Abstract

Lead halide perovskites are extensively investigated as efficient solution-processable materials for photovoltaic applications. The greatest stability and performance of these compounds are achieved by mixing different ions at all three sites of the APbX3 structure. Despite the extensive use of mixed lead halide perovskites in photovoltaic devices, a detailed and systematic understanding of the mixing-induced effects on the structural and dynamic aspects of these materials is still lacking. The goal of this review is to summarize the current state of knowledge on mixing effects on the structural phase transitions, crystal symmetry, cation and lattice dynamics, and phase diagrams of three- and low-dimensional lead halide perovskites. This review analyzes different mixing recipes and ingredients providing a comprehensive picture of mixing effects and their relation to the attractive properties of these materials.
1. Introduction
In the past decade, lead halide perovskites APbX3 (A = molecular or inorganic cation, X = halide anion) and related materials have received unprecedented attention due to their potential applications for photovoltaic devices such as efficient solar cells and light-emitting diodes (LEDs).1−7 The power conversion efficiency of solar cells based on these solution-processable compounds experienced an extraordinary boost and currently exceeds 25%,8,9 with the best performance and stability achieved using perovskites with mixed compositions.10,11
Depending on the type and relative size of the constituent ions, lead halide perovskites can form either three-dimensional (3D) or lower-dimensional structures.12,13 The stability of the 3D perovskite structure is determined by the tolerance factor first introduced by Goldschmidt for inorganic oxides14 and later modified for organic–inorganic perovskites by introduction of the effective ionic radius reff of the molecular cations.15,16 This semiempirical approach is based on the assumption of rotational freedom of the molecular cations around their center of mass and work surprisingly well for organic–inorganic perovskites, thus confirming the highly dynamical nature of the these cations within the perovskite cavities. The anisotropic size of the molecular cations, together with the number and the type of hydrogen donors for N–H···X hydrogen bonding, are important in shaping the distortions of the ordered or partially ordered low-temperature phases.17 When the tolerance factor falls in the range of 0.9 to 1.0, 3D perovskites of cubic symmetry are formed, while a tolerance factor of 0.80 to 0.89 typically leads to lower symmetry. Low-dimensional structures typically emerge, when the tolerance factor is higher than 1 or lower than 0.7.
The low-dimensional lead halides can crystallize in several stoichiometries and can be regarded as two-, one-, or zero-dimensional (2D, 1D, and 0D) structures. The 2D hybrid organic–inorganic lead halides are derived from 3D perovskites by removing inorganic layers along the (100), (110), or (111) directions, and the most common stoichiometry is A′2An–1PbnX3n+1 or A′′An–1PbnX3n+1, where A, A′, and A″ are a small monovalent, large monovalent and large divalent cation, respectively.18 1D structures have various stoichiometries, and they contain chains composed of corner-, edge-, or face-shared PbX6 octahedra separated by organic cations.19 0D structures also have various stoichiometries, for instance, A4PbX6, and they consist of completely isolated PbX6 octahedra or lead halide clusters.20,21 In most cases, the isolation is achieved by employment of large organic cations. As discussed by Akkerman and Manna, these types of lead halides, especially 1D and 0D, have little in common with the 3D perovskite structure, and therefore they should not be identified as perovskites.22 However, the term “perovskite” for these low-dimensional structures has been widely accepted in the literature, and in this review we will call such structures perovskites.
The 3D structures of lead halide perovskites have a typical APbX3 topology, where the A-site cation is enclosed in an inorganic lead-halide cage formed out of the corner-sharing PbX6 octahedra (Figure 1a). Currently, there are five known A-site cations (methylammonium (MA), formamidinium (FA), methylhydrazinium (MHy), aziridinium (AZR), and Cs+), which stabilize the 3D perovskite structure (Table 1). Among these, hybrid organic–inorganic MAPbI3 and FAPbI3, and all-inorganic CsPbX3 perovskites are the most promising for solar cell and related applications.8,9
Figure 1.

Mixing in the (a) 3D and (b) lower-dimensional (Ruddlesden–Popper) lead halide perovskites.
