity of ≈1 MBq sealed between two Mylar foils with a thickness Czech Republic
of 2 μm was used as a positron source. A digital spectrometer with a Petr Praus − Faculty of Mathematics and Physics, Institute of
time resolution of 145 ps (full width at half maximum (FWHM) of Physics, Charles University, Prague 2 CZ-121 16, Czech
the resolution function) was used for positron lifetime measurements. Republic
Fitting of positron lifetime spectra was performed using the PLRF
code.3 The source contribution to the LT spectra consisted of (i) a Kate Higgins − Joint Institute for Advanced Materials,
component with a lifetime of ≈368 ps and a relative intensity of Department of Materials Science and Engineering, University
≈10% representing a contribution of positrons annihilated in the of Tennessee, Knoxville, Tennessee 37996, United States;
source spot and inside the Mylar foil and (ii) a weak long-lived orcid.org/0000-0001-5503-2884
component with a lifetime of ≈1.5 ns and an intensity of ≈1% Bogdan Dryzhakov − Joint Institute for Advanced Materials,
originating from pick-off annihilation of ortho-positronium formed in Department of Materials Science and Engineering, University
the covering Mylar foil. The CDB studies were carried out using a of Tennessee, Knoxville, Tennessee 37996, United States
digital spectrometer equipped with two high-purity Ge detectors. The Andrii Kanak − Yuriy Fedkovych Chernivtsi National
CDB spectrometer was characterized by an energy resolution University, Chernivtsi 58012, Ukraine; orcid.org/0000-
(FWHM) of 0.9 keV at the annihilation line and a peak-to-
background ratio higher than 105. The results of the CDB 0001-9238-4029
measurements are presented in this paper as ratio curves related to Franck Sureau − Laboratoire Jean Perrin, Sorbonne
a well-annealed pure Al (99.9999%) reference sample. University, Paris 75005, France
Positron Annihilation Ab Initio Calculations. Positron Jindřich Pipek − Faculty of Mathematics and Physics, Institute
annihilation-related parameters (positron lifetimes and momentum of Physics, Charles University, Prague 2 CZ-121 16, Czech
distribution of annihilating electron−positron pairs) were calculated Republic
using density functional theory. Ab initio calculations of positron Eduard Belas − Faculty of Mathematics and Physics, Institute
parameters were performed within the so-called standard scheme.4 In of Physics, Charles University, Prague 2 CZ-121 16, Czech
this approximation, positron density is assumed to be everywhere Republic
vanishingly small and not affecting the bulk electron structure.
Modeling of point defects was performed using 320 atom-based Marián Betušiak − Faculty of Mathematics and Physics,
supercells consisting of 4 × 4 × 4 perovskite units. Vacancies were Institute of Physics, Charles University, Prague 2 CZ-121 16,
created simply by removing corresponding atoms. Details of Czech Republic
calculations of positron lifetimes are described in the Supporting Mykola Brynza − Faculty of Mathematics and Physics,
Information. Institute of Physics, Charles University, Prague 2 CZ-121 16,
Time-of-Flight Current Spectroscopy. To generate free carriers Czech Republic
and control their drift in perovskite single-crystal samples, we used a 1 Eric Lukosi − Joint Institute for Advanced Materials,
μs blue laser (450 nm) and 500 μs pulse bias. The working principle Department of Nuclear Engineering, University of Tennessee,
of the ToF setup is shown in Figure S2. The laser pulse generates free Knoxville, Tennessee 37996, United States
holes near the anode. The bias pulse, synchronized with a light pulse,
induces hole drift to the cathode. The free electrons have no influence Bin Hu − Joint Institute for Advanced Materials, Department
on CW because of their fast collection at the anode. Drifting holes can of Materials Science and Engineering, University of Tennessee,
interact with deep traps, which limit the lifetime of holes. This Knoxville, Tennessee 37996, United States; orcid.org/
phenomenon is observed by the exponential decay of CW. 0000-0002-1573-7625


*
ASSOCIATED CONTENT
sı Supporting Information
Complete contact information is available at:
https://pubs.acs.org/10.1021/acs.chemmater.1c03540

The Supporting Information is available free of charge at Notes


https://pubs.acs.org/doi/10.1021/acs.chemmater.1c03540. The authors declare no competing financial interest.
The views and conclusions contained in this document are
ToF and PL setup details, positron annihilation ab initio
those of the authors and should not be interpreted as
calculation evaluation, and details of charge transport
necessarily representing the official policies, either expressed
mobility and lifetimes (PDF)
or implied, of the U.S. Department of Homeland Security.

