X-Ray Diffraction and Raman-2
X-Ray Diffraction and Raman-2
X-Ray Diffraction and Raman-2
2
X-ray diffraction and Raman
spectroscopy for lead halide
perovskites
Mohammad Ziaur Rahman and Tomas Edvinsson
Department of Engineering Sciences, Solid State Physics, Ångström Laboratory,
Uppsala University, Uppsala, Sweden
2.1 Introduction
The broad field of perovskites have been attractive for condensed matters physics and
material sciences for many years due to the breath of physical properties that can be
found in the system. Oxide and fluoride perovskites have been extensively studied over
the years since they exhibit a plethora of intriguing photonic and electronic properties,
ionic conductivity, multiferroicity, piezoelectricity, superconductivity, and metal-
insulator transitions, where their properties can be tuned by the octahedral tilting with
inclusion of different metals [1]. Hybrid perovskite solar cells (HPSCs), based on
organic-inorganic lead halide materials, have recently emerged as a promising class of
materials in photovoltaic technology [2,3]. Following the report of solid state perovskite
solar cells with efficiencies up to 10% in 2012 [4], a great surge in world-wide research
efforts into perovskite solar cells research can be noticed- evident by more than 4000
publications on HPSCs in 2018. These efforts and the previous experience in the other
hybrid organic-inorganic solar cell systems (e.g. the dye-sensitized solar cell) lead to a
certified record efficiency surpassing 23% in lab scale devices in 2018 [5]. Other possible
application for the hybrid metal halide perovskite materials are hybrid light emitting
diodes (HLEDs) or utilization as X-ray scintillators [6].
Perovskites are materials with a crystal structure analogous with CaTiO3, having a gen-
eral chemical formula ABX3, where A is a large cation, B a smaller sized cation, and X is
an anion (Fig. 2.1). In the lead halide perovskites that are investigated for photovoltaics, B
is normally Pb21, X is a halide (i.e. I-, Br-, Cl-), while A is a small organic dipolar cation,
FIGURE 2.1 (A) Perovskite crystal structure ABX3 for a cubic symmetry and (B) the cubic structure of the
MAPbI3 hybrid perovskite [7]. Reproduced with permission from reference T.J. Jacobsson, M. Pazoki, A. Hagfeldt,
T. Edvinsson, Goldschmidt’s rules and strontium replacement in lead halogen perovskite solar cells: theory and preliminary
experiments on CH3NH3SrI3, J. Phys. Chem. C 119 (2015) 2567325683. Copyright 2015, American Chemical Society.
FIGURE 2.2 (Top) Optical photographs, (A) PL of films, and (B) band gap change with linear bandgap
tuning by I-to-Br halogen exchange. (C) Incident-photon-to-current-efficiency (IPCE) for solar cell devices with
Br fraction up to 10%. [9]. Adapted with permission from reference B.-w. Park, B. Philippe, S.M. Jain, X. Zhang,
T. Edvinsson, H. Rensmo, et al., Chemical engineering of methylammonium lead iodide/bromide perovskites: tuning of
opto-electronic properties and photovoltaic performance, J. Mater. Chem. A 3 (2015) 2176021771. Copyright 2015, Royal
Society of Chemistry.
semiconductor tandem systems. Toward the design of highly efficient tandem HPSCs, or
fully inorganic PSCs, it is crucial to quantify their crystalline structure and symmetries.
Because, such understanding is vital to control the structure as well as the fundamental
optoelectronic properties for a resilient device structure with a durable performance.
In this context, X-ray diffraction and Raman spectroscopy provide useful information of
the material properties, such as the crystal structure, symmetries, displacements, defects,
micro-crystallinity, orientations, vibrations, phonon lifetimes, and their interdependence
on material alterations. These techniques contribute with crucial information that is neces-
sary to find structure-to-property relationships from materials to function.
monochromatic X-rays that are elastically scattered, W.H Bragg and W.L Bragg
constructed a geometric analog satisfying the Laue condition, where the constructive dif-
fraction could be formulated into the Bragg equation
2dhkl sinθb 5 nλ (2.1)
Where dhkl is the distance between diffraction planes with Miller index (h k l ), λ is the
wavelength of the incoming X-ray, θB is the Bragg angle, and n is a positive integer
representing the diffraction order. The Laue constraint for the intensity I(R)- maximum
is now naturally fulfilled by the constructive interference for a fixed wavelength and
replacing the general angle θ with the angle for which the intensity is at maximum, θB.
The geometric situation is illustrated in Fig. 2.3A in a Bragg-Brentano setup where the
two extra path lengths of dhkl sinθ for the bottom lattice plane elastic scattering can be
seen and XRD diffractograms for cubic Pm-3m and tetragonal I4/mcm MAPbI3 (Fig. 2.3C).
