Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Article

Cite This: Chem. Mater. 2019, 31, 8563−8575 pubs.acs.org/cm

Layered Lead Iodide of [Methylhydrazinium]2PbI4 with a Reduced


Band Gap: Thermochromic Luminescence and Switchable Dielectric
Properties Triggered by Structural Phase Transitions
Mirosław Mączka,*,† Maciej Ptak,† Anna Gągor,† Dagmara Stefańska,† and Adam Sieradzki‡

Institute of Low Temperature and Structure Research, Polish Academy of Sciences, P.O. Box 1410, 50-950 Wrocław 2, Poland

Department of Experimental Physics, Wrocław University of Science and Technology, Wybrzeże Wyspiańskiego 27, 50-370
Wrocław, Poland
Downloaded via CTR FOR NANO & SOFT MATTER SCIENCES on October 10, 2023 at 12:51:52 (UTC).

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

ABSTRACT: Layered lead halides offer potential for light-


emitting and photovoltaic devices. Here, we report the synthesis
of two-dimensional (2D) lead iodide with single-octahedra slabs
separated by the smallest organic cation used to date in the
construction of A2PbI4 compounds. This compound, MHy2PbI4,
where MHy+ denotes methylhydrazinium cation, shows exception-
ally short separation of adjacent lead iodide layers and strong N−
H···I hydrogen bonds as well as unusually small interoctahedral
tilting of PbI6 octahedra, leading to exceptionally small band gap of
2.20 eV. Photoluminescence (PL) studies show that this compound exhibits at room temperature two emission bands at 561
and 610 nm related to free exciton and bound exciton. The latter one exhibits blue shift on cooling by around 37 nm in the 80−
300 K range. As a result, PL of MHy2PbI4 changes strongly with temperature from yellowish pink at 300 K to yellow-green at 80
K. MHy2PbI4 undergoes three structural phase transitions. The first phase transition occurs at 320 K, and it does not show any
evident change of Pmmn symmetry. The second phase transition that occurs at 298 K is associated with ordering of the MHy+
cations and change of symmetry to Pccn. This phase transition leads also to pronounced steplike change of dielectric
permittivity. Structural investigation indicates that major reorientation of MHy+ cations and tilts of PbI6 octahedra contribute to
this switchable dielectric property. The third phase transition observed at 262 K (233 K) on cooling (heating) leads to
distortion of the structure to triclinic, space group P1̅. MHy2PbI4 shows a low value of direct current conductivity σDC at low
temperatures, but the conductivity increases rapidly with increasing temperature and reaches 1.02 × 10−7 S m−1 at 298 K. We
also report temperature-dependent Raman scattering of MHy2PbI4. Analysis of Raman data revealed clear shifts and changes in
bandwidths at the phase transitions that provided deeper insight into mechanisms of the structural phase transitions in this
material. An interesting feature of the studied perovskite is also unusually large broadening of many bands on heating the sample
from 80 to 300 K that is much larger than typically observed in A2PbI4 compounds. This behavior proves that lattice dynamic
effects play very important role in MHy2PbI4 and that the nature of organic cation has a great impact on the lattice dynamics of
2D perovskites.

■ INTRODUCTION
Organic−inorganic metal halides have attracted a great interest
advantages for easy fabrication of ultrathin films and
nanocrystalline materials by solution-based method and
in recent years due to their remarkable and tunable properties mechanical exfoliation.11,12 Unfortunately, such 2D materials
that make them very suitable materials for a variety of exhibit much smaller power conversion efficiencies than 3D
applications.1−7 One group constitutes three-dimensional (3D) analogues due to the poor charge transport related to the
perovskites with formula ABX3, where A, B, and X denote presence of isolating organic spacers.3,10 However, 2D hybrids
monovalent cation (usually methylammonium (MA), for- are very promising materials for applications as photodetectors
mamidinium (FA), or alkali metal), divalent cation (Pb2+, Sn2+, and light-emitting diodes due to color tunability and often high
or Ge2+), and halogen (Cl−, Br−, I−).1,2,5,7 These compounds efficiency of photoluminescence (PL).3,6,13 Furthermore, 2D
have been rapidly developed for applications as solar cells with perovskites have often a notable optical nonlinearity.14
power conversion efficiencies exceeding 20% and as light- Two-dimensional layered halides are composed of metal
emitting diodes.2,5,7 Although 3D perovskites exhibit out- halide layers separated by organic cations. In contrast to 3D
standing optoelectronic properties, they exhibit low chemical
stability at ambient conditions.8 One of the strategies for Received: September 13, 2019
enhancing chemical stability is decreasing the dimensionality Revised: September 24, 2019
to two-dimensional (2D).9,10 Intrinsic layered structure is also Published: September 25, 2019

© 2019 American Chemical Society 8563 DOI: 10.1021/acs.chemmater.9b03775


Chem. Mater. 2019, 31, 8563−8575
Chemistry of Materials Article

perovskites, use of organic cations with different lengths, sizes, at RT and appeared below 170 K.25 Another representative of
shapes, and strengths of electrostatic interactions as well as (100) perovskites, CsGAPbI4, showed an asymmetric band at
hydrogen bonds (HBs) with the inorganic layers leads to a 77 K with maximum near 535 nm and small Stokes shift.21
remarkable structural tunability of 2D compounds.13 For This PL was not measurable at RT. Interestingly, this
instance, single-octahedra slabs may be flat or corrugated and compound showed also a weaker and very broad band with
the PbX6 octahedra may exhibit different degree of distortion. maximum near 730 nm. Such broad PL bands with large
It is worth adding that the inorganic layers contribute to the Stokes shifts appear often for compounds exhibiting large
semiconductor electronic structure, and due to the separation, degree of octahedral distortion, especially for (110) com-
they form quantum-like structures with 2D electronic confine- pounds, and they are usually attributed to self-trapped
ment. This leads to the existence of stable excitons at room excitons.21−23,25,29−34 Regarding (110) iodides with short
temperature (RT) with large binding energy of a few hundreds separation of inorganic layers, GA2PbI4 showed broad PL at
of meV, which is an order of magnitude higher compared to 598 nm (100 nm Stokes shift) while PL of FAGAPbI4 (FA =
3D lead halide perovskites, as well as to blue shift of band gap formamidinium) consisted of a few bands in the 600−700 nm
compared to 3D analogues.15−17 The presence of organic range (Stokes shift, 90−190 nm).22,23 It is worth adding that
spacers leads also to dielectric confinement that makes the broad PL is attractive for applications since it may lead to
exciton binding energy even larger.18 Such large exciton white emission.29−35
binding energy and large band gaps limit photovoltaic Apart of differences in structural connectivity and distortion,
performance of 2D perovskites. It is, therefore, important to PL and optoelectronic properties can also be affected by the
find materials with lower electronic and dielectric confinement. dynamic of the organic cations.36−38 It is therefore of great
For lead halide hybrids containing flat single-octahedra layers, interest to perform temperature-dependent studies of hybrid
the electronic confinement can be reduced by decreasing the halides, especially those undergoing order−disorder phase
organic cation length. For instance, studies of 2D A2PbI4(100) transitions, using various experimental methods that have
perovskites with A = benzylammonium, 2-phenethylammo- different probing length scales and sensitivities. One of such
nium (PEA), and 4-phenyl-1-butylammonium cations showed methods sensitive to the order−disorder phenomena, defects,
that decrease of the organic cation length leads to shift of the and local symmetry changes is Raman spectroscopy.17,38−40
band gap from 2.40 to 2.34 eV.13 The electronic confinement Literature data show that employment of small organic
is also affected by distortion of the PbX6 layer, i.e., it decreases cations in the synthesis of 2D lead halides may result in
with decreasing distortion.19,20 For instance, it was shown that discovery of novel materials with promising functional
weakly distorted (100) GAMAPbI4 and CsGAPbI4 perovskites properties. FA+ cation with ionic radius of 253 pm is small
(GA and MA denote guanidinium and methylammonium enough to stabilize 3D perovskite structure (tolerance factor,
cations, respectively) have smaller band gaps, i.e., 2.27 and 2.35 TF = 0.99, only slightly smaller than 1), while GA+ with ionic
eV,20,21 than strongly distorted (110) GA2PbI4, FAGAPbI4, or radii 278 pm is large enough to stabilize 2D A2PbI4 structure
ImEAPbI4 analogues (FA = formamidinium and ImEA = (TF = 1.04).1,22,41,42 There are four other organic ions with
imidazolium ethylammonium) that showed band gap of 2.45− ionic radii between FM+ and GA+, i.e., imidazolium (IM+, 258
2.54 eV.22−24 Reduction of the dielectric confinement can be pm), methylhydrazinium (MHy+, 264 pm), dimethylammo-
achieved by using high-dielectric-permittivity organic compo- nium (DMA+, 272 pm), and ethylammonium (EtA+, 274
nents.25 It was shown that lead iodide containing ethanolamine pm).1,43 EtA+, DMA+, and IM+ form iodides of APbI3
(EA) with high dielectric permittivity of 37 has strongly composition, but 2D compounds of A2PbI4 compositions are
reduced dielectric confinement and the exciton binding energy not known.44−46 MHy+ have not been yet employed in
20 times smaller compared to conventional 2D perovskites.25 construction of any lead halide hybrid. We have decided,
As a result, EA 2 PbI 4 is expected to have improved therefore, to undertake efforts to prepare novel lead iodide
optoelectronic properties, especially for solar cells and containing MHy+ cations. It is worth emphasizing that there
photodetectors.25 are only a few organic−inorganic hybrid perovskites containing
In addition to band gap tunability, A2PbX4 compounds with methylhydrazinium cations, i.e., methylhydrazinium metal
well-separated inorganic layers often exhibit strong excitonic formates that were shown to exhibit ferroelectric and
PL even at room temperature.13,26,27 When structural multiferroic properties.43,47,48 We will show that this small
distortion is weak, such PL bands related to free exciton and high-dielectric-permittivity organic cation of 2249 can be
(FE) are narrow and the Stokes shift is small.13,24,26,27 When employed in the synthesis of 2D layered perovskite of formula
structural distortion is important, red-shifted PL bands may MHy2PbI4. This compound undergoes three structural phase
appear. For instance, two bands were observed for cystamine- transitions, shows small band gap, and exhibits PL as well as
based lead iodide. The narrow band at 522 nm overlapping switchable dielectric behavior near 300 K. It also shows
with the excitonic absorption feature was attributed to FE, pronounced temperature dependence of PL and Raman bands.
while the broader one at 554 nm was attributed to bound We discuss the origin of this behavior and mechanism of the
exciton (BE).28 Situation may become much more compli- phase transitions.