■ AUTHOR INFORMATION
Corresponding Authors
■ ACKNOWLEDGMENTS
A.M., J.P., P.P, M.B., and E.B. acknowledge financial support
Artem Musiienko − Helmholtz-Zentrum Berlin für from the Grant Agency of the Czech Republic, Grant No.
Materialien und Energie GmbH, Berlin 12489, Germany; P102/19/11920S and the Grant Agency of Charles University,
orcid.org/0000-0002-2259-8387; projects No. 1234119 and 379621. A.M. thanks Helmholtz-
Email: artem.musiienko@helmholtz-berlin.de Zentrum Berlin for support and acknowledges financial
304 https://doi.org/10.1021/acs.chemmater.1c03540
Chem. Mater. 2022, 34, 297−306
Chemistry of Materials pubs.acs.org/cm Article

support from the German Science Foundation (DFG) in the Temperature Solution-Grown CsPbBr3 Single Crystals and Their
framework of the priority program “Perovskite Semiconduc- Characterization. Cryst. Growth Des. 2016, 16, 5717−5725.
tors: From Fundamental Properties to Devices” (SPP 2196). (13) He, Y.; Matei, L.; Jung, H. J.; McCall, K. M.; Chen, M.;
H.E. thanks the Ministry of Education, Youth, and Sports of Stoumpos, C. C.; Liu, Z.; Peters, J. A.; Chung, D. Y.; Wessels, B. W.;
the Czech Republic − National Program of Sustainability Wasielewski, M. R.; Dravid, V. P.; Burger, A.; Kanatzidis, M. G. High
spectral resolution of gamma-rays at room temperature by perovskite
(NPU LO1305). Computational resources were supplied by
CsPbBr3 single crystals. Nat. Commun. 2018, 9, 1609.
the project ″e-Infrastruktura CZ″ (e-INFRA LM2018140) (14) Feng, Y.; Pan, L.; Wei, H.; Liu, Y.; Ni, Z.; Zhao, J.; Rudd, P. N.;
provided within the program Projects of Large Research, Cao, L. R.; Huang, J. Low defects density CsPbBr3 single crystals
Development, and Innovations Infrastructures. A.K. gratefully grown by an additive assisted method for gamma-ray detection. J.
acknowledges Prof. Fochuk P.M. and Dr. Khalavka Y.B. for Mater. Chem. C 2020, 8, 11360−11368.
support. K.H., B.D., E.L., B.H., and M.A. acknowledge financial (15) Ahn, N.; Kwak, K.; Jang, M. S.; Yoon, H.; Lee, B. Y.; Lee, J.-K.;
support from U.S. Department of Homeland Security (grant # Pikhitsa, P. V.; Byun, J.; Choi, M. Trapped charge-driven degradation
2016-DN-077-ARI01). We are grateful to Prof. Sergei V of perovskite solar cells. Nat. Commun. 2016, 7, 13422.
Kalinin for insightful comments on the manuscript. (16) Duan, J.; Zhao, Y.; Wang, Y.; Yang, X.; Tang, Q. Hole-Boosted


Cu(Cr,M)O2 Nanocrystals for All-Inorganic CsPbBr3 Perovskite
ABBREVIATIONS Solar Cells. Angew. Chem., Int. Ed. 2019, 58, 16147−16151.
(17) Saidaminov, M. I.; Haque, M. A.; Almutlaq, J.; Sarmah, S.;
MHPs, metal-halide perovskites; FTIR, Fourier transform Miao, X.-H.; Begum, R.; Zhumekenov, A. A.; Dursun, I.; Cho, N.;
infrared; PAS, positron annihilation spectroscopy; PL, photo- Murali, B.; Mohammed, O. F.; Wu, T.; Bakr, O. M. Inorganic Lead
luminescence; CW-ToF, current waveform time of flight; Halide Perovskite Single Crystals: Phase-Selective Low-Temperature
TRPL, time-resolved photoluminescence; μτ, mobility-life- Growth, Carrier Transport Properties, and Self-Powered Photo-
time; CDB, coincidence Doppler broadening detection. Adv. Opt. Mater. 2017, 5, No. 1600704.