Although the maximum intensity at angle θ is instructive to infer the set of distances in
the crystalline materials, and thus reflecting the internal structure and symmetry, also the
peak shape contains information from the number of crystalline planes to form or to cancel
FIGURE 2.3 (A) Schematic illustration of a Bragg-Brentano geometry XRD setup, (B) illustration of a XRD
setup with multiple angles for generation of 2D q-range and construction of pole figures [16]. (C) Synchrotron
XRD patterns of MAPbI3 for the cubic Pm-3m and the tetragonal I4/mcm structure, and R-symmetry point super-
lattice peaks [17]. (D) GIWAXS images and structure orientation information of polycrystalline and (E) near
single-crystalline (BA)2(MA)3Pb4I13 RuddlesdenPopper phases [18]. (B) Adapted with permission from
reference K. Inaba, S. Kobayashi, K. Uehara, A. Okada, S.L. Reddy, T. Endo, High resolution X-ray diffraction analyses of
(La, Sr)MnO3/ZnO/Sapphire(0001) double heteroepitaxial films, Adv. Mater. Phys. Chem. 03 (2013) 7289. Copyright
2013, Scientific Research. (C) Reproduced with permission from reference P.S. Whitfield, N. Herron, W.
E. Guise, K. Page, Y.Q. Cheng, I. Milas, et al., Structures, phase transitions and tricritical behavior of the hybrid perovskite
methyl ammonium lead iodide, Sci. Rep. 6 (2016) 35685. Copyright 2016 Nature Publishing Group. (D and E) Reproduced
with permission from reference H. Tsai, W. Nie, J.-C. Blancon, C.C. Stoumpos, R. Asadpour, B. Harutyunyan, et al., High-
efficiency two-dimensional RuddlesdenPopper perovskite solar cells, Nature 536 (2016) 312. Copyright 2016 Nature
Publishing Group.
the non-constructive interference signal. Peak shape analysis can here be used to extract
local disorder, strain, and crystallite size. The ability to separate of distortion and particle
size broadening in X-ray patterns, however, as well as coupling effect from local disorder
and strain have to be taken into care when performing detailed peak shape analysis
[13,14]. When the sample is single crystalline or textured one can quantify the crystal sym-
metry or crystallite orientations by extending the diffraction analysis to other directions
compared to the Bragg angle, either by using a wide-angle detector, repositioning the
detector to angular ranges perpendicular to the Bragg angle, or rotate the sample
(Fig. 2.3B).
In the most general case, one can construct a pole figure to fully characterize the distri-
bution of diffractions from a textured sample with a variation in crystallite orientations by
rotating the sample axes while having the corresponding Bragg reflection of interest fixed,
successively change the Bragg angle, and add data from an additional perpendicular axis
rotations (rocking curve) of the sample. The two first parts of this are illustrated in
Fig. 2.3D and E where diffraction pattern from gracing incident wide angle X-ray scatter-
ing (GIWAXS) of polycrystalline and near single-crystalline perovskite derived
RuddlesdenPopper (RP) phases are shown. The crystallographic unit cell of RP phases
can be seen as an elongation of the dimensions in the perovskite structure, keeping layers
of perovskite symmetries (ABX3) in between with the general RP phase composition for-
mula An11BnX3n11 [15]. For more complicated A-site cation compositions in RP phases,
they can also be represented as An-1A0 2BnX3n11, where A and A0 (A-prime) represent two
different cations and B refers to the Pb, Bi, or Sn metal cations usually incorporated in the
hybrid metal halides. The A cations are located in the perovskite layer with a 12-fold
cuboctahedral coordination to the anions while the A0 cations have a 9-fold coordination
and are located at the interface between the perovskite layer and the intermediate layer
before the next perovskite layer.
Apart from a general structure determination and peak shape analysis, more details can
be extracted from diffraction by analyzing the diffraction intensities or shifts in more
detail. First, one should note that X-rays are diffracted from the electron clouds around
the nuclei forming the lattice planes where any anisotropic polarization will give an error
in the determination of the derived nuclei positions. In these cases, neutron diffraction
characterization is advised for a high resolved structure determination. The amplitude of
the scattering (within the Fraunhofer condition) from electrons around a nuclei can be
described by a phase factor e2iΔk r(i,j) containing the scattering vector Δk and the pair-
wise distance between all pairs of point charges r(i,j). It can be formulated into an integral
of the contribution from a charge distribution into an atomic form factor [19].