Table 1. The Most Popular A-site Cations Used for Mixing and Formation of 3D Lead Halide Perovskitesa.
| A-site cation | reff (pm) | tolerance factor | dipole moment (D) |
|---|---|---|---|
| Methylammonium (MA) CH3NH3+ | 217 | 0.91 | 2.26 |
| Hydrazinium (HY) H3N-NH2+ | 217 | 0.91 | 3.24 |
| Formamidinium (FA) HC(NH2)2+ | 253 | 0.99 | 0.22 |
| Imidazolium (IM) C3N2H5+ | 258 | 1.00 | 1.42 |
| Methylhydrazinium (MHy) CH3NH2NH2+ | 264 | 1.01 | 2.80 |
| Dimethylammonium (DMA) (CH3)2NH2+ | 272 | 1.03 | 1.52 |
| Ethylammonium (EA) (C2H5)NH3+ | 274 | 1.03 | 3.95 |
| Acetamidinium (ACE) H3C(NH2)2+ | 277 | 1.04 | 1.33 |
| Guanidinium (GA) C(NH2)3+ | 278 | 1.04 | 0 |
| Cs+ | 167 | 0.81 | - |
| Rb+ | 152 | 0.78 | - |
The effective cation radius reff and the tolerance factor (in the lead iodide framework) were calculated by following the procedure described in ref (15). The electric dipole moment was calculated for the isolated cation using DFT (B3LYP/6-31G*).
Mixing in 3D perovskites can be performed in all three sites of the lattice (Figure 1a),10,11 although the B-site (metal) mixing is significantly less frequent. The choice of organic and inorganic cations for the A-site mixing is rather broad with the most popular cations listed in Table 1 together with their corresponding effective ionic radii, tolerance factors, and dipole moments. The mixing ingredients at the X-site are limited to I–, Br–, and Cl– anions, which have different ionic radii and affinities to form H-bonds with the molecular cations.
Upon mixing with high amounts of large molecular cations, the 3D perovskite structure becomes unstable due to the increase of the tolerance factor.15,16 In this limit, typically a phase separation occurs in the 3D perovskite and low-dimensional perovskite-like structures. Currently, the low-dimensional perovskites are widely applied in fabrication of solar cells,23−25 LEDs,4−7,19,25−28 photodetectors,26,29 piezoelectric sensors,26,30 scintillators,19,31 nonlinear optical (NLO) switches,32,33 and dielectric switches.34,35 These compounds occur in various topologies comprising one or two organic cations, where examples of the former compounds are A2′PbX4 and A″PbX4 layered perovskites (A′ = monovalent cation, A″ = divalent cation).36−41 There are many reports on the formation of lead halide perovskites comprising two organic cations. One of these groups constitutes an alternating cation in the interlayer space (ACI) perovskites with the general formula A′A′′nPbnX3n+1 (n ≥ 1).42−44 Another and very large group constitutes quasi-layered Dion-Jacobson (DJ, A′′An–1PbnX3n+1) and Ruddlesden–Popper (RP, A′2An–1PbnX3n+1) phases with n ≥ 2, where A, A′, and A″ is a small “perovskitizer” cation, large spacer monovalent cation. and large spacer divalent cation, respectively (Figure 1b).23,45−47 Most of the small cations are the same as those used in the synthesis of pure and mixed-cation 3D perovskites, e.g., MA, FA, DMA, GA, etc., but the relaxed tolerance factor in such low-dimensional perovskites also allows the incorporation of larger cations such as isopropylammonium (i-PA) and 1,2,4-triazolium (TZ) into the perovskite cavities.48,49
There are numerous studies and reviews on the mixing effects on the photovoltaic performance, optical properties, and stability of lead halide perovskites (see e.g. refs (11, 50−52)). However, despite these properties highly depending on the atomistic picture of the mixed compositions,53−57 a systematic understanding of how mixing affects the structural and dynamics phenomena in these materials is still lacking. This is especially important in the context of classical inorganic perovskites, where mixing is known to drastically alter the long-range order and lattice dynamics, causing formation of exotic frustrated phases such as relaxors and electric dipole glasses.58−62
The main goal of this review is to provide the first comprehensive and systematic understanding and highlight common aspects of mixing effects on the structural and dynamic properties of lead halide perovskites. The introductory section of the review is structured to briefly cover the most important photovoltaic aspects of mixing followed by a brief summary of ion substitution effects in classical inorganic compounds and the most common tools employed to study the structural and dynamic properties of mixed lead halide perovskites. The subsequent sections provide an in-depth analysis of the structural phase transitions in nonmixed parent compounds forming a basis for the subsequent chapters, which deal with the A-, X-, and B-site mixing in 3D perovskites. Afterward, the mixing effects in the low-dimensional lead halide systems are discussed. The review ends with the generalization of the common aspects observed in different mixed systems followed by an outlook, which draws a roadmap for further studies of these intricate systems.
