You are on page 1of 6

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/314222583

Quantitative Phase-Change Thermodynamics and Metastability of


Perovskite-Phase Cesium Lead Iodide

Article  in  Journal of Physical Chemistry Letters · March 2017


DOI: 10.1021/acs.jpclett.7b00134

CITATIONS READS

44 1,773

6 authors, including:

Subham Dastidar Christopher J Hawley


Drexel University Drexel University
8 PUBLICATIONS   190 CITATIONS    14 PUBLICATIONS   227 CITATIONS   

SEE PROFILE SEE PROFILE

Alejandro Gutierrez-Perez Jonathan E Spanier


Drexel University Drexel University
5 PUBLICATIONS   54 CITATIONS    129 PUBLICATIONS   4,278 CITATIONS   

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Subham Dastidar on 16 November 2017.

The user has requested enhancement of the downloaded file.


Letter

pubs.acs.org/JPCL

Quantitative Phase-Change Thermodynamics and Metastability of


Perovskite-Phase Cesium Lead Iodide
Subham Dastidar,† Christopher J. Hawley,‡ Andrew D. Dillon,† Alejandro D. Gutierrez-Perez,‡
Jonathan E. Spanier,‡ and Aaron T. Fafarman*,†

Department of Chemical and Biological Engineering, Drexel University, 3141 Chestnut Street, Philadelphia, Pennsylvania 19104,
United States

Department of Materials Science & Engineering, Drexel University, Philadelphia, Pennsylvania 19104, United States
*
S Supporting Information

ABSTRACT: The perovskite phase of cesium lead iodide (α-CsPbI3 or “black” phase) possesses
favorable optoelectronic properties for photovoltaic applications. However, the stable phase at
room temperature is a nonfunctional “yellow” phase (δ-CsPbI3). Black-phase polycrystalline thin
films are synthesized above 330 °C and rapidly quenched to room temperature, retaining their
phase in a metastable state. Using differential scanning calorimetry, it is shown herein that the
metastable state is maintained in the absence of moisture, up to a temperature of 100 °C, and a
reversible phase-change enthalpy of 14.2 (±0.5) kJ/mol is observed. The presence of
atmospheric moisture hastens the black-to-yellow conversion kinetics without significantly
changing the enthalpy of the transition, indicating a catalytic effect, rather than a change in
equilibrium due to water adduct formation. These results delineate the conditions for trapping
the desired phase and highlight the significant magnitude of the entropic stabilization of this
phase.

apid advances in the field of organic−inorganic hybrid


R perovskites (OHPs) have significantly altered the photo-
voltaic research landscape. These materials boast exciting
metastable state for use as a photovoltaic absorber. Our group
recently demonstrated that chloride doping of CsPbI3, by
codeposition of a mixture of CsPbI3 and CsPbCl3 nanocrystals
photophysical properties and robust, inexpensive fabrication and subsequent chemical sintering, improves the phase stability
techniques and reported power conversion efficiencies up to of the resulting polycrystalline α-CsPbI3.18 Deliberate for-
22%.1 However, OHPs suffer from chemical instability due to mation of small crystal grains was associated with improved
the volatile2 and hygroscopic3,4 nature of the organic metastability in α-CsPbI3 thin films.5 This effect may also be at
component. Recently, perovskite-phase cesium lead iodide play in the demonstration that α-CsPbI3 quantum dots are
(CsPbI3) has emerged as an all-inorganic alternative to the stable under ambient conditions.6,19,20 Also, CsPbI3−xBrx mixed
OHPs with strikingly similar optoelectronic properties.5−7 crystals have been shown to be more stable than the pure-phase
Cesium has also been shown to play a critical role as a CsPbI3.21,22 Developing a quantitative understanding of the
substitutional dopant, imparting enhanced stability to the phase-change thermodynamics and kinetics of α-CsPbI3 will be
recently discovered mixed cation perovskites that form the basis critical to these ongoing efforts to improve its stability,
of the highest performing OHP solar cells to date.8−13 particularly in the presence of atmospheric moisture.
According to the Goldschmidt tolerance factor, an empirical Herein, the thermodynamics of the reversible phase
index of the structural stability of perovskites, cesium is almost transition between the black and yellow phases of CsPbI3 has
too small to form a stable perovskite with lead iodide,10,14 been examined, and the transition enthalpies are reported using
despite being the largest nonradioactive elemental cation. differential scanning calorimetry (DSC). The DSC results show
Nonetheless, at elevated temperature (>305 °C) CsPbI3 adopts a narrow endothermic peak around 320 °C and a much wider
a cubic perovskite structure with a band gap 1.73 eV,5,6,15,16 exothermic feature around 270 °C, signifying the yellow-to-
known as the “black” phase or α-CsPbI3 (Figure 1a). The black phase conversion and reappearance of the yellow phase,
equilibrium phase of CsPbI3 at room temperature is a high- respectively. The position and breadth of the peaks provide
bandgap, nonperovskite “yellow” phase (δ-CsPbI3) that is insights into the kinetics of the phase transition in the regime in
essentially nonfunctional as a photovoltaic material.17 It follows which the perovskite phase is metastable. Both indicators are
simply that the yellow δ-phase is enthalpically favored, while measured herein as a function of the rate of heat flow,
the functional, black α-CsPbI3 is entropically favored, however, demonstrating significant thermal hysteresis, which is, in turn,
the magnitudes of these competing contributions have been
heretofore unknown. Received: January 17, 2017
For practical applications, several strategies are being Accepted: March 3, 2017
explored to form the α-CsPbI3 phase and preserve it in a Published: March 3, 2017