ð
fatom 5
atom
ρe ðrÞ e2Δk r dr (2.2)
where ρe(r) is the electron density charge distribution around the atom, and can formally
be determined via the square of the electronic wave function. For a distribution of atoms,
one can combine the individual atomic form factors into a structure factor
PN
F5 fatom eiΔkrn where fatom is the integral in Eq. 2.2 and rn denotes the position of the
n51
n-th atom in the structure. Combining this with the Laue condition and the notation for
Miller indices, one could obtain
X
N
FðhklÞ 5 fatom e½i2πðhxn 1kyn 1lzn Þ (2.3)
n51
For the simple cubic systems, with N 5 1 and the atoms in the primitive unit cells at
(0,0,0), reflections can be seen for all miller index combinations. For the FCC lattice with
N 5 2 and cubic sub-symmetry with at the basis (0,0,0) and (1/2, 1/2, 1/2), the structure factor
becomes 4fatom when all hkl are even or odd and 0 otherwise. Of interest here is a general
form factor for the perovskite structure (ABX3), here exemplified for the cubic structure
with space group Pm3m with the A-site cation at (0,0,0), B cation at (1/2, 1/2, 1/2), and 3 X
FIGURE 2.4 (A) Illustration of how GIWAXRD and GISAXS can be used for texture analysis in thin films
depending on the required dimensions where (B) represents randomly oriented arrangements of crystallites with
circular diffraction rings, except for the slightly oriented (200) reflection. (C) corresponds to a textured and
preferentially oriented films with diffraction patterns as arcs of different intensities and (D) corresponds to a
highly oriented films produce more localized diffraction spots/ellipses [24]. Reproduced with permission from
reference J. Rivnay, S.C.B. Mannsfeld, C.E. Miller, A. Salleo, M.F. Toney, Quantitative determination of organic semicon-
ductor microstructure from the molecular to device scale, Chem. Rev. 112 (2012) 54885519. Copyright 2012, American
Chemical Society.
Thermal, light- and chemical induced phase transformations are also important to
characterize and relates to mechanical stability during day-night operation, and long-
time stability. Certainly, the promises of lead halide perovskites (LHPs) as a solar cell
material is undeniable but it is very important that the crystal structure integrity and
chemical composition can be retained during extended use. LHP undergoes degradation
in contact with water, humidity, light, and temperature [25,26]. The hybrid lead halogen
perovskites are water soluble salts where uncapsulated perovskites take on water very
FIGURE 2.5 (A) Illustration of the crystal structure of PbI2 and MAPbI3 at different temperatures together
with the αFAPbI3 and δFAPbI3 phase and (B and C) comparison between the experimental and theoretical
XRD patterns [36]. Reproduced with permission from reference A. Binek, F.C. Hanusch, P. Docampo, T. Bein, Stabilization
of the trigonal high-temperature phase of formamidinium lead iodide, J. Phys. Chem. Lett. 6 (2015) 12491253. Copyright
2015, American Chemical Society.
of FA [36]. This can be observed in X-ray diffraction measurements by a shift of the dif-
fraction pattern to lower angles. The XRD patterns of the synthesized yellow FAPbI3 and
stabilized black FAPbI3 and their theoretical patterns are illustrated in Fig. 2.5B and C.
Temperature-dependent XRD measurements is widely practiced to investigate the
phase transformation between the between different phases of LHP. In XRD patterns, a
few peaks disappear when the tetragonal phase transforms into the cubic structure whilst
some double peaks merge into single peaks. A peak shift in the XRD with the temperature
not only gives information about the phase transformation, but also about thermal expan-
sion coefficients. With this information, one can evaluate how the materials behaves under
thermal stress. A contour plot of individual groups of peaks in XRD measurements as a
function of the temperature between 35 and 95 C is given in in Fig. 2.6, in which the
phase transformation between the tetragonal and cubic phases is clearly seen. At room
temperature, the data perfectly agree with a tetragonal structure. However, some peaks
disappear and some double peaks merge into single peaks at higher temperatures that
matches the cubic structure, as is seen in Fig. 2.6. Noticeably, the phase transformation
FIGURE 2.7 (A) Comparison between the measured (a) and calculated (b) Raman spectra of MAPbI3 in the
orthorhombic structure. (B) Schematic representation of the most important Raman-active modes of the PbI3
network in MAPbI3. The symmetry and calculated frequency of each normal mode are indicated at the top of the
each panel, and the solid black lines indicate the unit cell [52]. Reproduced with permission from reference M.A. Pérez-
Osorio, Q. Lin, R.T. Phillips, R.L. Milot, L.M. Herz, M.B. Johnston, et al., Raman spectrum of the organicinorganic halide
perovskite CH3NH3PbI3 from first principles and high-resolution low-temperature raman measurements, J. Phys. Chem.