(18) Zhang, H.; Liu, X.; Dong, J.; Yu, H.; Zhou, C.; Zhang, B.; Xu,
Y.; Jie, W. Centimeter-Sized Inorganic Lead Halide Perovskite
REFERENCES CsPbBr3 Crystals Grown by an Improved Solution Method. Cryst.
(1) Wei, H.; DeSantis, D.; Wei, W.; Deng, Y.; Guo, D.; Savenije, T. Growth Des. 2017, 17, 6426−6431.
J.; Cao, L.; Huang, J. Dopant compensation in alloyed (19) Wang, F.; Zhang, H.; Sun, Q.; Hafsia, A. B.; Chen, Z.; Zhang,
CH3NH3PbBr3−xClx perovskite single crystals for gamma-ray B.; Xu, Y.; Jie, W. Low-Temperature Solution Growth and
spectroscopy. Nat. Mater. 2017, 16, 826−833. Characterization of Halogen (Cl, I)-Doped CsPbBr3 Crystals. Cryst.
(2) Wei, H.; Huang, J. Halide lead perovskites for ionizing radiation Growth Des. 2020, 20, 1638−1645.
detection. Nat. Commun. 2019, 10, 1066. (20) Zhang, H.; Wang, F.; Lu, Y.; Sun, Q.; Xu, Y.; Zhang, B.-B.; Jie,
(3) Sutton, R. J.; Eperon, G. E.; Miranda, L.; Parrott, E. S.; Kamino, W.; Kanatzidis, M. G. High-sensitivity X-ray detectors based on
B. A.; Patel, J. B.; Hörantner, M. T.; Johnston, M. B.; Haghighirad, A. solution-grown caesium lead bromide single crystals. J. Mater. Chem.
A.; Moore, D. T.; Snaith, H. J. Bandgap-Tunable Cesium Lead Halide C 2020, 8, 1248−1256.
Perovskites with High Thermal Stability for Efficient Solar Cells. Adv. (21) Musiienko, A.; Pipek, J.; Praus, P.; Brynza, M.; Belas, E.;
Energy Mater. 2016, 6, No. 1502458. Dryzhakov, B.; Du, M.-H.; Ahmadi, M.; Grill, R. Deciphering the
(4) Yakunin, S.; Dirin, D. N.; Shynkarenko, Y.; Morad, V.; effect of traps on electronic charge transport properties of
Cherniukh, I.; Nazarenko, O.; Kreil, D.; Nauser, T.; Kovalenko, M. methylammonium lead tribromide perovskite. Sci. Adv. 2020, 6,
V. Detection of gamma photons using solution-grown single crystals No. eabb6393.
of hybrid lead halide perovskites. Nat. Photonics 2016, 10, 585−589. (22) Musiienko, A.; Moravec, P.; Grill, R.; Praus, P.; Vasylchenko, I.;
(5) Shi, D.; Adinolfi, V.; Comin, R.; Yuan, M.; Alarousu, E.; Buin, A.; Pekarek, J.; Tisdale, J.; Ridzonova, K.; Belas, E.; Landová, L.; Hu, B.;
Chen, Y.; Hoogland, S.; Rothenberger, A.; Katsiev, K.; Losovyj, Y.; Lukosi, E.; Ahmadi, M. Deep levels, charge transport and mixed
Zhang, X.; Dowben, P. A.; Mohammed, O. F.; Sargent, E. H.; Bakr, O. conductivity in organometallic halide perovskites. Energy Environ. Sci.
M. Low trap-state density and long carrier diffusion in organolead
2019, 12, 1413−1425.
trihalide perovskite single crystals. Science 2015, 347, 519−522.
(23) Guin, S. N.; Banerjee, S.; Sanyal, D.; Pati, S. K.; Biswas, K.