© 2017 American Chemical Society 1278 DOI: 10.1021/acs.jpclett.7b00134


J. Phys. Chem. Lett. 2017, 8, 1278−1282
The Journal of Physical Chemistry Letters Letter

Figure 1. (a) Crystal structure of black (α) and yellow (δ) phases of CsPbI3 at their thermodynamically stable temperatures. (b) XRD of CsPbI3
drop-cast films in black and yellow phase at room temperature. (c) UV−vis transmittance spectra of a thin film of CsPbI3, displaying complete
transition from black to yellow phase within a period of 75 min after exposing to relative humidity (RH) of 33% at 23 °C. (d) Microscopic images of
a thin film of CsPbI3, showing the generation of a yellow nucleation site, subsequent growth, and complete conversion within 15 min. The images
were taken at 2, 5, 10, and 15 min after the exposure, respectively, at RH 50% and 23 °C.

critical to the formation of a trapped state. In the literature, metastable “parent” black phase (Figure 1d). Subsequently, the
thermal hysteresis is observed in OHPs as well. 23,24 yellow phase continues to grow and eventually converts the
Interestingly, in the formamidinium lead iodide system, it has entire film.
been associated with conformational entropy of the A-site The same general phenomena are observed if metastable
molecular cation,24 whereas there can be no such contribution films are instead subjected to elevated temperature while
from cesium. In the current work, additional calorimetric maintaining an inert nitrogen environment, as shown in Figure
studies with and without moisture make it possible to establish 2. In this experiment, transmittance data are recorded as the
the parameters necessary to kinetically trap the black phase of
CsPbI3 in its metastable form and clarify the deleterious role
played by moisture.
Films and powders are fabricated according to a modified
literature procedure5,25 in which solvent evaporation of a
solution of CsI and PbI2 leads to the formation of the
nonfunctional orthorhombic or “yellow” phase (δ-CsPbI3).
Similar to bulk CsPbI3 crystals16 the polycrystalline samples
undergo phase change above 305 °C and convert into a
photoactive black phase (α-CsPbI3). To maintain the desired α-
CsPbI3 phase the films are cooled rapidly (“quenched”) to
capture the metastable black phase at room temperature. These
steps are performed in the inert atmosphere of a nitrogen
glovebox because otherwise, in the ambient atmosphere, α-
CsPbI3 transforms into δ-CsPbI3 in minutes. For powder
samples, a drop-cast film is scraped off the substrate using a
spatula and then ground to a powder. (Details can be found in
the Supporting Information (SI).) Figure 2. Contour plot of transmittance spectra as a function of
X-ray diffraction measurements are performed for the drop- temperature, taken for a thin film of metastable, black-phase α-CsPbI3.
cast films, indicating that quenched samples, encapsulated in Spectra are taken in 5 °C increments over the course of several hours,
nitrogen and black to the eye, are indeed the metastable, cubic achieving an effective scan rate of <1 °C/min. See text labels for phase
descriptions.
α-phase (Figure 1b, black curve). Samples at equilibrium with
ambient atmosphere and yellow to the eye are the
orthorhombic, nonperovskite δ-phase (Figure 1b, red curve).
The moisture dependence of phase transition is shown in temperature is raised in 5 °C increments over several hours.
Figure 1c, where transmittance spectra (shown as the negative Beginning at room temperature, the initially black films absorb
logarithm of transmittance) are collected over a period of 75 across the visible spectrum. Above 200 °C the metastable phase
min for a thin film of CsPbI3 exposed to a 33% relative transforms to a wide-bandgap structure. This indicates thermal
humidity (RH) nitrogen atmosphere at 23 °C. At the start, the destabilization of the metastable phase, in favor of the
spectrum (black curve) represents black α-CsPbI3 with a equilibrium phase for this temperature range, the yellow δ-
bandgap ∼1.73 eV. In contrast, as the film transforms to yellow phase. Importantly, this phenomenon occurs even in the
δ-phase, a strong absorption feature is observed at 420 nm absence of moisture. Above the phase-transition temperature,
(orange curve). The phase transition proceeds via the the black phase becomes the equilibrium phase and the
formation of a nucleation site of yellow δ-phase in the spectrum transitions back to its original state. To understand
1279 DOI: 10.1021/acs.jpclett.7b00134
J. Phys. Chem. Lett. 2017, 8, 1278−1282
The Journal of Physical Chemistry Letters Letter