C 122 (2018) 2170321717. Copyright 2018, American Chemical Society.
Halide substitution in metal halide perovskites (MHPs) with the nominal composition
CH3NH3PbI2X, where X is I, Br, or Cl, influences the morphology, charge quantum yield,
and local interaction with the organic MA cation. Raman spectroscopies combined with
theoretical vibration calculations supported the early findings that iodide-chloride
perovskites phase separate and have also revealed that halide substitution with halides
forming a shorter bond length to Pb can delocalizes the charge to the MA cation;
therefore, liberate the vibrational movement of the MA cation that lead to a more adap-
tive organic phase [60].
Raman spectroscopy with 532 nm laser excitation are performed within the absorption
profile of MHP materials and are thus under resonance (electronically exciting) conditions,
providing information on vibrations in excited state and can also give clues for the charge
separation/transfer mechanisms. Fundamental vibrations in the isolated clusters and the
trends in the splitting of the degenerate states when different halogens are included
were investigated with low-frequency Raman measurements (down to 10 cm21) and non-
periodic DFT calculations with an emphasis on the ordering of the peaks to determine
Raman properties of MHPs [60].
The experimental Raman spectrum for MAPbI3 shows vibration peaks at 40, (54), (63),
71, 94, 108, 135, and 145 cm21, whereas MAPbI2Cl shows corresponding peaks at 40, NA,
NA, 71, 97, 110, and a broad peak at 166 cm21 as seen in Fig. 2.8. The NA (not applicable)
notation, emphasize that the peaks cannot be resolved and the spectra instead show a
shoulder in that area. Besides the strongest Raman peaks at 6973, 9497, and
108110 cm21, peaks also at 40 and 54 cm21 depends on halide composition during
synthesis.
Periodic DFT calculations provides a suitable model system for single-crystalline
materials where the different orientation of the cations must be considered to be
periodic. For non-periodic or non-crystalline materials, cluster calculations can instead
be informative for the study of local effects in the inorganic octahedron and the behavior
of the organic cation. In the present example, an inorganic octahedron unit cluster PbX6
with Oh symmetry is used with two MA-dipole canceling cations. A PbI6 octahedron has
15 internal degrees of freedom (3 N3, where N is the number of iodine atoms).
According to group theoretical representation [61], it can be written as
A1 g 1 Eg 1 2T1 u 1 T2 g 1 T2 u, where A1 g, Eg, and T2 g are Raman active, the two T1 u
modes are IR active, and T2 u is a silent mode (neither Raman or IR active). PbI6 as a
molecular unit within the lattice belongs to the Oh symmetry group. Deviation from this
symmetry would result in splitting of the degenerate states, and eventually complete
removal of symmetry and 15 different bands.
Calculated Raman spectra for PbX6 and (MA)2PbX6 clusters are shown in Fig. 2.8AG,
and experimental data for MAPbI3 and MAPbI2Cl are shown in Fig. 2.8H and I. The
calculations reveal three different groups of vibrational modes, a triply degenerate
asymmetric X-Y, X-Z, and Y-Z vibrations (mode A), a double-degenerate asymmetric
“breathing” (mode B), and non-degenerated symmetric “breathing” (mode C), as well as
the MA vibrations (rotation, wagging, MAMA stretch) shown in Fig. 2.8J. A shift in the
Raman peaks in (MA)2PbX6 clusters to higher wavenumbers (energy) compared to those
in the PbX6 clusters were observed, caused by the organic cations that extend the motion
of X from the Pb21 atom. The calculated Raman spectrum of (MA)2PbCl6 (see Fig. 2.8C)
differs somewhat from the other clusters while mode A shifts to lower wavenumbers
and mode B appears as two peaks. Notably, all (MA)2PbX6 structures showed Raman
activity of the MA groups between 140 and 180 cm21 and Pb-I vibrations at lower
wavenumbers.