cated in the case of 2D perovskites with short separation of
inorganic layers due to reduced quantum and dielectric EXPERIMENTAL SECTION
confinement. For instance, narrow excitonic PL with a few
nanometers Stokes shifts was reported at 558 and 550 nm for Synthesis. PbI2 (99%, Sigma-Aldrich), methylhydrazine (98%,
Sigma-Aldrich), hydroiodic acid (57 wt % in H2O, 99.95% Sigma-
(100) perovskites 3AMPPbI4 [3AMP = 3-(aminomethyl)- Aldrich), methylacetate (99.5%, Sigma-Aldrich), and ethanol (95%,
piperidinium] and GAMAPbI4, respectively, but this PL was Sigma-Aldrich) were commercially available and used without further
weak at room temperature.20,27 (100) perovskite EA2PbI4 also purification. Single crystals were grown using antisolvent vapor-
revealed excitonic PL with very small Stokes shift, but due to assisted crystallization, in which the appropriate antisolvent is slowly
extremely reduced dielectric confinement, this PL was absent diffused into a solution containing the crystal precursors. The crystal

8564 DOI: 10.1021/acs.chemmater.9b03775


Chem. Mater. 2019, 31, 8563−8575
Chemistry of Materials Article

precursor was prepared by adding 3 mmol PbI2 to a solution 7. The triclinic phase was heavily twinned with overlapping diffraction
containing 9 mmol methylhydrazine dissolved in 6 mL of HI. The peaks from at least three domains. The model of the structure was
clear solution obtained after stirring for 15 min was placed into a glass obtained from one dominating domain indexing ∼50% of diffracted
vial, which was placed in a second larger glass vial containing ethanol intensities. This resulted in disturbed intensities and high (max ∼ 4 e
and methylacetate in 1:1 ratio. The lid of the outer vial was A−3) residual densities around heavy atoms and negative anisotropic
thoroughly sealed, but the lid of the inner vial was loosened to allow displacement parameters of the organic part. Thus, all C and N atoms
diffusion of the ethanol and methylacetate into the precursor solution. were refined using the isotropic displacement parameters.
Red platelets were harvested after 4 days, filtered from the mother Raman and IR Studies. Temperature-dependent Raman spectra
liquid, and dried at room temperature. A good match of their powder on a single crystal were measured using a Renishaw InVia Raman
X-ray diffraction (XRD) patterns with the calculated ones based on spectrometer equipped with a confocal DM 2500 Leica optical
the single-crystal data (Figure S1, Supporting Information) confirmed microscope and a thermoelectrically cooled CCD as a detector.
the phase purity of the bulk sample. Additional measurements of low-wavenumber modes (down to 5
Differential Scanning Calorimetry (DSC) and Thermogravi- cm−1) were performed on the same instrument using an eclipse filter.
metric Analysis (TGA) Measurements. Heat capacity was Since excitation with an argon laser operating at 488 or 514 nm
measured using a Mettler Toledo DSC-1 calorimeter with high resulted in a very strong luminescence background, a diode laser
resolution of 0.4 μW. Nitrogen was used as a purging gas, and the operating at 830 nm was used for Raman studies. The temperature
heating and cooling rate was 5 K min−1. The sample weight was 26.11 was controlled using Linkam cryostat cell. The spectral resolution was
mg. The excess heat capacity associated with the phase transitions was 2 cm−1.
evaluated by subtraction from the data the baseline representing
variation in the absence of the phase transitions. TGA study was
performed in the temperature range 300−1073 K using PerkinElmer
TGA 4000. The sample weight was ca. 23.4 mg, and the heating speed
■ RESULTS AND DISCUSSION
DSC and TGA. The DSC measurements show the presence
was 10 K min−1. Pure nitrogen gas as an atmosphere was used. of three heat anomalies at T1 = 320 K (318 K), T2 = 299 K
X-ray Powder Diffraction. Powder XRD pattern was obtained on (295 K), and T3 = 262 K (233 K) during heating (cooling)
an X’Pert PRO X-ray diffraction system equipped with a PIXcel (Figures 1 and S2). Strongly symmetric shapes of these
ultrafast line detector and Soller slits for Cu Kα1 radiation (λ =
1.54056 Å). The powders were measured in the reflection mode, and
the X-ray tube settings were 30 mA and 40 kV.
Dielectric Properties. The complex impedance (Z*) measure-
ments were done using a Novocontrol Alpha analyzer. The complex
dielectric permittivity (ε*), electric modulus (M*), and alternating
current (AC) and direct current (DC) conductivities (σ) have been
evaluated from the measured complex impedance. The platelike single
crystal of 3.1 × 1.7 mm2 with thickness 0.4 with crystallographic
orientation perpendicular to the (001) cleavage plane (notation for
the Pmmn phase) was used. The silver paste was coated on both
parallel surfaces of the sample to ensure a good electrical contact. A
sinusoidal voltage with amplitude of 1 V and frequency in the range
101−106 Hz was applied across the sample. The temperature was
controlled by the Novocontrol Quattro system, by using a nitrogen
gas cryostat. The measurements were taken every 1 K in the
temperature range of 140−350 K. The temperature stability of the
samples was better than 0.1 K. Owing to the shape of the sample
crystal, measurements in other crystallographic directions were not
possible. We have, therefore, performed additional dielectric measure-
ments for pellet with a diameter of 6 mm and thickness of 0.4 mm
made of well-dried sample.
Optical Studies. RT diffuse reflectance spectrum of the powdered
sample was measured using the Varian Cary 5E UV−vis−NIR
spectrophotometer. Temperature-dependent emission spectra under
405 nm excitation line from a diode laser were measured with the Figure 1. Change in (a) Cp and (b) S related to the phase transitions
Hamamatsu photonic multichannel analyzer PMA-12 equipped with a in the heating (red) and cooling (blue) runs.
BT-CCD linear image sensor. The temperature of the sample was
controlled using Linkam THMS 600 heating/freezing stage. The
quantum efficiency was measured on Hamamatsu Absolute PL
anomalies, sharp changes of entropy, and large thermal
quantum yields (QYs) measurement system C9920-02G. To record hysteresis (especially the lowest-temperature phase transition
time-resolved PL spectra and decay times, a femtosecond laser at T3) point to the first-order character of these phase
(Coherent Model Libra) was used as an excitation source. transitions. The associated changes in enthalpy ΔH and
Single-Crystal X-ray Diffraction. Single-crystal X-ray diffraction entropy ΔS were estimated to be, respectively, ∼120 J mol−1
was collected at Xcalibur diffractometer operating with Atlas charge- and ∼0.41 J mol−1 K−1 for the phase transition at T1, ∼744 J
coupled device (CCD) detector and Mo Kα radiation. In all phases, mol−1 and ∼2.65 J mol−1 K−1 for the phase transition at T2,
empirical absorption correction using spherical harmonics, imple- and ∼56 J mol−1 and ∼0.23 J mol−1 K−1 for the phase
mented in SCALE3 ABSPACK scaling algorithm, was done using transition at T3 (average values).
CrysAlis PRO 1.171.38.46 (Rigaku Oxford Diffraction, 2015). The
According to our X-ray diffraction data, the MHy+ cations
structures were solved using SHELXT 2014/5 and refined in
SHELXL2018/3. Due to the disorder of MHy+ in phases I, II, and show 3-fold disorder above T2 and complete order below T3.
III, the displacement parameters of split nitrogen atoms were refined In the intermediate phase III, only one of three MHy+ ions
isotopically, additionally not all hydrogen atoms were incorporated in shows 2-fold disorder. Thus, the total change in entropy at T2
the model. In the last phase, phase IV, all hydrogen atoms were and T3 is expected to be R ln 3 = 9.13 J mol−1 K−1. The large
included from geometry; terminal NH2 groups were refined with afix (small) experimental value of ΔS at T2 (T3) is consistent with
8565 DOI: 10.1021/acs.chemmater.9b03775
Chem. Mater. 2019, 31, 8563−8575
Chemistry of Materials Article

Figure 2. (a) Details of the crystal structure of phase I, T = 330 K. The MHy+ cation is disordered over three positions. Possible N−H···I hydrogen
bonds are marked as dashed lines. (b) Tilting of PbI64− octahedra in subsequent phases, as seen along the b axis. (c) Temperature changes of the
interlayer distance between the lead iodide sheets. (d) Details of the crystal structure in the triclinic LT polymorph, T = 100 K, P1̅. All MHy+ ions
are ordered and interact with the inorganic substructure via N−H···I HBs and intermolecular N−H···N HB interactions. Only 12 independent
molecules from the asymmetric unit are drawn for the picture clarity.