spectral changes seen as higher temperatures, we turn to


thermogravimetric analysis (TGA).
TGA analysis reveals (Figure 3, black curve) that the thermal
decomposition of CsPbI3 begins above 400 °C. This is

Figure 4. DSC curves for the CsPbI3 powders, sealed under N2


atmosphere, heating at 2 °C/min. The “black” and “yellow” labels
indicate the starting phase of the powders (α-CsPbI3 and δ-CsPbI3,
respectively). The curves are offset for clarity. The inset shows an
Figure 3. TGA heating profiles for CsPbI3 and PbI2 in black, and red, expanded and baseline-corrected view of the temperature range where
respectively. The weight loss is calculated relative to the weight at the the quenched metastable black phase transforms to a stable yellow
final temperature of 800 °C, before which both curves exhibit a phase. The arrows indicate the direction of heating/cooling.
plateau.
peak due to the yellow-to-black phase transition. On cooling,
significantly higher than, for example, MAPbI3, in which the the existence of a broad exothermic feature centered about 277
initial decomposition is governed by the release of HI and °C signifies the reemergence of the yellow phase. The
CH3NH2 and begins by ∼300 °C.26 For CsPbI3 the initial mass asymmetry of the exothermic peak could be due to the
loss begins in approximately the same temperature range as presence of some transient intermediate crystal structure, as
PbI2 sublimation (Figure 3, red curve). The two distinct mass- attributed in previous studies of MAPbI3.23,26
loss features in the TGA curve for CsPbI 3 coincide The enthalpy changes of the forward and the reverse phase
approximately with literature values for the temperatures at transition can be quantified by measuring the area under the
which pure PbI2 and CsI reach an appreciable vapor pressure exothermic or endothermic features. From DSC analysis (see
(arbitrarily defined as 200 Pa), at 493 and 762 °C, Table 1) the transition enthalpies are calculated to be 14.2
respectively.27 This analysis allows us to assign the spectral (±0.5) kJ/mol for both types of samples and for both
changes above 450 °C in Figure 2 to the formation of residual exothermic and endothermic peaks. We have executed three
CsI, coinciding with the loss of PbI2. We speculate that the consecutive scanning cycles for the same sample and found that
broadly absorbing material observed in the temperature range the individual peaks overlap and the transition energies are
between 420 to 450 °C is due to a PbI2-depleted intermediate identical within the experimental certainty (Figure S1). This
that forms transiently. As a consequence of the high- repeatability confirms the reversibility of the phase transitions
temperature threshold for sublimation, it is possible to directly and the absence of significant material losses, either to
measure the reversible enthalpies of phase changes using DSC. evaporation or chemical reaction with the crucible walls. To
To quantify the free-energy landscape of the CsPbI3 crystal the best of our knowledge this represents the first measurement
phases we employed DSC. Aluminum DSC crucibles are filled of this energy difference, and as such it can form a valuable
with powder samples, and except where otherwise noted, benchmark for first-principles calculations of stability. The large
crimped inside a N2-filled glovebox to control the composition magnitude of this enthalpy is interesting from two perspectives:
of the head space of the crucible. The yellow sample is obtained First, it demonstrates that the observed metastability of the
by exposing a black-phase powder sample to moist ambient black phase at room temperature persists despite a significant
(22−23 °C, RH 20−40%) and subsequently transferring it back driving force to convert to yellow (>5kBT per formula unit at
into the glovebox prior to sealing the crucible. The heating and room temperature, where kB is the Boltzmann constant);
cooling curves are obtained sequentially by heating the sample second, it highlights the fact that the entropic favorability of the
from 20 to 350 °C and then cooling it back to the starting black phase must be quite substantial to reverse the relative
point. Figure 4 represents the DSC thermograms for both black stability of the two phases at the moderate phase-transition
and yellow samples at a 2 °C/min scan rate. While the yellow temperature of ∼300 °C.
sample, being the stable phase, shows no sign of thermal To draw further insight from calorimetry, we note that
activity at temperatures below 200 °C, the quenched black although the phase transition is reversible, the exothermic and
sample undergoes an exothermic change (i.e., positive heat flow endothermic peak centers are not aligned and there is a
out of the sample) between 90 and 150 °C, magnified in the considerable amount of thermal hysteresis. To understand how
inset to Figure 4. The negative enthalpy of change indicates the rate of temperature change and atmospheric moisture affect
that the rise in temperature provides enough thermal energy to the phase change kinetics and particularly the kinetic trapping
overcome the energy barrier and allows the metastable black of desired black phase, DSC analysis is performed at a higher
phase to convert into a stable yellow phase. We speculate that temperature ramp rate (10 °C/min). At the faster scan rate, we
the relatively greater range of metastability (>200 °C) observed see qualitatively similar behavior to that shown in Figure 4 but
in Figure 2 is due to the formation of smaller grain sizes5 in with substantial differences in the transition peak locations,
spincast films in comparison with powders made by peak width, and onset temperature of the phase change (Figure
dropcasting. As temperature continues to rise, at 321 °C both S2; see also Table 1). The exothermic peak corresponding to
samples shown in Figure 4 exhibit a strong narrow endothermic the destabilization of the quenched black-phase sample now
1280 DOI: 10.1021/acs.jpclett.7b00134
J. Phys. Chem. Lett. 2017, 8, 1278−1282
The Journal of Physical Chemistry Letters Letter