DFT calculations on mixed halide clusters, such as (MA)2PbI5Cl and (MA)2PbI4Cl2 has
previously shown that a single Cl substitution does not result in large differences in calcu-
lated Raman vibrational spectra compared to the (MA)2PbI6 cluster, while larger changes
are found for two introduced Cl atoms. These results carries valuable information for the
possibility to identify doped phases. Fig. 2.8F and G show “Raman activities compared to
the (MA)2PbI6 cluster, which is assigned to asymmetric vibrations of PbI between 40 and
95 cm21. Moreover, there are two small appearances of additional vibration modes, for
instance N10 (green color, NCl stretch via H at 240 cm21) and D2 (green color, PbCl
stretch at 190 cm21) in Fig. 2.8F. The case of (MA)2PbI4Cl2 with double Cl substitution
in the octahedral unit in Fig. 2.8G has rather different vibration modes, viz., mode D4
(violet color, asymmetric PbCl stretch, 7582 cm21) and highly increased intensity of
mode D20 (violet color, asymmetric PbCl stretch at 185 cm21) compared with mode D2
(green colored solid line in Fig. 2.8F). Moreover, in Fig. 2.8G the Raman activity of mode
M4 (violet color, at 138143 cm21) is shown to shift to lower wavenumber compared to
mode M3 (green color, Fig. 2.8F), but mode N20 (violet color, at 247 cm21) is shifted to
higher wavenumber for mode M20 (green color, Fig. 2.8F)” [60].
Observing the new vibration signatures with relative intensity shifts in the Raman
activities of (MA)2PbI4Cl2 compared to the undoped analog, it is thus possible to identify
local halide substitution or different phases. For example, both the single- and double-Cl-
substituted OMHPs have shown commonly a decreased Raman intensity of modes C3 and
C4 (green and violet colors) compared to mode C0 (black color).
The DFT-calculated Raman spectra for three vibrational modes and the experimental
Raman signal of MAPbI3 (Fig. 2.8A, H, and I) showed three shoulders or peaks in the
70120 cm21 associated to modes A, B, and C. Additional four peaks (m, d, e, and f)
appear between 140 and 400 cm21 are assigned to MA rotation, MA wagging, and sym-
metric MAMA stretch [60,62]. Because of the limitation of cluster models to emulate the
full symmetry of the crystal, cluster analysis should only be taken as representative of the
local modes of the crystalline system, but can as such be used to analyze the details of
chloride effects during crystallization [63] or effects from local charge delocalization and
transfer [60]. If one analyze the Raman spectrum of MAPbI2Cl in more detail, it can be
seen that the mode B as peak “b0 ” shows peak shift to lower intensity. This is correlated
with disordered inorganic frameworks where higher Raman activities have been observed
previously from the due to increased electronphonon interaction between the central cat-
ionic metal and anionic oxygen [64]. This same effect holds for MAPbI3 and the
subMAPbI3 crystal in MAPbI2Cl on modes A and C. In the range of 136150 cm21, that
MAPbI3 sample shows 1.3 times higher Raman intensity at 143 cm21 than that of
MAPbI2Cl at B145 cm21 as is seen in Fig. 2.8.
During resonance excitation at 532, absence of the feature at 143 cm21 nm implies the
disappearance of initial and final state polarization for the corresponding Raman
rotation/wagging in the MA cation. The difference at 143145 cm21 between MAPbI3
and MAPbI2Cl samples seems well correlated with the local charge transfer yield [60]
and could be part of the origin of the improvements in the mixed halogen systems
[65,66]. DFT calculations for OMHPs display the transitions of charges from the HOMO
to LUMO (with LUMO 1 1 and LUMO 1 2). For example, the HOMO of 2MAPbI6
showed an I 5p π-bonding orbital while its LUMO, LUMO 1 1, and LUMO 1 2 were
decomposed to “Pb(6 s)I(5p)” σ-antibonding, “Pb(6p)I(5p)” σ-antibonding, and “Pb
(6 s)I(5p)” σ-antibonding orbitals, respectively [60]. The observed 7 times lower inten-
sity for the 143 cm21 mode in the MAPbI2Cl sample corresponds to a loss of polarization
on mode M of internal MAPbI3. A photoexcited state that cannot change polarizability of
an MA cation radical or neutral MA molecule can then play an important role as an
organic cation charge stabilizer or a neutral dipolar molecule in the cage of the inorganic
framework [67]. The understanding of these phenomena lies deep within the origin of
the relative intensities in Raman measurements that comes from the losses of the electro-
magnetic energy in the incident light to an excited vibration level and the change in
FIGURE 2.9 (A) Room temperature optical bleaching of CH3NH3PbI3 perovskite by pyridine vapor. (B)
Sequential resonance Raman spectroscopy of the recovery of CH3NH3PbI3 film in ambient air after bleaching
with pyridine. The marked yellow region show vibrations characteristic of isolated plumbates in the intermediate
phase. (C) The ν(NC) breathing mode of the pyridine ring unperturbed and bonded and the calculated shift
when interacting with possible species in the intermediate bleached phase: (pyr)IH and (pyr)PbI2. Experimental
and theoretical assignments of the symmetric ring breathing and H-wagging modes of pyridine (top) and experi-
mental Raman shifts for pyridine in liquid capillary and when interacting with the perovskite during bleaching.