large (small) change in disorder of MHy+ cations. However, and the space group changes to Pccn. The last phase transition
the total change in entropy (∼2.88 J mol−1 K−1) is smaller than to phase IV introduces small triclinic distortion of the large
expected. As discussed in the literature, some residual entropy unit cell. The triclinic angles are very close to 90° nevertheless,
is always left over if a phase transition has some relaxor the splitting of diffraction peaks clearly confirms this drastic
character.50 Thus, our results showing less-than-expected symmetry reduction. Figure S4 shows the reconstruction of the
entropy change indicate that much residual entropy is left reciprocal space before and after the transition to phase IV.
over, implying relaxor character of the phase transition. The Bragg peak profiles are the convolution of at least three
The anomaly at T1 leads to small changes in entropy, maxima from dominating domains. As the symmetry reduction
suggesting weak contribution of an order−disorder mechanism to monoclinic would result in maximum two domain states, the
to this phase transition. structure of phase IV was resolved in the triclinic system of P1̅
The thermogravimetric plot indicates that MHy2PbI4 starts space group.
to decompose at about 480 K (Figure S3). In this plot, a The crystal structure of all phases is built of the inorganic 2D
weight loss of about 44% takes place between 480 and 660 K, perovskite substructure, which can be derived from 3D
corresponding to release of methylhydrazinium iodide (the perovskites by slicing of the lattice along (100) crystallographic
calculated value is 43.1%). On further heating, PbI2 starts to planes.23,54 The PbI64− octahedra share four corners and
sublimate, and this process ends near 900 K. The expand as layers in b and c directions. Thus, in all octahedra,
decomposition temperature of MHy2PbI4 falls within the there are two apical iodides and four bridging ones. The a
range typically observed for A2PbI4 iodides (558 K for n- direction corresponds to the interlayer distance, which is equal
butylammonium, 51 543 K for EA 2 PbI 4 , 52 478 K for to a/2. The presence of n-glide plane results in the (1/2, 1/2,
[NH3(CH2)4NH3]PbI4 and [NH3(CH2)6NH3]PbI4,53 and 0) translation of neighboring layers. The MHy+ ions are
528 K for FAGAPbI4).23 located between the planes and interact with them via an
Single Crystal X-ray Diffraction. To obtain information extended system of N−H···I HB interactions and intra-
on symmetries of the observed phases, we employed X-ray molecular N−H···N HBs. In phases I and II, all cations are
diffraction method. The crystals of MHy2PbI4 possess four disordered within three positions, which are related to the
different polymorphs depending on the temperature: two high- rotations of molecules along the long axis (Figure 2a). With
temperature (HT) phases II and I, room-temperature phase III temperature lowering, the movements gradually cease. In phase
expanding down to 233 K on cooling, and low-temperature III, only one of the three MHy+ cations is disordered over two
(LT) polymorph IV, below. Phases II and I are isostructural sites; in phase IV, all cations are ordered.
and adopt orthorhombic Pmmn symmetry with the unit cell In phases I and II, the lead iodide sheets consist of only
volume near 770 Å3 (Table S1). The phase transition to phase slightly tilted octahedra. There are two distinct bridging angles
III leads to a substantial increase of the unit cell volume to Pb(1)−I(1)−Pb(1) of 171° and Pb(1)−I(3)−Pb(1) of 180°.
4575 Å3 because of a tripling of b and doubling of c lattice The I(1)−Pb(1)−I(1) angle, where I(1) is an apical iodide,
parameters. The symmetry of phase III remains orthorhombic, equals 179° (Tables 1 and S2). The high symmetry of the
8566 DOI: 10.1021/acs.chemmater.9b03775
Chem. Mater. 2019, 31, 8563−8575
Chemistry of Materials Article

Table 1. Distortion Parameters Resulting from the Single-Crystal X-ray Structures of MHy2PbI4 Perovskite in Subsequent
Polymorphic Phases
phase T (K) Pb−I−Pb direction maxa θtilt 180 − θin 180 − θout max Δd max σ2
−6
I 330 [001] 180 0 0 6.8 × 10 5.6
[010] 170.7(1) 9.3 0
II 305 [001] 180 0 0 1.3 × 10−5 8.8
[010] 170.3(1) 9.7 0
III 280 [001] 166.87(1) 13.13 0.24 1.4 × 10−4 11.9
[010] 170.99(1) 1.55 7.92
IV 100 [001] 163.43(1) 16.50 1.69 3.9 × 10−4 18.1
[010] 166.49(1) 2.53 13.47
a
Maximum distortion parameters are given because the structures of LT phases contain two (phase III) and six (phase IV) inequivalent [PbI6]4−
octahedra. Δdbond length distortion; σ2octahedral angle variance.56

Figure 3. (a) Photograph of MHy2PbI4 single crystals showing orientation of the crystallographic axes in notation for phase I (space group Pmmn).
(b−e) Details of the Raman spectra at different temperatures in a heating run for z(xx + xy)z polarization.

individual octahedra (C2v) may result from the lack of 170.7(1) to 166.49(1)° (along the [010] direction) (Tables
preference in directional bonding due to the temperature- 1 and S4).
induced dynamic distribution of NH2 and NH2+ donor groups. Phase IV is entirely ordered, Figure 2d, and all 24
The presence of HBs has a strong impact on the position of inequivalent iodides act as acceptors in HB interactions
iodide ligands.55 Those iodides, which act as acceptors for the (Table S5). Additionally, the structure is stabilized by N−
hydrogen atoms, shift from the ideal linear placements between H···N intermolecular HBs with terminal N atoms serving as
the lead atoms. The II−III phase transition brings distortions donors and acceptors for the hydrogen atoms and in-chain
within the lead iodide layers and the changes in the cations NH2 groups being hydrogen atoms donors.
distribution. The unit cell expands in both b and c directions It is important to compare structural data of A2PbI4(100)
accommodating the distortions arising from the interoctahe- perovskites with short separation of inorganic layers. According
dral tilting. Figure 2 illustrates the tilting of the PbI64− to literature data, the distances between layers are 9.288,
octahedra in subsequent phases and the temperature changes 9.4129, 9.9003, 10.2261, 10.2263, and 10.4999 Å for
of interlayer distance. The local structural distortions result in CsGAPbI4, GAMAPbI4, PyrEAPbI4 (PyrEA = pyridinium
two inequivalent PbI64− octahedra of lower symmetry, C2 and ethylammonium), 3AMPPbI 4 , EA 2 PbI 4 , and 4AMPPbI 4
C1 in phase III. The number of inequivalent cations increases [4AMP = 4-(aminomethyl)piperydinium], respectively
to three, of which every third is disordered over two possible (Table S6).20,21,24,25,27 For MHy2PbI4, this value is 9.3635 Å
sites with 0.7/0.3 ratio at 280 K. Figure S5 illustrates the (at 305 K, Table S1), i.e., smaller value is observed only for
selected details of the structure of phase III. mixed-cation CsGAPbI4. For CsGAPbI4, GAMAPbI4, Pyr-
During the last phase transition, the interoctahedral EAPbI4, 3AMPPbI4, EA2PbI4, and 4AMPPbI4 compounds, the
distortions increase and the local octahedral symmetry Pb−I−Pb angles are 149.74 + 165.19, 159.04 + 180.0, 159−
decreases. The triclinic unit cell accommodates 6 inequivalent 174, 163.39−168.83, 159.93 + 161.06, and 154.2−154.59°,
PbI64− octahedra of C1 symmetry and 12 ordered cations. The respectively (Table S6).20,21,24,25,27 Significant deviation from
octahedral distortion parameters (Table 1) increase with the 180° is also observed for many A2PbI4(100) perovskites with
temperature lowering, as it is expected from the symmetry large separation between inorganic layers, such as NBT2PbI4
reduction. The values and the differences between the phases (NBT = n-butylammonium) and PEA2PbI4 (Table S6).57,58
are, however, rather small, indicating only slight deviations Thus, MHy2PbI4 shows the smallest deviation of the Pb−I−Pb
from the ideal octahedral coordination. The main modifica- angles from 180°, and this deviation is observed for phases I
tions between the phases concern the variations in the and II only along the [010] direction.
interoctahedral tilting, which changes from 180 to max Regarding phase transitions, they have been reported for a
163.43(1)° (for Pb−I−Pb along the [001]) and from number of A2PbI4(100) perovskites with large separation
8567 DOI: 10.1021/acs.chemmater.9b03775
Chem. Mater. 2019, 31, 8563−8575
Chemistry of Materials Article

between inorganic layers due to the presence of bulky organic decreases, the Raman spectra exhibit weak changes down to
cations containing organic rings or alkyl chains.57,59−61 The 300 K. Detailed analysis of Raman wavenumbers and full width
mechanism of these transitions was related mainly to dynamic at half-maximum (FWHM) values as a function of temperature
rotational disordering of the NH3 head group or melting of the shows, however, clear anomalies at T1 (Figure S9). First, some
chains.59,61 Among (100) iodide perovskites with short bands exhibit weak discontinuous shifts at T1. The shifts are
separation between inorganic layers, phase transition was negligible for the Pb−I stretching and PbI6 librational modes,
reported only for EA2PbI4, but the structure of the HT phase but clear shifts up to 1.6 cm−1 are observed for the 1358 and
was not solved.52 As discussed above, phase transitions in 1322 cm−1 τ(NH2) modes (Figure S9a). Furthermore, the
MHy2PbI4 are associated with rotational disordering of the δ(NH2) band at 1574 cm−1 splits into two bands. Second, the
whole MHy+ cations associated with tilts and distortion of PbI6 δ(NH2) band narrows at T1 (Figure S9b). These changes
octahedral units. indicate that the phase transition at T1 is associated with subtle
Raman and IR Studies. Further information on the reorganization of MHy+ cations, which is too small to lead to
mechanism of the phase transitions and lattice dynamics can be any clear distortion of the inorganic layers and space group
obtained by analyzing temperature-dependent Raman spectra. change.
Orientation of the single crystal is shown in Figure 3a, and The phase transition at T2 is very clearly observed when the
nonresonant Raman spectra are presented in Figures 3b−e, 4, temperature decreases from 300 to 295 K. First, the PbI6
librational bands observed at 15.8 and 33.4 cm−1 split into 13.0
+ 19.8 and 29.3 + 40.8 cm−1 bands (Figures 4a and S9a),
respectively. Second, the ρ(NH2) band shifts from 857.1 to
855.3 cm−1. Furthermore, a new band appears at 866 cm−1.
Weak upshifts are also observed for the τ(NH2) modes from
1322.5 and 1357.9 cm−1 to 1324.6 and 1359.4 cm−1, and for
the τ(CH3) mode, from 337.1 to 339.1 cm−1 (Figure S9a).
Third, the 339.1 and 1357.9 cm−1 modes exhibit decrease of
FWHM values from 65.4 and 20.8 cm−1 to 52.0 and 18.2 cm−1
(Figure S9b). The observed narrowing, splitting, and shift of
bands are consistent with the first-order character of the phase
transition at T2, ordering of MHy+ cations, tilts of PbI6
octahedra, and increase of the formula units in the unit cell.
On further decrease of temperature, the Pb−I stretching
modes observed at 110.1 and 88.5 cm−1 (at 295 K) exhibit
softening to 106.1 and 83.1 cm−1 (at 260 K), followed by
Figure 4. Details of the low-wavenumber Raman spectra at different abrupt shift to 97.1 and 80.7 cm−1, respectively, at 250 K.
temperatures in a heating run for (a) z(xx + xy)z and (b) z(yy + yx)z Furthermore, a new band appears at 250 K at 76.1 cm−1.
polarization configurations. Interestingly, librations of PbI6 octahedra do not show any
significant changes in the 295−250 K range (Figure S9a).
and S6−S8. The observed Raman modes and their assignment Softening is also observed for the δ(CNN) mode, which shifts
are listened in Table S7. Former studies of lead halide from 442.6 cm−1 at 295 K to 442.0 cm−1 at 260 cm−1 and
perovskites showed that lattice modes are observed below 300 437.9 cm−1 at 250 K. The δ(NH2), νas(CNN), and νs(CNN)
cm−1, and the most intense Raman bands can be attributed to modes also exhibit significant shifts (see changes near 1580,
vibrations involving Pb2+ and I− ions.38,60,62,63 In particular, the 1090, and 885 cm−1 in Figure S9a). FWHM values show weak
Pb−I stretching modes were observed in the 70−110 cm−1 anomalies at T3. It is worth noting that the bands are still broad
range, while modes below 80 cm−1 were attributed to at 250 K, indicating significant motional freedom of MHy+
librational and deformational modes of PbI6 octahedra.60−63 cations. However, all bands related to internal vibrations of
We assign, therefore, the strong bands at 120, 112, and 90 MHy+ cations show pronounced narrowing on cooling to 80 K.
cm−1 as well as very strong bands at 34, 27, and 13 cm−1 (at The fact that the L(PbI6) modes are barely affected whereas
370 K, Figures 3, 4, and S6−S8) to the Pb−I stretching and large anomalies are observed for the Pb−I stretching modes
PbI6 librational modes, respectively. One should also expect to indicates that the phase transition at T3 is associated with
observe translational and librational modes of MHy+ ions. significant distortion of the PbI6 octahedra, whereas tilts of
These modes are overlapped by much stronger ones these octahedra seem to be weakly affected. The distortion of
corresponding to the Pb−I stretching modes and are not the inorganic layers is certainly associated with changes of the
resolved at 370 K. However, some weak bands appear at 80 K interaction strength between MHy+ and I− ions, and changes
in the 123−195 cm−1 range and at 302 cm−1 (Table S7) that in the C−N−N bond angle, as evidenced by the wavenumber
can be most likely attributed to translational and librational shifts observed for the δ(NH2), δ(CNN), νs(CNN), and
modes of MHy+. The remaining weak- and medium-intensity νas(CNN) modes.
bands observed above 200 cm−1 can be attributed to internal In summary, Raman data show that when going from 80 to
vibrations of MHy+ cation. Assignment of these modes to 370 K, the largest broadening is observed for the τ(CH3),
respective motions of atoms can be proposed based on our ρ(NH2), and δ(NH2) modes. This behavior proves that
previous Raman and IR studies of methylhydrazinium metal dynamics of MHy+ cations plays a very important role in this
formates and theoretical data reported for methylhydra- perovskite. However, significant broadening is also observed
zine.43,64,65 for the lattice modes, and this behavior can be attributed to
Thermal evolution of the Raman spectra, presented in softness of the structure and interactions of the organic layers
Figures 3, 4, and S6−S8, shows that when temperature with inorganic layers mediated by the HBs. Such dynamic
8568 DOI: 10.1021/acs.chemmater.9b03775
Chem. Mater. 2019, 31, 8563−8575
Chemistry of Materials Article