Table 1. Phase-Transition Enthalpies of CsPbI3 Powders under Varied Conditions


exothermic endothermic
phase/atmosphere rate (°C/min) peak (°C) fwhm (°C)a |ΔH| (kJ/mol) |ΔS| (J/K·mol)b peak (°C) fwhm (°C) |ΔH| (kJ/mol) |ΔS| (J/K ·mol)b
black − N2 10 256.7 14.6 15.1 29 324.5 3.8 13.8 23
2 278.7 6.6 14.4 27 321.7 2.3 14.0 24
yellow − N2 10 259.9 19.4 15.0 28 325.3 3.6 13.8 23
2 275.1 6.5 14.6 27 321.8 2.2 13.9 23
yellow − air 10 267.6 8.9 14.6 27 325.8 3.7 14.0 23
2 281.8 4.5 14.2 26 323.1 2.0 13.7 23
a
fwhm: full width at half-maximum. bApproximate entropy values were calculated from ΔG = ΔH − TΔS using the temperature at the transition
peak.

extends over a temperature range of 110−185 °C. To facilitate increased sharpness of the peaks at either scan rate. At 2 °C/
the comparison for different scanning rates, the ordinate (y min, in the presence of water molecules, the reappearance of
axis) is transformed to heat capacity rather than heat flow. the yellow phase from the black phase (Figure S3b) appears 4
Figure 5 exhibits a significant shift toward lower onset to 5 °C before the transition temperatures in a dry atmosphere.
Most importantly, however, the enthalpy of the phase change is
the same with or without water molecules present, and no new
distinct enthalpic features are observed anywhere else in the
thermogram. By contrast, in OHPs, irreversible enthalpic
features in the thermogram have been observed and
hypothesized to originate from H2O desorption.23 In computa-
tional studies of moisture-induced instability of OHPs, the
emphasis has been on the energetics of complex formation
between the perovskite and H2O.28,29 For the present results,
no such H2O adducts are evidenced, and the observed effect
can be explained simply as a transitory stabilization of the
transition state, with no effect on the equilibrium enthalpies of
the two phases.
Figure 5. Effect of cooling rate and moisture on phase transition. The The enthalpy and the temperature of the reversible phase
exothermic transition for yellow phase δ-CsPbI3 powders, sealed either transition between the functional perovskite α-phase of CsPbI3
under N2 atmosphere (red curves) or in moist atmosphere (blue and the nonfunctional yellow δ-phase was measured as a
curves) at cooling rate of 2 °C/min (solid lines) and 10 °C/min function of atmosphere, heating rate, and sample history. The
(dashed lines), respectively. Arrow indicates temperature scanning black α-phase is intrinsically unstable at room temperature,
direction. even in a moisture-free atmosphere. Only above 321 °C is it
spontaneously formed from the δ-phase. In the absence of
temperature and increased breadth of the transition feature moisture, the α-phase can be rapidly cooled or “quenched” to
with faster cooling. At all temperatures below the phase- form a metastable functional material at room temperature,
transition temperature, the yellow phase is thermodynamically which remains metastable for significant periods of time, up to
favored, yet the system is kinetically trapped in the black phase ∼100 °C. Moisture, which was previously known to destabilize
over the temperature interval defined by the magnitude of the the α-phase, is shown here to play a primarily catalytic role,
hysteresis, which itself depends on the cooling rate. At the limit lowering the kinetic barrier to the phase transition without