Reproduced with permission from reference S.M. Jain, Z. Qiu, L. Häggman, M. Mirmohades, M.
B. Johansson, T. Edvinsson, et al., Frustrated Lewis pair-mediated recrystallization of CH3NH3PbI3 for improved optoelec-
tronic quality and high voltage planar perovskite solar cells, Energy Environ. Sci. 9 (2016) 37703782. Copyright Royal
Society of Chemistry 2016.
2.4 Conclusions
X-ray diffraction and Raman spectroscopy provide key structural and vibrational
information of LHP materials, containing information on the crystal structure, symme-
tries, displacements, defects, micro-crystallinity, orientations, thermal expansion, phase
transitions, vibrations, effects from doping, and local charge transfer effects and their
interdependence on material alterations. XRD techniques are more or less necessary to
elucidate the structural peripteries in modern material science where Raman spectros-
copy can be used as an important complement that provide vibrations characterization
under non-resonant and resonant conditions and aid in determination of structure-to-
property relationships from materials to function. The techniques can provide crystalline
and chemical features from sub-Ångström variations in the structure useful for both
understanding the molecular chemistry during creation and the structure and properties
of the resulting LHP crystalline materials. The chapter briefly introduces different com-
monly used XRD approaches as well as some recent results from their applications.
Raman spectroscopy is here a complementing technique that do not only provides infor-
mation about the chemical composition and phases in crystalline materials, but can also
be used for characterization of vibration in amorphous structures, molecular materials,
solvents, and gases. With this knowledge, one can extract chemical identities and pro-
cesses during syntheses as well as details in the crystalline materials via the observed
local or extended lattice vibrations, bonding interactions, orientations, symmetries, and
phonon modes to complement the structural characterization from XRD. A more detailed
analysis of relative intensities in resonance Raman measurements, related to the losses of
the electromagnetic energy in the incident light to an excited vibration level and the
change in polarizability of the electron cloud during the vibration, additionally allows
investigations of the primary light-matter interaction and subsequent charge
delocalization.
Acknowledgment
We acknowledge financial support from the Swedish Research Council (201503814), the Swedish Research
Council for sustainable development (2016-00908), and the Swedish Energy Agency (P44648-1).
References
[1] T.R.J.D. Perovskites, Structure-Property Relationships, John Wiley & Sons, Singapore, 2016.
[2] M.A. Green, A. Ho-Baillie, Perovskite solar cells: the birth of a new era in photovoltaics, ACS Energy Lett. 2
(2017) 822830.
[3] M.A. Green, A. Ho-Baillie, H.J. Snaith, The emergence of perovskite solar cells, Nat. Photonics 8 (2014) 506.
[4] M.M. Lee, J. Teuscher, T. Miyasaka, T.N. Murakami, H.J. Snaith, Efficient hybrid solar cells based on meso-
superstructured organometal halide perovskites, Science 338 (2012) 643647.
[5] https://www.nrel.gov/pv/assets/images/efficiency-chart.png.
[6] Y. Zhao, K. Zhu, Organicinorganic hybrid lead halide perovskites for optoelectronic and electronic applica-
tions, Chem. Soc. Rev. 45 (2016) 655689.
[7] T.J. Jacobsson, M. Pazoki, A. Hagfeldt, T. Edvinsson, Goldschmidt’s rules and strontium replacement in lead
halogen perovskite solar cells: theory and preliminary experiments on CH3NH3SrI3, J. Phys. Chem. C 119
(2015) 2567325683.
[8] J.H. Noh, S.H. Im, J.H. Heo, T.N. Mandal, S.I. Seok, Chemical management for colorful, efficient, and
stable inorganic-organic hybrid nanostructured solar cells, Nano Lett. 13 (2013) 17641769.
[9] B.-w. Park, B. Philippe, S.M. Jain, X. Zhang, T. Edvinsson, H. Rensmo, et al., Chemical engineering of
methylammonium lead iodide/bromide perovskites: tuning of opto-electronic properties and photovoltaic
performance, J. Mater. Chem. A 3 (2015) 2176021771.
[10] T.A. Berhe, W.-N. Su, C.-H. Chen, C.-J. Pan, J.-H. Cheng, H.-M. Chen, et al., Organometal halide perovskite
solar cells: degradation and stability, Energy Environ. Sci. 9 (2016) 323356.
[11] A. Mei, X. Li, L. Liu, Z. Ku, T. Liu, Y. Rong, et al., A hole-conductorfree, fully printable mesoscopic perov-
skite solar cell with high stability, Science 345 (2014) 295298.