disorder was shown to play an important role in the


optoelectronic and photovoltaic properties of these perov-
skites.36,38,66 It was especially large for 3D perovskites that
were even regarded as phonon glass materials.17 Regarding
A2PbX4 2D analogues, only a few temperature-dependent
lattice dynamics studies were reported to date, i.e., non-
resonant Raman scattering spectra were analyzed for
(FC6H4C2H4NH3)2PbI4 in the 300−600 cm−1 range and
300−380 K as well as PEA2PbI4 and NBT2PbI4 in the 65−250
cm−1 range and 15−295 K.61,63 No information on temper-
ature dependence of Raman modes below 65 cm−1 was
reported, although a few such modes have been observed
recently at 15 K in PEA2PbI4 and NBT2PbI4 by high-resolution
resonant impulsive stimulated Raman spectroscopy.38 Our
results on MHy2PbI4 show the presence of a few modes below
40 cm−1. They also show pronounced broadening of both
intramolecular MHy+ modes and modes related to inorganic
layers. This broadening is much more pronounced for
MHy2PbI4 compared to PEA2PbI4 and NBT2PbI4. For
instance, the 119 cm−1 band shows increase of FWHM from Figure 5. (a) Dielectric permittivity, (b) dielectric loss, (c) real part,
9.8 cm−1 at 80 K to 25.8 cm−1 at 300 K, whereas for PEA2PbI4 and (d) imaginary part of electric modulus spectra as a function of
and NBT2PbI4, the FWHM values were less than 15 cm−1 for temperature in MHy2PbI4 single crystal. The representative curves are
the Pb−I stretching modes. Broadening of the τ(CH3) mode is plotted in frequency decades between 1 Hz and 1 MHz. The vertical
also larger (FWHM, 5.1 cm−1 at 80 K and 65.4 cm−1 at 300 K) lines correspond to the phase-transition temperatures obtained from
compared to the broadening of the torsional mode in the DSC measurements.
NBT2PbI4 observed at 239 cm−1 (FWHM < 25 cm−1). This
behavior proves that the nature of organic cation has a great
higher frequencies, confirming thermally induced activation
impact on the dynamic disorder in 2D lead halide perovskites
process. The observed dipolar process deviates from the
and that this disorder is larger in MHy2PbI4 than in PEA2PbI4
classical Debye behavior. All data were parameterized in the
and NBT2PbI4, and it resembles more that observed in 3D
vicinity of the peak maximum with the use of the single
perovskites. We argue that this difference reflects smaller
Havriliak−Negami function. To describe relaxation dynamics
separation of the inorganic layers and thus stronger coupling
of the observed dipolar processes, we compare the temper-
between the organic moieties and the inorganic perovskite
ature-dependent behavior of its dielectric relaxation times as a
layers mediated by the HBs.
function of 1000/T (see Figure S10c). This relaxation map
Electric Measurements. Since the DSC, X-ray diffraction,
depicts the dynamic properties of this compounds and the
and Raman data revealed the presence of phase transitions, we
characteristic electronic behavior of the investigated structure.
decided to check if these transitions are related with some
The thermal activation of their relaxation time τs spans
dielectric and electric conductivity anomalies. The temperature through almost 4 decades in frequency. In the investigated
dependence of the real ε′ and imaginary ε″ parts of the temperature range, the relaxation times τs exhibit linear
dielectric permittivity is given in Figure 5a,b. Both ε′ and ε″ tendencies versus the inverse temperature. Therefore, the
increase monotonically with increasing temperature and

ij E yz
relaxation times can be modeled using the Arrhenius relation

τ = τ0 expjjj a zzz
exhibit distinct frequency dependence starting from about

j kBT z
250 K, with clear anomalies near 295 and 320 K. The observed
k {
frequency dependence of ε′ and ε″ implies some relaxation
ordering process. Similar behavior was previously noted for a
number of APbI3 and [(CH3)4N]4Pb3Cl10 hybrid perovskites, where τ0, kB, and Ea are relaxation time at the high temperature
but we are not aware of any reports on dielectric dispersion in limit, Boltzmann constant, and activation energy, respectively.
A2PbX4 (X = Cl, Br, I) layered perovskites.67−70 To better Taking into account the X-ray diffraction data, the physical
visualize the relaxation dynamics and eliminate the contribu- nature and mechanism of the observed dielectric relaxation
tion of electrode polarization, the electric modulus M* = 1/ε* process in MHy2PbI4 can be attributed to the reorientational
representation was used.71 The temperature-dependent com- motion of MHy+ cations in the structure. No sizable dielectric
plex electric modulus spectra at various frequencies are shown relaxation in phase IV corroborates well with well-ordered
in Figure 5c,d. The value of M′ reaches almost a constant value MHy+ cations in this phase. With increasing temperature,
at lowest temperatures, but it approaches zero at high thermal motions of MHy+ cations become important and
temperatures. Each of the temperature dependence of M″ molecular dipole orientation can be achieved under the applied
exhibits a well-defined peak. The peak maximum shifts toward AC field, especially in phases II and I, in which MHy+ cations
higher temperatures with increasing frequency, indicating the exhibit rotational disorder. The obtained activation energies
presence of a structural relaxation. a r e s m a l l e r t h a n r e p o r t e d f o r o n e - d im e n s i o n a l
We also analyzed the frequency-dependent electric modulus (triethylpropylammonium)PbI3 for which Ea was 0.65 eV in
of MHy2PbI4 for several isotherms (Figure S10a,b). The the HT phase and 1.09 eV in the LT phase.69 Structural data
isothermal complex electric modulus spectra also reveal the allow also to understand temperature dependence of ε′. Due to
occurrence of dipolar relaxation process above 250 K. As the orientational disorder of MHy+ in phase I, the dipole
temperature increases, the modulus peak maxima move toward moment of the cation can easily reorient under AC field
8569 DOI: 10.1021/acs.chemmater.9b03775
Chem. Mater. 2019, 31, 8563−8575
Chemistry of Materials Article