of the extremely fast cooling (>200 °C/min) used to quench impacting the enthalpy significantly. CsPbI3 shows promise as
the samples in the metastable black phase, the system is an all-inorganic alternative to the hybrid halide perovskite
depleted of the thermal energy it would require to overcome photovoltaic materials; however, to realize this potential,
the activation barrier on the time scale of observation. Thus the methods such as those demonstrated herein will be critical to
system becomes frozen in a metastable state. It is interesting to understand the phase stability of α-CsPbI3 and related halide
note here that only small hysteresis values (1 to 2 °C) have perovskites.
been previously reported for MAPbI3 for the tetragonal to cubic
phase transition, and even those have been ascribed to presence
of moisture.23

*
ASSOCIATED CONTENT
S Supporting Information
It is known that moisture has the effect of destabilizing the
The Supporting Information is available free of charge on the
metastable black phase of CsPbI3; however, to identify the
ACS Publications website at DOI: 10.1021/acs.jpclett.7b00134.
mechanistic role of moisture in the phase transition, DSC
analysis is repeated in the presence of moisture. Figure 5 Experimental procedures, additional DSC plots, and
depicts the exothermic transition during the cooling cycle for method for baseline correction and Gaussian fitting for
yellow phase δ-CsPbI3 with and without a humid atmosphere. transition peaks. (PDF)


To achieve a humid environment, a yellow phase CsPbI3
powder sample is sealed in a DSC crucible in open atmosphere
AUTHOR INFORMATION
(22 to 23 °C, RH = 40%). The striking differences between the
N2- (red curve) and air- (blue curve) filled samples are the Corresponding Author
earlier transition onsets (i.e., at higher temperature) and the *E-mail: fafarman@drexel.edu.
1281 DOI: 10.1021/acs.jpclett.7b00134
J. Phys. Chem. Lett. 2017, 8, 1278−1282
The Journal of Physical Chemistry Letters Letter

ORCID (14) Goldschmidt, V. M. Die Gesetze der Krystallochemie.


Subham Dastidar: 0000-0002-3025-4787 Naturwissenschaften 1926, 14, 477−485.
(15) Møller, C. K. Crystal Structure and Photoconductivity of
Notes Cæsium Plumbohalides. Nature 1958, 182, 1436−1436.
The authors declare no competing financial interest. (16) Trots, D. M.; Myagkota, S. V. High-Temperature Structural

■ ACKNOWLEDGMENTS
S.D. and A.T.F. acknowledge funding from CBET-1604293.
Evolution of Caesium and Rubidium Triiodoplumbates. J. Phys. Chem.
Solids 2008, 69, 2520−2526.
(17) Choi, H.; Jeong, J.; Kim, H.-B.; Kim, S.; Walker, B.; Kim, G.-H.;
Kim, J. Y. Cesium-Doped Methylammonium Lead Iodide Perovskite
A.D.D. acknowledges funding from CMMI-1463412. C.J.H., Light Absorber for Hybrid Solar Cells. Nano Energy 2014, 7, 80−85.
A.D.G.-P., and J.E.S. acknowledge support from ONR under (18) Dastidar, S.; Egger, D. A.; Tan, L. Z.; Cromer, S. B.; Dillon, A.
N00014-14-1-0761. The TGA was performed in Prof. Giuseppe D.; Liu, S.; Kronik, L.; Rappe, A. M.; Fafarman, A. T. High Chloride
R. Palmese’s lab. We thank Weichun Huang from Prof. Doping Levels Stabilize the Perovskite Phase of Cesium Lead Iodide.
Christopher Li’s group for helping with the DSC measurement. Nano Lett. 2016, 16, 3563−3570.