[12] W. Chen, Y. Wu, Y. Yue, J. Liu, W. Zhang, X. Yang, et al., Efficient and stable large-area perovskite solar cells
with inorganic charge extraction layers, Science 350 (2015) 944948.
[13] B.E. Warren, B.L. Averbach, The separation of cold-work distortion and particle size broadening in X-Ray
patterns, J. Appl. Phys. 23 (1952). 497-497.
[14] J. Rivnay, R. Noriega, R.J. Kline, A. Salleo, M.F. Toney, Quantitative analysis of lattice disorder and crystallite
size in organic semiconductor thin films, Phys. Rev. B 84 (2011) 045203.
[15] S.N. Ruddlesden, P. Popper, The compound Sr3Ti2O7 and its structure, Acta Crystallogr. 11 (1958) 5455.
[16] K. Inaba, S. Kobayashi, K. Uehara, A. Okada, S.L. Reddy, T. Endo, High resolution X-ray diffraction analyses
of (La, Sr)MnO3/ZnO/Sapphire(0001) double heteroepitaxial films, Adv. Mater. Phys. Chem. 03 (2013)
7289.
[17] P.S. Whitfield, N. Herron, W.E. Guise, K. Page, Y.Q. Cheng, I. Milas, et al., Structures, phase transitions and
tricritical behavior of the hybrid perovskite methyl ammonium lead iodide, Sci. Rep. 6 (2016) 35685.
[18] H. Tsai, W. Nie, J.-C. Blancon, C.C. Stoumpos, R. Asadpour, B. Harutyunyan, et al., High-efficiency two-
dimensional RuddlesdenPopper perovskite solar cells, Nature 536 (2016) 312.
[19] R.W. James, The Optical Principles of the Diffraction of X-rays, Cornell University Press, New York, 1965.
[20] A.M. Glazer, The classification of tilted octahedra in perovskites, Acta Crystallogr. Sect. B 28 (1972) 33843392.
[21] A. Glazer, Simple ways of determining perovskite structures, Acta Crystallogr. Sect. A 31 (1975) 756762.
[22] C.J. Howard, H.T. Stokes, Group-theoretical analysis of octahedral tilting in perovskites, Acta Crystallogr.
Sect. B 54 (1998) 782789.
[23] C.J. Howard, H.T. Stokes, Structures and phase transitions in perovskites - a group-theoretical approach.
This article is dedicated to Helen D. Megaw (1907-2002), in appreciation of her many contributions to the
crystallography of inorganic and mineral compounds, including her seminal studies of perovskites. Some of
the material was presented by CJH in the Megaw memorial session at the 21st European Crystallography
Meeting, Durban, South Africa, August 2003, Acta Crystallogr. Sect. A 61 (2005) 93111.
[24] J. Rivnay, S.C.B. Mannsfeld, C.E. Miller, A. Salleo, M.F. Toney, Quantitative determination of organic semi-
conductor microstructure from the molecular to device scale, Chem. Rev. 112 (2012) 54885519.
[25] C.C. Boyd, R. Cheacharoen, T. Leijtens, M.D. McGehee, Understanding degradation mechanisms and
improving stability of perovskite photovoltaics, Chem. Rev. 119 (2019) 34183451.
[26] C.C. Boyd, R. Cheacharoen, K.A. Bush, R. Prasanna, T. Leijtens, M.D. McGehee, Barrier design to prevent
metal-induced degradation and improve thermal stability in perovskite solar cells, ACS Energy Lett. 3 (2018)
17721778.
[50] A. Maalej, Y. Abid, A. Kallel, A. Daoud, A. Lautié, F. Romain, Phase transitions and crystal dynamics in the
cubic perovskite CH3NH3PbCl3, Solid State Commun. 103 (1997) 279284.
[51] J.D. Jackson, Classical Electrodynamics, 3rd ed., Wiley, New York, 1999.
[52] M.A. Pérez-Osorio, Q. Lin, R.T. Phillips, R.L. Milot, L.M. Herz, M.B. Johnston, et al., Raman spectrum of the
organicinorganic halide perovskite CH3NH3PbI3 from first principles and high-resolution low-temperature
raman measurements, J. Phys. Chem. C 122 (2018) 2170321717.
[53] D.L. Rousseau, R.P. Bauman, S.P.S. Porto, Normal mode determination in crystals, J. Raman Spectrosc. 10
(1981) 253290.
[54] P.Y. Yu, M. Cardona, Vibrational properties of semiconductors, and electron-phonon interactions, in: P.Y.
Yu, M. Cardona (Eds.), Fundamentals of Semiconductors: Physics and Materials Properties, Springer Berlin
Heidelberg, Berlin, Heidelberg, 1996, pp. 99147.