(6) Wen, X.; Feng, Y.; Huang, S.; Huang, F.; Cheng, Y.-B.; Green,
Origin of the Order−Disorder Transition and the Associated
M.; Ho-Baillie, A. Defect trapping states and charge carrier
Anomalous Change of Thermopower in AgBiS2 Nanocrystals: A
recombination in organic−inorganic halide perovskites. J. Mater.
Chem. C 2016, 4, 793−800. Combined Experimental and Theoretical Study. Inorg. Chem. 2016,
(7) Zheng, X.; Chen, B.; Dai, J.; Fang, Y.; Bai, Y.; Lin, Y.; Wei, H.; 55, 6323−6331.
Zeng, X. C.; Huang, J. Defect passivation in hybrid perovskite solar (24) Guin, S. N.; Pan, J.; Bhowmik, A.; Sanyal, D.; Waghmare, U. V.;
cells using quaternary ammonium halide anions and cations. Nat. Biswas, K. Temperature Dependent Reversible p−n−p Type
Energy 2017, 2, 17102. Conduction Switching with Colossal Change in Thermopower of
(8) Kang, J.; Wang, L. W. High Defect Tolerance in Lead Halide Semiconducting AgCuS. J. Am. Chem. Soc. 2014, 136, 12712−12720.
Perovskite CsPbBr3. J. Phys. Chem. Lett. 2017, 8, 489−493. (25) Grebennikov, D.; Ovchar, O.; Belous, A.; Mascher, P.
(9) Dirin, D. N.; Cherniukh, I.; Yakunin, S.; Shynkarenko, Y.; Application of positron annihilation and Raman spectroscopies to
Kovalenko, M. V. Solution-Grown CsPbBr3 Perovskite Single the study of perovskite type materials. J. Appl. Phys. 2010, 108,
Crystals for Photon Detection. Chem. Mater. 2016, 28, 8470−8474. 114109.
(10) Stoumpos, C. C.; Malliakas, C. D.; Peters, J. A.; Liu, Z.; (26) Pansara, P. R.; Meshiya, U. M.; Makadiya, A. R.; Raval, P. Y.;
Sebastian, M.; Im, J.; Chasapis, T. C.; Wibowo, A. C.; Chung, D. Y.; Modi, K. B.; Nambissan, P. M. G. Defect structure transformation
Freeman, A. J.; Wessels, B. W.; Kanatzidis, M. G. Crystal Growth of during substitution in quadruple perovskite CaCu3-xTi4-xFe2xO12
the Perovskite Semiconductor CsPbBr3: A New Material for High- studied by positron annihilation spectroscopy. Ceram. Int. 2019, 45,
Energy Radiation Detection. Cryst. Growth Des. 2013, 13, 2722−2727. 18599−18603.
(11) Ding, J.; Du, S.; Zuo, Z.; Zhao, Y.; Cui, H.; Zhan, X. High (27) Sundar, C. S. Positron annihilation spectroscopy in materials
Detectivity and Rapid Response in Perovskite CsPbBr3 Single-Crystal science. Bull. Mater. Sci. 1994, 17, 1215−1232.
Photodetector. J. Phys. Chem. C 2017, 121, 4917−4923. (28) Dhar, J.; Sil, S.; Dey, A.; Sanyal, D.; Ray, P. P. Investigation of
(12) Rakita, Y.; Kedem, N.; Gupta, S.; Sadhanala, A.; Kalchenko, V.; Ion-Mediated Charge Transport in Methylammonium Lead Iodide
Böhm, M. L.; Kulbak, M.; Friend, R. H.; Cahen, D.; Hodes, G. Low- Perovskite. J. Phys. Chem. C 2017, 121, 5515−5522.