(19) Protesescu, L.; Yakunin, S.; Bodnarchuk, M. I.; Krieg, F.;
REFERENCES Caputo, R.; Hendon, C. H.; Yang, R. X.; Walsh, A.; Kovalenko, M. V.
Nanocrystals of Cesium Lead Halide Perovskites (CsPbX3, X = Cl, Br,
(1) Saliba, M.; Matsui, T.; Domanski, K.; Seo, J.-Y.; Ummadisingu, and I): Novel Optoelectronic Materials Showing Bright Emission with
A.; Zakeeruddin, S. M.; Correa-Baena, J.-P.; Tress, W. R.; Abate, A.; Wide Color Gamut. Nano Lett. 2015, 15, 3692−3696.
Hagfeldt, A.; et al. Incorporation of Rubidium Cations into Perovskite (20) Wang, C.; Chesman, A. S. R.; Jasieniak, J. J. Stabilizing the Cubic
Solar Cells Improves Photovoltaic Performance. Science 2016, 354, Perovskite Phase of CsPbI3 Nanocrystals by Using an Alkyl
206−209. Phosphinic Acid. Chem. Commun. 2017, 53, 232−235.
(2) Misra, R. K.; Aharon, S.; Li, B.; Mogilyansky, D.; Visoly-Fisher, I.; (21) Beal, R. E.; Slotcavage, D. J.; Leijtens, T.; Bowring, A. R.; Belisle,
Etgar, L.; Katz, E. A. Temperature- and Component-Dependent R. A.; Nguyen, W. H.; Burkhard, G. F.; Hoke, E. T.; McGehee, M. D.
Degradation of Perovskite Photovoltaic Materials under Concentrated Cesium Lead Halide Perovskites with Improved Stability for Tandem
Sunlight. J. Phys. Chem. Lett. 2015, 6, 326−330. Solar Cells. J. Phys. Chem. Lett. 2016, 7, 746−751.
(3) Cheng, Z.; Lin, J. Layered Organic−inorganic Hybrid Perov- (22) Sutton, R. J.; Eperon, G. E.; Miranda, L.; Parrott, E. S.; Kamino,
skites: Structure, Optical Properties, Film Preparation, Patterning and B. A.; Patel, J. B.; Hörantner, M. T.; Johnston, M. B.; Haghighirad, A.
Templating Engineering. CrystEngComm 2010, 12, 2646−2662. A.; Moore, D. T.; et al. Bandgap-Tunable Cesium Lead Halide
(4) Noh, J. H.; Im, S. H.; Heo, J. H.; Mandal, T. N.; Seok, S. I. Perovskites with High Thermal Stability for Efficient Solar Cells. Adv.
Chemical Management for Colorful, Efficient, and Stable Inorganic− Energy Mater. 2016, 6, 1502458.
Organic Hybrid Nanostructured Solar Cells. Nano Lett. 2013, 13, (23) Baikie, T.; Fang, Y.; Kadro, J. M.; Schreyer, M.; Wei, F.;
1764−1769. Mhaisalkar, S. G.; Graetzel, M.; White, T. J. Synthesis and Crystal
(5) Eperon, G. E.; Paternò, G. M.; Sutton, R. J.; Zampetti, A.; Chemistry of the Hybrid Perovskite (CH3NH3)PbI3 for Solid-State
Haghighirad, A. A.; Cacialli, F.; Snaith, H. J. Inorganic Caesium Lead Sensitised Solar Cell Applications. J. Mater. Chem. A 2013, 1, 5628−
Iodide Perovskite Solar Cells. J. Mater. Chem. A 2015, 3, 19688− 5641.
19695. (24) Chen, T.; Foley, B. J.; Park, C.; Brown, C. M.; Harriger, L. W.;
(6) Swarnkar, A.; Marshall, A. R.; Sanehira, E. M.; Chernomordik, B. Lee, J.; Ruff, J.; Yoon, M.; Choi, J. J.; Lee, S.-H. Entropy-Driven
D.; Moore, D. T.; Christians, J. A.; Chakrabarti, T.; Luther, J. M. Structural Transition and Kinetic Trapping in Formamidinium Lead
Quantum Dot−induced Phase Stabilization of α-CsPbI3 Perovskite for Iodide Perovskite. Sci. Adv. 2016, 2, e1601650.