[55] R.G. Niemann, A.G. Kontos, D. Palles, E.I. Kamitsos, A. Kaltzoglou, F. Brivio, et al., Halogen effects on
ordering and bonding of CH3NH3 1 in CH3NH3PbX3 (X 5 Cl, Br, I) hybrid perovskites: a vibrational spec-
troscopic study, J. Phys. Chem. C 120 (2016) 25092519.
[56] J.-X. Cheng, X.S. Xie, Coherent anti-stokes raman scattering microscopy: instrumentation, theory, and appli-
cations, J. Phys. Chem. B 108 (2004) 827840.
[57] J. Even, L. Pedesseau, J.-M. Jancu, C. Katan, Importance of spinorbit coupling in hybrid organic/inorganic
perovskites for photovoltaic applications, J. Phys. Chem. Lett. 4 (2013) 29993005.
[58] D.B. Mitzi, Templating and structural engineering in organicinorganic perovskites, J. Chem. Soc., Dalton
Trans. (2001) 112.
[59] A. Poglitsch, D. Weber, Dynamic disorder in methylammoniumtrihalogenoplumbates (II) observed by milli-
meter-wave spectroscopy, J. Chem. Phys. 87 (1987) 63736378.
[60] B.-w. Park, S.M. Jain, X. Zhang, A. Hagfeldt, G. Boschloo, T. Edvinsson, Resonance Raman and excitation
energy dependent charge transfer mechanism in halide-substituted hybrid perovskite solar cells, ACS Nano
9 (2015) 20882101.
[61] G. Blasse, A.F. Corsmit, Electronic and vibrational spectra of ordered perovskites, J. Solid State Chem. 6
(1973) 513518.
[62] C. Quarti, G. Grancini, E. Mosconi, P. Bruno, J.M. Ball, M.M. Lee, et al., The Raman spectrum of the
CH3NH3PbI3 hybrid perovskite: interplay of theory and experiment, J. Phys. Chem. Lett. 5 (2014) 279284.
[63] G. Grancini, S. Marras, M. Prato, C. Giannini, C. Quarti, F. De Angelis, et al., The impact of the crystallization
processes on the structural and optical properties of hybrid perovskite films for photovoltaics, J. Phys. Chem.
Lett. 5 (2014) 38363842.
[64] G. Coslovich, B. Huber, W.S. Lee, Y.D. Chuang, Y. Zhu, T. Sasagawa, et al., Ultrafast charge localization in a
stripe-phase nickelate, Nat. Commun. 4 (2013) 2643.
[65] J.M. Frost, K.T. Butler, F. Brivio, C.H. Hendon, M. van Schilfgaarde, A. Walsh, Atomistic origins of high-
performance in hybrid halide perovskite solar cells, Nano Lett. 14 (2014) 25842590.
[66] T. Umebayashi, K. Asai, T. Kondo, A. Nakao, Electronic structures of lead iodide based low-dimensional
crystals, Phys. Rev. B 67 (2003).
[67] H.L. Bostrom, J.A. Hill, A.L. Goodwin, Columnar shifts as symmetry-breaking degrees of freedom in molecu-
lar perovskites, Phys. Chem. Chem. Phys. 18 (2016) 3188131894.
[68] S.M. Jain, B. Philippe, E.M.J. Johansson, B.-w. Park, H. Rensmo, T. Edvinsson, et al., Vapor phase conversion
of PbI2 to CH3NH3PbI3: spectroscopic evidence for formation of an intermediate phase, J. Mater. Chem. A 4
(2016) 26302642.
[69] R. Gottesman, L. Gouda, B.S. Kalanoor, E. Haltzi, S. Tirosh, E. Rosh-Hodesh, et al., Photoinduced reversible
structural transformations in free-standing CH3NH3PbI3 perovskite films, J. Phys. Chem. Lett. 6 (2015)
23322338.
[70] S.M. Jain, Z. Qiu, L. Häggman, M. Mirmohades, M.B. Johansson, T. Edvinsson, et al., Frustrated Lewis pair-
mediated recrystallization of CH3NH3PbI3 for improved optoelectronic quality and high voltage planar
perovskite solar cells, Energy Environ. Sci. 9 (2016) 37703782.
[71] B.R. Vincent, K.N. Robertson, T.S. Cameron, O. Knop, Alkylammonium lead halides. Part 1. Isolated PbI64 2
ions in (CH3NH3)4PbI6•2H2O, Can. J. Chem. 65 (1987) 10421046.
[72] D.W. Stephan, G. Erker, Frustrated Lewis pair chemistry: development and perspectives, Angew. Chem. Int.
Ed. Engl. 54 (2015) 64006441.