305 https://doi.org/10.1021/acs.chemmater.1c03540
Chem. Mater. 2022, 34, 297−306
Chemistry of Materials pubs.acs.org/cm Article

(29) Boroński, E.; Nieminen, R. M. Electron-positron density-


functional theory. Phys. Rev. B 1986, 34, 3820−3831.
(30) Gebauer, J.; Rudolf, F.; Polity, A.; Krause-Rehberg, R.; Martin,
J.; Becker, P. On the sensitivity limit of positron annihilation:
detection of vacancies in as-grown silicon. Appl. Phys. A 1999, 68,
411−416.
(31) Cha, J. H.; Noh, K.; Yin, W.; Lee, Y.; Park, Y.; Ahn, T. K.;
Mayoral, A.; Kim, J.; Jung, D. Y.; Terasaki, O. Formation and
Encapsulation of All-Inorganic Lead Halide Perovskites at Room
Temperature in Metal-Organic Frameworks. J. Phys. Chem. Lett. 2019,
10, 2270−2277.
(32) Wang, K.-H.; Li, L.-C.; Shellaiah, M.; Wen Sun, K. Structural
and Photophysical Properties of Methylammonium Lead Tribromide
(MAPbBr3) Single Crystals. Sci. Rep. 2017, 7, 13643.
(33) Calistru, D. M.; Mihut, L.; Lefrant, S.; Baltog, I. Identification
of the symmetry of phonon modes in CsPbCl3 in phase IV by Raman
and resonance-Raman scattering. J. Appl. Phys. 1997, 82, 5391−5395.
(34) Shibata, K.; Yan, J.; Hazama, Y.; Chen, S.; Akiyama, H. Exciton
Localization and Enhancement of the Exciton−LO Phonon
Interaction in a CsPbBr3 Single Crystal. J. Phys. Chem. C 2020,
124, 18257−18263.
(35) Cha, J. H.; Han, J. H.; Yin, W.; Park, C.; Park, Y.; Ahn, T. K.;
Cho, J. H.; Jung, D. Y. Photoresponse of CsPbBr3 and Cs4PbBr6
Perovskite Single Crystals. J. Phys. Chem. Lett. 2017, 8, 565−570.
(36) Xie, L. Q.; Zhang, T. Y.; Chen, L.; Guo, N.; Wang, Y.; Liu, G.
K.; Wang, J. R.; Zhou, J. Z.; Yan, J. W.; Zhao, Y. X.; Mao, B. W.; Tian,
Z. Q. Organic-inorganic interactions of single crystalline organolead
halide perovskites studied by Raman spectroscopy. Phys. Chem. Chem.
Phys. 2016, 18, 18112−18118.
(37) Batista, A. N. L.; Batista, J. M., Jr.; Bolzani, V. S.; Furlan, M.;
Blanch, E. W. Selective DMSO-induced conformational changes in
proteins from Raman optical activity. Phys. Chem. Chem. Phys. 2013,
15, 20147−20152.
(38) Mahapatra, A.; Parikh, N.; Kumari, H.; Pandey, M. K.; Kumar,
M.; Prochowicz, D.; Kalam, A.; Tavakoli, M. M.; Yadav, P. Reducing
ion migration in methylammonium lead tri-bromide single crystal via
lead sulfate passivation. J. Appl. Phys. 2020, 127, 185501. Recommended by ACS
(39) Xie, L. Q.; Chen, L.; Nan, Z. A.; Lin, H. X.; Wang, T.; Zhan, D.
P.; Yan, J. W.; Mao, B. W.; Tian, Z. Q. Understanding the Cubic
Phase Stabilization and Crystallization Kinetics in Mixed Cations and Charge Carriers Trapping by the Full-Configuration Defects
Halides Perovskite Single Crystals. J. Am. Chem. Soc. 2017, 139, in Metal Halide Perovskites Quantum Dots
3320−3323. Xiao-Yi Liu, Zi-Wu Wang, et al.
(40) Zhang, M.; Zheng, Z.; Fu, Q.; Chen, Z.; He, J.; Zhang, S.; SEPTEMBER 19, 2022
Chen, C.; Luo, W. Synthesis and single crystal growth of perovskite THE JOURNAL OF PHYSICAL CHEMISTRY LETTERS READ
semiconductor CsPbBr 3. J. Cryst. Growth 2018, 484, 37−42.
(41) Schötz, K.; Askar, A. M.; Peng, W.; Seeberger, D.; Gujar, T. P.; Tuning Defects in a Halide Double Perovskite with Pressure
Thelakkat, M.; Köhler, A.; Huettner, S.; Bakr, O. M.; Shankar, K.;
Nathan R. Wolf, Hemamala I. Karunadasa, et al.
Panzer, F. Double peak emission in lead halide perovskites by self-
NOVEMBER 07, 2022
absorption. J. Mater. Chem. C 2020, 8, 2289−2300. JOURNAL OF THE AMERICAN CHEMICAL SOCIETY READ
(42) Fang, Y.; Wei, H.; Dong, Q.; Huang, J. Quantification of re-
absorption and re-emission processes to determine photon recycling
efficiency in perovskite single crystals. Nat. Commun. 2017, 8, 14417. Accurately Determining the Phase Transition Temperature
(43) Goetz, K. P.; Taylor, A. D.; Paulus, F.; Vaynzof, Y. Shining of CsPbI3 via Random-Phase Approximation Calculations
Light on the Photoluminescence Properties of Metal Halide and Phase-Transferable Machine Learning Potentials
Perovskites. Adv. Funct. Mater. 2020, 30, No. 1910004. Tom Braeckevelt, Veronique Van Speybroeck, et al.
(44) Ding, J.; Lian, Z.; Li, Y.; Wang, S.; Yan, Q. The Role of Surface SEPTEMBER 22, 2022
Defects in Photoluminescence and Decay Dynamics of High-Quality CHEMISTRY OF MATERIALS READ
Perovskite MAPbI3 Single Crystals. J. Phys. Chem. Lett. 2018, 9,
4221−4226. Enhanced Light Emission through Symmetry Engineering of
(45) Weiss, T. P.; Bissig, B.; Feurer, T.; Carron, R.; Buecheler, S.; Halide Perovskites
Tiwari, A. N. Bulk and surface recombination properties in thin film
Yoonhoo Ha, Hyungjun Kim, et al.
semiconductors with different surface treatments from time-resolved
DECEMBER 27, 2021
photoluminescence measurements. Sci. Rep. 2019, 9, 5385. JOURNAL OF THE AMERICAN CHEMICAL SOCIETY READ

Get More Suggestions >

306 https://doi.org/10.1021/acs.chemmater.1c03540
Chem. Mater. 2022, 34, 297−306

You might also like