leading to weakening of the field, which show the larger hexane, and 1,8-diaminooctane.53 It is also worth noting that
dielectric permittivity and a plateau above 320 K at high the conductivity of 2D MHy2PbI4 is an order of magnitude
frequencies. In phase IV (III), the cations are ordered (nearly smaller than that of 3D CH3NH3PbI3 (1.11 × 10−6 S m−1).73
ordered), leading to small dielectric permittivity in the AC Figure 6 shows that all phase transitions observed in DSC
field. However, the dielectric permittivity increases continu- measurements are also well visible in the temperature
ously on heating in phase II, and this behavior suggests gradual dependence of σDC. The estimated activation energy is 0.06
increase in reorientational motions of MHy+ cations. It is eV at the lowest temperatures, and it increases to 0.38 eV for
worth adding that the ε′ and ε″ show jumps at 295 K. This phase II. These values are of the same order of magnitude as
jump is 17.9 for ε′ at 1 Hz, and based on our X-ray diffraction those found for the migration of iodide ion vacancies in
data, it can be most likely attributed to abrupt change in the tilt CH3NH3PbI3 and ion migration in 2D BA2MA3Pb4I13, where
of PbI6 octahedra and orientation of the MHy+ cations. BA and MA denote n-butylammonium and methylammonium
The AC conductivity at different frequencies and temper- cation, respectively.74−76
atures was determined from the dielectric data using the To check the electrical anisotropy of MHy2PbI4, we have
Jonscher law72 performed additional measurements for the polycrystalline
sample (pellet). The obtained results, both for dielectric
σAC = σDC + Aωn + ε0ε″ω
permittivity and electric modulus, look very similar to those
where σDC corresponds to the frequency-independent direct observed for the single crystal (see Figures S11 and S12),
current conductivity, A is the temperature-dependent constant, suggesting small electrical anisotropy. The confirmation of this
n is the power exponent representing the degree of interaction assumption is the relaxation map for the pellet sample, which is
between mobile ions and the surroundings, and ω = 2πf [f very close to that obtained for the single crystal with
(Hz)] is the angular frequency. The third term defines comparable value of activation energies (Figure S13).
dielectric losses related to the dielectric relaxation process, Optical Properties. Diffuse reflectance spectrum, pre-
where ε0 is the vacuum permittivity. sented in Figure S14, shows a band at 561 nm that, in analogy
The frequency dependence of the AC conductivity in the with other 2D perovskites, can be attributed to excitonic
log−log scale is depicted in Figure 6a. It is known that the absorption.24 The excitonic absorption is among the most red-
shifted in the family of A2PbI4(100) layered perovskites. For
instance, other (100) perovskites with short separation of
inorganic layers exhibit the excitonic peaks at 553,21 544,24 and
53652 nm for CsGAPbI4, PyrEAPbI4, and EA2PbI4, respec-
tively, while for NBT2PbI4 and PEA2PbI4 with large separation
of inorganic layers, these peaks were observed at 513 and 517
nm, respectively (Table S6).58,62
The room-temperature diffuse reflectance spectrum was
used to determine the energy band gap (Eg) of MHy2PbI4 with
the Kubelka−Munk relation77

(1 − R )2
F (R ) =
2R
Figure 6. Variation of (a) AC conductivity with frequency in the where R denotes reflectance. The graphical analysis of the
temperature range 160−340 K and (b) DC conductivity vs energy band gap from UV−vis reflectance transformed to
temperature for MHy2PbI4. The dashed line depicts the phase- [F(R)hν]2 and plotted versus hν, which is presented in the
transition temperatures obtained from the DSC measurements. inset to Figure S14. The value of Eg in MHy2PbI4 is estimated
Activation energies in the selected temperature ranges are Ea1 = to be 563 nm (2.20 eV). For CsGAPbI4, GAMAPbI4,
0.23 eV; Ea2 = 0.38 eV; Ea3 = 0.26 eV; and Ea4 = 0.06 eV. PyrEAPbI4, 3AMPPbI4, 4AMPPbI4, and PEA2PbI4, the Eg
values are 2.35,21 2.27,20 2.48,24 2.23,27 2.38,27 and 2.7 eV,58
mechanism of conductivity can be explained taking into respectively (Table S6). As can be noted, MHy2PbI4 shows the
account the temperature and frequency dependence of n. The smallest energy of Eg among all A2PbI4 layered perovskites.
hopping mechanism of conductivity is represented by n close Interestingly, the band gap of MHy2PbI4 is smaller than the
to 1.72 The variation of the exponent n as a function of band gap of CsGAPbI4 analogue, in spite of larger separation
temperature is presented in the inset of Figure 6a. In the between inorganic layers. According to former reports, smaller
investigated temperature range, the n is equal to 1 at the lowest distortion of the structure should lead to red shift of the band
temperatures and decreases to almost zero in the HT phase, gap.78 Inspection of the crystal structures shows that Pb−I−Pb
which proves that the conductivity mechanism in MHy2PbI4 angles between the octahedral units are larger for MHy2PbI4
strongly depends on temperature. (170.3 and 180° at 305 K, Table 1) than CsGAPbI4 (149.74
The plot of direct current conductivity σDC versus and 165.19°).21 Thus, the red shift of Eg for MHy2PbI4
temperature is presented in Figure 6b. MHy2PbI4 shows compared to CsGAPbI4 is consistent with more ideal Pb−I−
quite a low values of σDC at low temperatures (1.01 × 10−11 S Pb bond angles.
m−1 at 160 K), but the conductivity increases rapidly with PL spectra of MHy2PbI4 consist of two bands (Figures 7a
increasing temperature and reaches 1.02 × 10−7 S m−1 at 298 and S15). The P1 band, observed at 561 nm at RT, overlaps
K and 3.5 × 10−6 S m−1 at 340 K. A similar RT conductivity of with the excitonic absorption (Figure 7b) and exhibits weak
1.3 × 10−7 S m−1 was observed for other layered A2PbI4 temperature dependence (Figure 8). Its FWHM is only 21
compounds with protonated 1,4-diaminobutane, 1,6-diamino- meV at 80 K (Figure 8). Such minimally Stokes shifted and
8570 DOI: 10.1021/acs.chemmater.9b03775
Chem. Mater. 2019, 31, 8563−8575
Chemistry of Materials Article

permanent lattice defects but is intrinsic feature of the studied


sample.
Figures 7a, 9, and S15 show that intensity of the P1 and P2
bands decreases quickly on heating the sample, indicating that

Figure 7. (a) Temperature-dependent PL spectra measured every 10 Figure 9. (a) Temperature-dependent integrated intensities for P1
K from 80 to 300 K, (b) normalized absorption and emission spectra, (FE) and P2 (BE) bands and (b) CIE coordinates of MHy2PbI4 at
(c) configuration coordinate diagram for FE and BE states, and (d) different temperatures.
temperature dependence of bands’ intensity (contour map).
the thermal activation of nonradiative processes is activated at
higher temperatures. A closer inspection of the PL spectra
shows that PL intensity drops by nearly 92% when heating
from 80 to 220 K and to 0.8% at 290 K. Former studies of
EA2PbI4 and PEA2PbI4 showed a drop of about 30% at RT for
the latter compound and to almost zero near 220 K for
EA2PbI4.25 This behavior of MHy2PbI4 is intermediate
between that observed for PEA2PbI4 with strong dielectric
confinement and EA2PbI4 with strongly reduced dielectric
confinement. We can, therefore, suppose that the excitonic
activation energy of MHy2PbI4 is higher than the 13 meV value
found for EA2PbI4 and smaller than 250 meV found for
Figure 8. Temperature dependence of band center positions and PEA2PbI4.25 To check this assumption, FE activation energy
FWHM for FE and BE bands. has been estimated using temperature dependence of P1 band
intensity and assuming validity of the following equation
narrow PL is characteristic of FE recombination.13,58,62,79 The I0
second broader band P2 is observed at 610 nm at RT, and this I (T ) =
band exhibits blue shift on cooling (to 573 nm at 80 K). Its 1 + e−Ea / kBT
separation from the P1 band is only 12.3 nm (47 meV) at 80 where I0, Ea, and kB correspond to the emission intensity at low
K, but it increases to 49 nm (178 meV) at RT. Additional red- temperature, activation energy, and Boltzmann constant. From
shifted bands have been observed for many A 2 PbI 4 the fit of the low-temperature experimental data, approximate
compounds, and they were usually attributed to bound activation energy of 59.2 meV was extracted for MHy2PbI4
excitons,28,34 phonon replicas of the FE,33 and out-of-plane- (Figure S17). As expected, this value is among the smallest
oriented magnetic dipole transition.80 Bound excitons were found for A2PbI4 perovskites. It is also smaller than 96.5 meV
found red-shifted at RT by around 32 nm for cystamine lead found for PyrEAPbI4, which showed a power conversion
iodide and 30 nm for PEA2PbI4.28,34 Furthermore, study of efficiency of 1.43%.24 This behavior is consistent with short
PEA2PbI4 revealed significant blue shift of the BE PL on separation between the inorganic layers (shorter than for
cooling.34 The band assigned to phonon replica in trans-2,5- PyrEAPbI4 and EA2PbI4) and high dielectric constant of
dimethylpiperazinium lead iodide was red-shifted by 33 nm MHy+. It is worth adding that although Ea may not correspond
and the separation from the FE band changed weakly on exactly to the exciton binding energy, its small value suggests
cooling.33 In the case of butylammonium lead iodide, the band that the exciton binding energy is also small. Indeed, according
red-shifted by 20 nm was attributed to the out-of-plane- to Cheng et al., for A2PbI4 compounds with organic cation
oriented magnetic dipole transition.80 The strong decrease of possessing dielectric constant of 22 (like methylhydrazine), the
the separation between P1 and P2 bands in MHy2PbI4 is not exciton binding energy should be about 30 meV.25 Thus,
consistent with phonon replica, but such behavior as well as similarly to EA2PbI4, MHy2PbI4 also has significantly reduced
significantly larger width of the P2 band could indicate that this dielectric confinement, is more like a 3D perovskite in
emission comes from BE states. PL quantum yield (PLQY) of character, and might be promising material for photovoltaic
MHy2PbI4 reaches 0.8% at RT. Such a low value of PLQY at applications.
RT is typical for 2D iodides.33,34 We also studied PL PL of MHy2PbI4 is yellowish pink at RT with CIE
dependence on excitation intensity. The PL intensity of FE coordinate (0.43, 038) (Figure 9b). However, due to
and BE states increases linearly with excitation power density significant shift of P2 band on cooling and different
from 4.97 to 92.2 mW cm−2 at 80 K (Figure S16). This result temperature dependence of its intensity, chromacity exhibits
indicates that the broader emission is not associated with any strong temperature dependence. That is, the color of the PL
8571 DOI: 10.1021/acs.chemmater.9b03775
Chem. Mater. 2019, 31, 8563−8575
Chemistry of Materials Article

Figure 10. (a−c) Color maps present intensity of PL and lifetimes of MHy2PbI4 at 80, 193, and 233 K under excitation at 400 nm using
femtoseconds laser. (d−f) PL emission spectra at selected temperatures. (g−i) Emission decay curves for FE and BE bands at given temperatures.