High-Efficiency Photovoltaics. Science 2016, 354, 92−95. (25) Eperon, G. E.; Stranks, S. D.; Menelaou, C.; Johnston, M. B.;
(7) Frolova, L. A.; Anokhin, D. V.; Piryazev, A. A.; Luchkin, S. Y.; Herz, L. M.; Snaith, H. J. Formamidinium Lead Trihalide: A Broadly
Dremova, N. N.; Stevenson, K. J.; Troshin, P. A. Highly Efficient All- Tunable Perovskite for Efficient Planar Heterojunction Solar Cells.
Inorganic Planar Heterojunction Perovskite Solar Cells Produced by Energy Environ. Sci. 2014, 7, 982−988.
Thermal Coevaporation of CsI and PbI2. J. Phys. Chem. Lett. 2017, 8, (26) Dualeh, A.; Gao, P.; Seok, S. I.; Nazeeruddin, M. K.; Grätzel, M.
67−72. Thermal Behavior of Methylammonium Lead-Trihalide Perovskite
(8) Lee, J.-W.; Kim, D.-H.; Kim, H.-S.; Seo, S.-W.; Cho, S. M.; Park, Photovoltaic Light Harvesters. Chem. Mater. 2014, 26, 6160−6164.
N.-G. Formamidinium and Cesium Hybridization for Photo- and (27) Kaye, G. W. C.; Laby, T. H. Tables of Physical and Chemical
Moisture-Stable Perovskite Solar Cell. Adv. Energy Mater. 2015, 5, Constants and Some Mathematical Functions, 16th ed..; Longman Sc &
1501310. Tech: New York, 1995.
(9) Li, W.; Li, J.; Niu, G.; Wang, L. Effect of Cesium Chloride (28) Tong, C.-J.; Geng, W.; Tang, Z.-K.; Yam, C.-Y.; Fan, X.-L.; Liu,
Modification on the Film Morphology and UV-Induced Stability of J.; Lau, W.-M.; Liu, L.-M. Uncovering the Veil of the Degradation in
Planar Perovskite Solar Cells. J. Mater. Chem. A 2016, 4, 11688− Perovskite CH3NH3PbI3 upon Humidity Exposure: A First-Principles
11695. Study. J. Phys. Chem. Lett. 2015, 6, 3289−3295.
(10) Li, Z.; Yang, M.; Park, J.-S.; Wei, S.-H.; Berry, J. J.; Zhu, K. (29) Zhang, L.; Ju, M.-G.; Liang, W. The Effect of Moisture on the
Stabilizing Perovskite Structures by Tuning Tolerance Factor: Structures and Properties of Lead Halide Perovskites: A First-
Formation of Formamidinium and Cesium Lead Iodide Solid-State Principles Theoretical Investigation. Phys. Chem. Chem. Phys. 2016,
Alloys. Chem. Mater. 2016, 28, 284−292. 18, 23174−23183.
(11) Kulbak, M.; Gupta, S.; Kedem, N.; Levine, I.; Bendikov, T.;
Hodes, G.; Cahen, D. Cesium Enhances Long-Term Stability of Lead
Bromide Perovskite-Based Solar Cells. J. Phys. Chem. Lett. 2016, 7,
167−172.
(12) McMeekin, D. P.; Sadoughi, G.; Rehman, W.; Eperon, G. E.;
Saliba, M.; Hörantner, M. T.; Haghighirad, A.; Sakai, N.; Korte, L.;
Rech, B.; et al. A Mixed-Cation Lead Mixed-Halide Perovskite
Absorber for Tandem Solar Cells. Science 2016, 351, 151−155.
(13) Niemann, R. G.; Gouda, L.; Hu, J.; Tirosh, S.; Gottesman, R.;
Cameron, P. J.; Zaban, A. Cs + Incorporation into CH 3 NH 3 PbI 3
Perovskite: Substitution Limit and Stability Enhancement. J. Mater.
Chem. A 2016, 4, 17819−17827.

1282 DOI: 10.1021/acs.jpclett.7b00134


J. Phys. Chem. Lett. 2017, 8, 1278−1282

View publication stats

You might also like