changes with decreasing temperature to yellow-green at 80 K


(0.45, 0.54) (Figure 9b). MHy2PbI4 exhibits, therefore,
■ CONCLUSIONS
We report the synthesis of 2D lead iodide with single-
thermochromic properties. It is worth adding that significant octahedra slabs of MHy2PbI4 composition. This compound has
changes of the temperature-dependent emission intensity the smallest organic cation among all reported A2PbI4
indicate that this perovskite has high potential for noncontact compounds, and we show that this feature leads to unusual
temperature sensing. The potential of the metal-organic properties of this material. First, interoctahedral distortion of
frameworks with perovskite-type architecture for noncontact the inorganic lattice in MHy2PbI4 is among the smallest among
thermometry has been recently reported by Ptak et al. for A2PbI4 perovskites. Due to this feature and high dielectric
formate-based materials.81 constant of organic layers, MHy2PbI4 shows exceptionally
To further understand the origin of PL bands, we have also small band gap and is expected to exhibit strongly reduced
performed femtosecond-time-resolved measurements at vari- dielectric confinement. Second, Raman studies have revealed
ous temperatures (from 80 to 233 K) under 400 nm excitation that this compound shows significantly higher dynamic
line generated by femtoseconds laser. Two bands were disorder than 2D A2PbI4 analogues with larger organic cations.
observed (Figures 10 and S18). It can be seen that at 80 K, Similarity of the dynamic behavior to 3D analogues can be
the emission at 560 nm decays very quickly (0.37 ns) while attributed to small separation between the lead iodide layers
that at 573 nm relaxes more slowly (0.46 ns) (Figures 10g−i, and strong HBs mediated coupling between inorganic and
S19, and S20). Former study of cystamine lead iodide showed organic layers. Third, MHy2PbI4 shows two narrower bands in
that the decay time was slightly longer for BE (0.81 ns) than the 560−610 nm range that exhibit strong change in intensity
FE emission (0.51 ns).28 On the other hand, the decay times of in the 80−300 K range. Furthermore, the red-shifted band
the FE and red-shifted band of trans-2,5-dimethylpiperazinium exhibits pronounced blue shift on cooling. Thus, MHy2PbI4
lead iodide, assigned to phonon replica, were very similar (0.24 shows thermochromic behavior. Fourth, MHy2PbI4 shows
nm).33 The same decay times were also reported for the FE switchable dielectric behavior related to an order−disorder
and red-shifted band in butylammonium lead iodide attributed phase transition at 298 K. It also exhibits ionic conductivity
to out-of-plane-oriented magnetic dipole transition (0.296 that is however an order of magnitude smaller than the
ns).80 Thus, decay times of MHy2PbI4 are consistent with conductivity of famous MAPbI3 perovskite.


assignment of the P2 band to weakly bound exciton. With
increasing temperature, the decay times of both emissions in ASSOCIATED CONTENT
MHy2PbI4 increase to 0.70 ns, i.e., they are significantly longer
than the lifetimes of 3AMPPbI4, 4AMPPbI4, and PEA2PbI4 *
S Supporting Information

(0.08, 0.04, and 0.17 ns, respectively).27,82 This indicates The Supporting Information is available free of charge on the
slower carrier combination in MHy2PbI4, which is beneficial ACS Publications website at DOI: 10.1021/acs.chemma-
for photovoltaic applications. It is worth adding that the ter.9b03775.
lifetimes of MHy2PbI4 are comparable to the lifetime of Crystal data; selected geometrical and hydrogen-bond
PyrEAPbI4(100) perovskite (0.41 ns) that has been recently parameters for the studied compound at different
shown to be promising solar cell material with a power temperatures; structural and optical parameters for
conversion efficiency of 1.43%.24 selected A2PbI4(100) perovskites; and Raman wave-
8572 DOI: 10.1021/acs.chemmater.9b03775
Chem. Mater. 2019, 31, 8563−8575
Chemistry of Materials Article

numbers together with the proposed assignments Perovskite Nanosheets with Bright, Tunable Photoluminescence and
(Tables S1−S7); powder XRD pattern; DSC traces; High Stability. Angew. Chem., Int. Ed. 2017, 56, 4252−4255.
TGA plot; reconstruction of hk2 reciprocal layer and 2D (12) Yaffe, O.; Chernikov, A.; Norman, Z. M.; Zhong, Y.;
profiles; hydrogen-bond interactions; Raman spectra; Velauthapillai, A.; Van der Zande, A.; Owe, J. S.; Heinz, T. F.
Excitons in Ultrathin Organic-Inorganic Perovskite Crystals. Phys. Rev.
temperature dependence of Raman wavenumbers;
B 2015, 92, No. 045414.
FWHM values and dielectric permittivity; map of (13) Gan, L.; Li, J.; Fang, Z.; He, H.; Ye, Z. Effect of Organic Cation
dipolar relaxation process; diffuse reflectance spectrum; Length on Exciton Recombination in Two-Dimensional Layered Lead
PL spectra; power dependence of PL; temperature- Iodide Hybrid Perovskite Crystals. J. Phys. Chem. Lett. 2017, 8, 5177−
dependent decay times (Figures S1−S20) (PDF) 5183.
CIF files for structures at 100 K (CIF) (14) Saouma, F. O.; Stoumpos, C. C.; Wong, J.; Kanatzidis, M. Z.;
CIF files for structures at 280 K (CIF) Jang, J. I. Selective Enhancement of Optical Nonlinearity in Two-
CIF files for structures at 305 K (CIF) Dimensional Organic-Inorganic Lead Iodide Perovskites. Nat.
Commun. 2017, 8, No. 742.
CIF files for structures at 330 K (CIF) (15) Chong, W. K.; Thirumal, K.; Giovanni, D.; Goh, T. W.; Liu, X.;

■ AUTHOR INFORMATION
Corresponding Author
Mathews, N.; Mhaisalkar, S. G.; Sum, T. C. Dominant Factors
Limiting Optical Gain in Layered Two-Dimensional Halide Perov-
skites Thin Films. Phys. Chem. Chem. Phys. 2016, 18, 14701−14708.
(16) Ishihara, T.; Takahashi, J.; Goto, T. Exciton State in Two-
*E-mail: m.maczka@int.pan.wroc.pl. Phone: +48-713954161.
Dimensional Perovskite Semiconductor (C10H21NH3)2PbI4. Solid
Fax: +48-713441029. State Commun. 1989, 69, 933−936.
ORCID (17) Miyata, K.; Atallah, T. L.; Zhu, X.-Y. Lead Halide Perovskites:
Mirosław Mączka: 0000-0003-2978-1093 Crystal-Liquid Duality, Phonon Glass Electron Crystals, and Large
Maciej Ptak: 0000-0002-4639-2367 Polaron Formation. Sci. Adv. 2017, 3, No. e1701469.
(18) Hong, X.; Ishihara, T.; Nurmikko, A. Dielectric Confinement
Adam Sieradzki: 0000-0003-4136-5754
Effect on Excitons in PbI4-Based Layered Semiconductors. Phys. Rev.
Author Contributions B 1992, 45, 6961−6964.
The manuscript was written through contributions of all (19) Pedesseau, L.; Sapori, D.; Traore, B.; Robles, R.; Fang, H.-H.;
authors. All authors have given approval to the final version of Loi, M. A.; Tsai, H.; Nie, W.; Blancon, J.-C.; Neukirch, A.; Tretiak, S.;
the manuscript. Mohite, A. D.; Katan, C.; Even, J.; Kapenekian, M. Advances and
Notes Promises of Layered Halide Hybrid Perovskite Semiconductors. ACS
The authors declare no competing financial interest. Nano 2016, 10, 9776−9786.


(20) Soe, C. M. M.; Stoumpos, C. C.; Kapenekian, M.; Traore, B.;
Tsai, H.; Nie, W.; Wang, B.; Katan, C.; Seshadri, R.; Mohite, A. S.;
REFERENCES Even, J.; Marks, T. J.; Kanatzidis, M. G. New Type of 2D Perovskites
(1) Hoefler, S. F.; Trimmel, G.; Rath, T. Progress on Lead-Free with Alternating Cations in the Interlayer Space, (C(NH2)3)-
Metal Halide Perovskites for Photovoltaic Applications: a Review. (CH3NH3)nPbnI3n+1: Structure, Properties, and Photovoltaic Per-
Monatsh. Chem. 2017, 148, 795−826. formance. J. Am. Chem. Soc. 2017, 139, 16297−16309.
(2) Wang, F.; Cao, Y.; Chen, C.; Chen, Q.; Wu, X.; Li, X.; Qin, T.; (21) Nazarenko, O.; Kotyrba, M. R.; Wörle, M.; Cuervo-Reyes, E.;
Huang, W. Materials Toward the Upscaling of Perovskite Solar Cells: Yakunin, S.; Kovalenko, M. V. Luminescent and Photoconductive
Progress, Challenges, and Strategies. Adv. Funct. Mater. 2018, Layered Lead Halide Perovskite Compounds Comprising Mixtures of
No. 1803753. Cesium and Guanidinium Cations. Inorg. Chem. 2017, 56, 11552−
(3) Zhou, C.; Lin, H.; Lee, S.; Chaaban, M.; Ma, B. Organic- 11564.
Inorganic Metal Halide Hybrids Beyond Perovskites. Mater. Res. Lett. (22) Daub, M.; Haber, C.; Hillebrecht, H. Synthesis, Crystal
2018, 6, 552−569. Structures, Optical Properties, and Phase Transitions of the Layered
(4) Brochard-Garnier, S.; Paris, M.; Genois, R.; Han, Q.; Liu, Y.;
Guanidinium-Based Hybrid Perovskites [(C(NH2)3]2MI4; M=Sn, Pb.
Massuyeau, F.; Gautier, R. Screening Approach for the Discovery of
Eur. J. Inorg. Chem. 2017, 1120−1126.
New Hybrid Perovskites with Efficient Photoemission. Adv. Funct.
(23) Nazarenko, O.; Kotyrba, M. R.; Yakunin, S.; Aebli, M.; Raino,
Mater. 2018, No. 1806728.
G.; Benin, B. M.; Wörle, M.; Kovalenko, M. V. Guanidinium-
(5) Zhao, X.; Ng, J. D. A.; Friend, R. H.; Tan, Z.-K. Opportunities
and Challenges in Perovskite Light-Emitting Devices. ACS Photonics Formamidinium Lead Iodide: A Layered Perovskite-Related Com-
2018, 5, 3866−3875. pound with Red Luminescence at Room Temperature. J. Am. Chem.
(6) Cai, P.; Wang, X.; Seo, H. J.; Yan, X. Bluish-White-Light- Soc. 2018, 140, 3850−3853.
Emitting Diodes Based on Two-Dimensional Lead Halide Perovskite (24) Febriansyah, B.; Koh, T. M.; Lekina, Y.; Jamaludin, N. F.;
(C6H5C2H4NH3)2PbCl2Br2. Appl. Phys. Lett. 2018, 112, No. 153901. Bruno, A.; Ganguly, R.; Zhen, Z. X.; Mhaisalkar, S. G.; England, J.
(7) Li, W.; Wang, Z.; Deschler, F.; Gao, S.; Friend, R. H.; Cheetham, Improved Photovoltaic Efficiency and Amplified Photocurrent
A. K. Chemically Diverse and Multifunctional Hybrid Organic- Generation in Mesoporous n=1 Two-dimensional Lead-Iodide
Inorganic Perovskites. Nat. Rev. Mater. 2017, 2, No. 16099. Perovskite Solar Cells. Chem. Mater. 2019, 31, 890−898.
(8) Chouhan, L.; Ghimire, S.; Biju, V. Blinking Beats Bleaching: The (25) Cheng, B.; Li, T.-Y.; Maity, P.; Wei, P.-C.; Nordlund, D.; Ho,
Control of Superoxide Generation by Photo-ionized Perovskite K.-T.; Lien, D.-H.; lin, C.-H.; Liang, R.-Z.; Miao, X.; Ajia, I. A.; Yin, J.;
Nanocrystals. Angew. Chem., Int. Ed. 2019, 58, 4875−4879. Sokaras, D.; Javey, A.; Roqan, I. S.; Mohammed, O. F.; He, J. H.
(9) Huo, C.; Cai, B.; Yuan, Z.; Ma, B.; Zheng, H. Two-Dimensional Extremely Reduced Dielectric Confinement in Two-Dimensional
Metal Halide Perovskites: Theory, Synthesis and Optoelectronics. Hybrid Perovskites with Large Polar Organics. Commun. Phys. 2018,
Small Methods 2017, 1, No. 1600018. 1, No. 80.
(10) Chao, L.; Wang, Z.; Xia, Y.; Chen, Y.; Huang, W. Recent (26) Lekina, Y.; Shen, Z. X. Excitonic States and Structural Stability
Progress on Low Dimensional Perovskite Solar Cells. J. Energy Chem. in Two-Dimensional Hybrid Organic-Inorganic Perovskites. J. Sci.:
2018, 27, 1091−1100. Adv. Mater. Devices 2019, 4, 189−200.
(11) Yang, S.; Niu, W.; Wang, A.-L.; Fan, Z.; Chen, B.; Tan, C.; Lu, (27) Mao, L.; Ke, W.; Pedesseau, L.; Wu, Y.; Katan, C.; Even, J.;
Q.; Zhang, H. Ultrathin Two-Dimensional Organic-Inorganic Hybrid Wasilewski, M. R.; Stoumpos, C. C.; Kanatzidis, M. G. Hybrid Dion-

8573 DOI: 10.1021/acs.chemmater.9b03775


Chem. Mater. 2019, 31, 8563−8575
Chemistry of Materials Article

Jacobson Lead Iodide Perovskites. J. Am. Chem. Soc. 2018, 140, Synthesis, Structural and Optical Characterization of APbX3
3775−3783. (A=methylammonium, dimethylammaonium, trimethylammonium;
(28) Krishnamurthy, S.; Kour, P.; Katre, A.; Gosavi, S.; Chakraborty, X=I, Br, Cl) Hybrid Organic-Inorganic Materials. J. Solid State
S.; Ogale, S. Cystamine Configured Lead Halide Based 2D Hybrid Chem. 2016, 240, 55−60.
Molecular Crystals: Synthesis and Photoluminescence Systematics. (46) Seth, C.; Khushalani, D. Non-Perovskite Hybrid Material,
APL Mater. 2018, 6, No. 114204. Imidazolium Lead Iodide, with Enhanced Stability. ChemNanoMat
(29) Zhou, C.; Lin, H.; Lee, S.; Chaaban, M.; Ma, B. Organic- 2019, 5, 85−91.
Inorganic Metal Halide Hybrids Beyond Perovskites. Mater. Res. Lett. (47) Simenas, M.; Balčiu̅nas, S.; Trzebiatowska, M.; Ptak, M.;
2018, 6, 552−569. Mączka, M.; Völkel, G.; Pöppl, A.; Banys, J. Electron Paramagnetic
(30) Dohner, E. R.; Jaffe, A.; Bradshow, L. R.; Karunadasa, H. I. Resonance and Electric Characterization of a [CH3NH2NH2][Zn-
Intrinsic White-Light Emission from Layered Hybrid Perovskites. J. (HCOO)3] Perovskite Metal Formate Framework. J. Mater. Chem. C
Am. Chem. Soc. 2014, 136, 13154−13157. 2017, 5, 4526−4536.
(31) Yangui, A.; Garrot, D.; Lauret, J. S.; Lusson, A.; Bouchez, G.; (48) Sieradzki, A.; Maczka, M.; Simenas, M.; Zareba, J. K.; Gagor,
Deleporte, E.; Pillet, S.; Bendeif, E. E.; Castro, M.; Triki, S.; Abid, Y.; A.; Balciunas, S.; Kinka, M.; Ciupa, A.; Nyk, M.; Samulionis, V.;
Boukheddaden, K. Optical Investigation of Broadband White-Light Banys, J.; Paluch, M.; Pawlus, S. On the Origin of Ferroelectric
Emission in Self-Assembled Organic-Inorganic Perovskite Structural Phases in Perovskite-like Metal-Organic Formate. J. Mater.
(C6H11NH3)2PbBr4. J. Phys. Chem. C 2015, 119, 23638−23647. Chem. C 2018, 6, 9420−9429.
(32) Mao, L.; Wu, Y.; Stoumpos, C. C.; Wasilewski, M. R.; (49) Delil, A. A. M. Characterisation of the Dielectric Properties of
Kanatzidis, M. R. White-Light Emission and Structural Distortion in the Propellants MON & MMH. AIP Conf. Proc. 1998, 420, 258.
New Corrugated Two-Dimensional Lead Bromide Perovskites. J. Am. (50) Samantaray, R.; Clark, R. J.; Choi, E. S.; Dalal, N. S. Elucidating
Chem. Soc. 2017, 139, 5210−5215. the Mechanism of Multiferroicity in (NH4)3Cr(O2)4 and Its Tailoring
(33) Gautier, R.; Paris, M.; Massuyeau, F. Massuyeau, Exciton Self- by Alkali Metal Substitution. J. Am. Chem. Soc. 2012, 134, 15953−
Trapping in Hybrid Lead Halides: Role of Halogen. J. Am. Chem. Soc. 15962.
2019, 141, 12619−12623. (51) Mitzi, D. B. Synthesis, Crystal Structure, and Optical and
(34) Yu, J.; Kong, J.; Hao, W.; Guo, X.; He, H.; Leow, W. R.; Liu, Z.; Thermal Properties (C4H9NH3)2MI4 (M=Ge, Sn, Pb). Chem. Mater.
Cai, P.; Qian, G.; Li, S.; Chen, X.; Chen, X. Broadband Extrinsic Self- 1996, 8, 791−800.
Trapped Exciton Emission in Sn-doped 2D Lead-Halide Perovskites. (52) Mercier, N.; Poiroux, S.; Riou, A.; Batail, P. Unique Hydrogen
Adv. Mater. 2018, 31, No. 1806385. Bonding Correlating with a Reduced Band Gap and Phase Transitions
(35) Smith, M. D.; Karunadasa, H. I. White-Light Emission from in the Hybrid Perovskites (HO(CH2)2NH3)PbX4 (X=I, Br). Inorg.
Layered halide Perovskites. Acc. Chem. Res. 2018, 51, 619−627.
Chem. 2004, 43, 8361−8366.
(36) Thouin, F.; Neutzner, S.; Cortecchia, D.; Dragomir, V. A.; Soci, (53) Safdari, M.; Svensson, P. H.; Hoang, M. T.; Oh, I.; Kloo, L.;
C.; Salim, T.; Lam, Y. M.; Leonelli, R.; Petrozza, A.; Kandada, A. R.
Gardner, J. M. Layered 2D Alkyldiammonium Lead Iodide Perov-
S.; Silva, C. Stable Biexcitons in Two-Dimensional Metal Halide
skites: Synthesis, Characterization, and Use in Solar Cells. J. Mater.
Perovskites with Strong Dynamic Lattice Disorder. Phys. Rev. Mater.
Chem. A 2016, 4, 15638−15646.
2018, 2, No. 034001.
(54) Katan, C.; Mercier, N.; Even, J. Quantum and Dielectric
(37) Neutzner, S.; Thouin, F.; Cortecchia, D.; Petrozza, A.; Silva, C.;
Confinement Effects in Lower-Dimensional Hybrid Perovskite
Kandada, A. R. S. Exciton-Polaron Spectral Structures in Two-
Semiconductors. Chem. Rev. 2019, 119, 3140−3192.
Dimensional Hybrid Lead-Halide Perovskites. Phys. Rev. Mater. 2018,
(55) Szklarz, P.; Gagor, A.; Jakubas, R.; Zieliński, P.; Piecha-Bisiorek,
2, No. 064605.
(38) Thouin, F.; Valverde-Chavez, D. A.; Quarti, C.; Cortecchia, D.; A.; Cichos, J.; Karbowiak, M.; Bator, G.; Ciżman, A. Lead-free Hybrid
Bargigia, I.; Beljonne, D.; Petrozza, A.; Silva, C.; Kandada, A. R. S. Ferroelectric Material Based on Formamidine: [NH2CHNH2]3Bi2I9. J.
Phonon Coherences Reveal the Polaronic Character of Excitons in Mater. Chem. C 2019, 7, 3003−3014.
Two-Dimensional Lead Halide Perovskites. Nat. Mater. 2019, 18, (56) Fleet, M. E. Distortion Parameters for Coordination Polyhedra.
349−356. Mineral. Mag. 1976, 40, 531−533.
(39) Ciupa, A.; Maczka, M.; Gągor, A.; Sieradzki, A.; Trzmiel, J.; (57) Billing, D. G.; Lemmerer, A. Synthesis, Characterization and
Pikul, A.; Ptak, M. Temperature-dependent Studies of [(CH3)2NH2]- Phase Transitions in the Inorganic-Organic Layered Perovskite-type
[FeIIIMII(HCOO)6] Frameworks (MII=Fe and Mg): Structural, Hybrids [(CnH2n+1NH3)2PbI4], n=4, 5 and 6. Acta Crystallogr., Sect. B:
Magnetic, Dielectric and Phonon Properties. Dalton Trans. 2015, Struct. Sci. 2007, 63, 735−747.
44, 8846−8854. (58) Du, K.-z.; Tu, Q.; Zhang, X.; Han, Q.; Liu, J.; Zauscher, S.;
(40) Mączka, M.; Pietraszko, A.; Macalik, B.; Hermanowicz, K. Mitzi, D. B. Two-Dimensional Lead(II) Halide-Based Hybrid
Structure, Phonon Properties, and Order-Disorder Transition in the Perovskites Templated by Acene Alkylamines: Crystal Structures,
Metal Formate Framework of [NH4][Mg(HCOO)3]. Inorg. Chem. Optical Properties and Piezoelectricity. Inorg. Chem. 2017, 56, 9291−
2014, 53, 787−794. 9302.
(41) Weller, M. T.; Weber, O. J.; Frost, J. M.; Walsh, A. Cubic (59) Barman, S.; Venkataraman, N. V.; Vasudevan, S.; Seshadri, S.
Perovskite Structure of Black Formamidinium Lead Iodide, α- Phase transitions in the Anchored Bilayers of Long-Chain
[HC(NH2)2]PbI3, at 298 K. J. Phys. Chem. Lett. 2015, 6, 3209−3212. Alkylammonium Lead Iodides (CnH2n+1NH3)2PbI4. J. Phys. Chem. B
(42) Szafrański, M.; Katrusiak, A. Phase Transitions in the Layered 2003, 107, 1875−1883.
Structure of Diguanidinium Tetraiodoplumbate. Phys. Rev. B 2000, (60) Yangui, A.; Pillet, S.; Garrot, D.; Triki, S.; Abid, Y.;
61, 1026−1035. Boukheddaden, K. Evidence and Detailed Study of a Second Order
(43) Mączka, M.; Gągor, A.; Ptak, M.; Paraguassu, W.; da Silva, T. Phase Transition in the (C6H11NH3)2PbI4 Organic-Inorganic
A.; Sieradzki, A.; Pikul, A. Phase Transitions and Coexistence of Hybrid Material. J. Appl. Phys. 2015, 117, No. 115503.
Magnetic and Electric Orders in the Methylhydrazinium Metal (61) Koubba, M.; Dammak, T.; Garrot, D.; Castro, M.; Codjovi, E.;
Formates. Chem. Mater. 2017, 29, 2264−2275. Mlayah, A.; Abid, Y.; Boukheddaden, K. Thermally-Induced First-
(44) Im, J. H.; Chung, J.; Kim, S.-J.; Park, N.-G. Synthesis, Structure, Order Phase Transition in the (FC6H4C2H4NH3)2[PbI4]. J. Appl.
and Photovoltaic Property of a Nanocrystalline 2H Perovskite-type Phys. 2012, 111, No. 053521.
Novel Sensitizer (CH3CH2NH3)PbI3. Nanoscale Res. Lett. 2012, 7, (62) Cortecchia, D.; Neutzner, S.; Kandada, A. R. S.; Mosconi, E.;
No. 353. Meggiolaro, D.; De Angelis, F.; Soci, C.; Petrozza, A. Broadband
(45) Mancini, A.; Quadrelli, P.; Amoroso, G.; Milanese, C.; Emission in Two-Dimensional Hybrid Perovskites: The Role of
Boiocchi, M.; Sironi, A.; Patrini, M.; Guizzetti, G.; Malavasi, L. Structural Deformation. J. Am. Chem. Soc. 2017, 139, 39−42.

8574 DOI: 10.1021/acs.chemmater.9b03775


Chem. Mater. 2019, 31, 8563−8575
Chemistry of Materials Article

(63) Dragomir, V. A.; Neutzner, S.; Quarti, C.; Cortecchia, D.; Halide Perovskites. 2019, arXiv:1901.05136. arXiv.org e-Print archive.
Petrozza, A.; Roorda, S.; Beljonne, D.; Leonelli, R.; Kandada, A. R. S.; https://export.arxiv.org/pdf/1901.05136 (accessed April 15, 2019).
Silva, C. Lattice Vibrations and Dynamic Disorder in Two- (81) Ptak, M.; Dziuk, B.; Stefanska, D.; Hermanowicz, K. The
Dim ens ional H ybri d Le ad Ha li de P erov skit es . 2 018 , structural, Phonon and Optical Properties of [CH 3 NH 3 ]-
arXiv:1812.05255. arXiv.org e-Print archive. https://arxiv.org/abs/ M0.5CrxAl0.5‑x(HCOO)3 (M=Na, K; x=0. 0.025, 0.5) Metal-Organic
1812.05255. Framework Perovskites for Luminescence Thermometry. Phys. Chem.
(64) Durig, J. R.; Harris, W. C.; Wertz, D. W. Infrared and Raman Chem. Phys. 2019, 21, 7965−7972.
Spectra of Substituted Hydrazines. I. Methylhydrazine. J. Chem. Phys. (82) Peng, W.; Yin, J.; Ho, K.-T.; Ouellette, O.; De Bastiani, M.;
1969, 50, 1449−1461. Murali, B.; El Tall, O.; Shen, C.; Miao, X.; Pan, J.; Alarousu, E.; He, J.-
(65) Durig, J. R.; Gounev, T. K.; Zheng, C.; Choulakian, A.; Verma, H.; Ooi, B. S.; Mohammed, O. F.; Sargent, E.; Bakr, O. M. Ultralow
V. N. Conformational Stability from Temperature-Dependent FT-IR Self-Doping in Two-dimensional Hybrid Perovskite Single Crystals.
Spectra of Liquid Xenon Solutions, ab Initio Calculations, and r0 Nano Lett. 2017, 17, 4759−4767.
Parameters for Methylhydrazine. J. Phys. Chem. A 2002, 106, 3395−
3402.
(66) Herz, L. M. How lattice Dynamics Moderate Electronic
Properties of Metal-Halide Perovskites. J. Phys. Chem. Lett. 2018, 9,
6853−6863.
(67) Fabini, D. H.; Hogan, T.; Evans, H. A.; Stoumpos, C. C.;
Kanatzidis, M. G.; Seshadri, S. Dielectric and Thermodynamic
Signatures of Low-Temperature Glassy Dynamics in the Hybrid
Perovskites CH3NH3PbI3 and HC(NH2)2PbI3. J. Phys. Chem. Lett.
2016, 7, 376−381.
(68) Xue, C.; Yao, Z.-Y.; Zhang, J.; Liu, W.-L.; Liu, J.-L.; Ren, X.-M.
Extra Thermo- and Water-Stable One-Dimensional Organic-Inorganic
Hybrid Perovskite [N-methyldabconium]PbI3 Showing Switchable
Dielectric Behaviour, Conductivity and Bright Yellow-Green
Emission. Chem. Commun. 2018, 54, 4321−4324.
(69) Wang, M.-J.; Chen, X.-R.; Tong, Y.-B.; Yuan, G.-J.; Ren, X.-M.;
Liu, J.-L. Phase Transitions, Dielectrics, Single-Ion Conductance and
Thermochromic Luminescence of a Inorganic-Organic Hybrid of
[Triethylpropylammonium][PbI3]. Inorg. Chem. 2017, 56, 9525−
9534.
(70) Xue, C.; Wang, S.; Liu, W.-L.; Ren, X.-M. Two-Step Structure
Phase Transition, Dielectric Anomalies, and Thermochromic
Luminescence Behaviour in a Direct Band Gap 2D Corrugated
Layer Lead Chloride Hybrid of [(CH3)4N]4Pb3Cl10. Chem. − Eur. J.
2019, 25, 5280−5287.
(71) Molak, A.; Paluch, M.; Pawlus, S.; Klimontko, J.; Ujma, Z.;
Gruszka, I. Electric Modulus Approach to the Analysis of Electric
Relaxation in Highly Conducting (Na0.75Bi0.25)(Mn0.25Nb0.75)O3
Ceramics. J. Phys. D: Appl. Phys. 2005, 38, 1450−1460.
(72) Jonscher, A. K. A New Understanding of the Dielectric
Relaxation of Solids. J. Mater. Sci. 1981, 16, 2037−2060.
(73) Safdari, M.; Fischer, A.; Xu, B.; Kloo, L.; Gardner, J. M.
Structure and Function Relationship in Alkylammonium Lead(II)
Iodide Solar Cells. J. Mater. Chem. A 2015, 3, 9201−9207.
(74) Eames, C.; Frost, J. M.; Barnes, P. R. F.; O’Regan, B. C.; Walsh,
A.; Islam, M. S. Ionic Transport in Hybrid Lead Iodide Pervskite Solar
Cells. Nat. Commun. 2015, 6, No. 7497.
(75) Yang, T.-Y.; Gregori, G.; Pellet, N.; Grätzel, M.; Maier, J. The
Significance of Ion Conduction in a Hybrid Organic-Inorganic Leqad-
Iodide-Based Perovskite Photosensitizer. Angew. Chem., Int. Ed. 2015,
54, 7905−7910.
(76) Lin, Y.; Bai, Y.; Fang, Y.; Wang, Q.; Deng, Y.; Huang, J.
Suppressed Ion Migration in Low-Dimensional Perovskites. ACS
Energy Lett. 2017, 2, 1571−1572.
(77) Kubelka, P.; Munk, F. Ein Beitrag zur Optik der Farbanstriche.
Z. Tech. Phys. 1931, 12, 593−601.
(78) Amat, A.; Mosconi, E.; Ronca, E.; Quarti, C.; Umari, P.;
Nazeeruddin, M. K.; Grätzel, M.; De Angelis, F. Cation-Induced
Band-Gap Tuning in Organohalide Perovskites: Interplay of Spin-
Orbit Coupling and Octahedra Tilting. Nano Lett. 2014, 14, 3608−
3616.
(79) Smith, M. D.; Connor, B. A.; Karunadasa, H. I. Tuning the
Luminescence of Layered Halide Perovskites. Chem. Rev. 2019, 119,
3104−3139.
(80) DeCrescent, R. A.; Venkatesan, N. R.; Dahlman, C. J.; Kennard,
R. M.; Zhang, X.; Li, W.; Du, X.; Chabinyc, M. L.; Zia, R.; Schuller, J.
A. Bright Magnetic Dipole Radiation from Two-Dimensional Lead-

8575 DOI: 10.1021/acs.chemmater.9b03775


Chem. Mater. 2019, 31, 8563−8575

You might also like