1.1. Benefits of Mixing in Lead Halide Perovskites and Related Compounds
It is well-known that one of the main limiting factors of the most popular 3D perovskites MAPbI3 and FAPbI3 is their low chemical stability in a humid environment and under exposure to light as well as low thermal stability. One of the main strategies to increase stability is the A-site mixing with larger organic cations like GA, DMA, IM, EA, HY, ACE, hydroxyethylammonium (HEA), or thioethylammonium (TEA)63−74 (see Table 1 for the most popular A-site cations). For instance, mixing of MAPbI3 with 5% of GA led to a 62% decrease in the degradation rate under illumination.63 Another very important benefit of mixing at the A-sites is a better photovoltaic,63−67,69−71,75−80 photodetecting,68,73,74,81 and photoluminescence (PL) performance63 as well as tunability of the NLO properties.82 For example, mixing of MAPbI3 with GA increased PL lifetimes and enhanced efficiency of solar cell devices.63
There are numerous experimental and theoretical reports showing that the A-site mixing affects the electronic structure including band gap and electron–phonon coupling.69,80,81,83−87 Studies of MA1–xCsxPbI3 and FA1–xCsxPbI3 showed that the band gap gradually increased with increasing Cs content, and this effect was attributed to shortening of the Pb–I bond lengths, when the larger MA cation is replaced by the smaller Cs+.83,87 Mixing of MAPbX3 compounds with larger cations also revealed a gradual increase of the band gap with increasing level of mixing. For instance, in the MA1–xDMAxPbBr3 solid solution, the band gap showed a change from 2.17 eV for x = 0 to 2.23 eV for x = 0.3,86 while in the MA1–xGAxPbI3 system the change was from 1.49 eV (x = 0) to 1.53 eV x = 0.22.80 This behavior was attributed to local distortions of the inorganic sublattice, especially alteration of the X–Pb–X bond angles mediated by the new H-bonds.69,84 Computations performed for the MA1–xDMAxPbI3 and MA1–xGAxPbI3 solid solutions suggested that the A-site mixing effectively suppressed the strength of the electron–phonon coupling and that this effect is stronger for the GA-based system.81 It was suggested that the reduced strength of the electron–phonon coupling may also contribute to the longer carrier lifetimes in the mixed perovskite.81 Latter temperature-dependent studies of the PL properties of MA1–xFAxPbI3 solid solutions showed that, while the effective phonon energy is almost independent of the FA content with an average value of about 5 meV, the electron–phonon coupling strength adopts two different values: about 10 meV for the end members and about 17 meV for the intermediate compositions.88 Thus, for this system the mixing led to the increase of the electron–phonon coupling.
Improved photovoltaic performance, stability, and tunability of PL color and intensity can also be realized by doping at the B-sites with Sn2+71,89−97 or other metal cations.98−101 The improved properties after mixing at the B-site have been attributed to change of the band gap, decrease of the electron–phonon coupling strength, increase in the material defect formation energy, and increased charge carrier lifetimes due to changes in the cation dynamics.67,84,89,91,93−97,102 It is worth adding that band gaps of tin-based perovskites are narrower compared to lead-based analogues. Consequently, for many B-site mixed perovskites comprising divalent lead and tin, for instance CsPb1–xSnxBr3 and FA0.75Cs0.25Pb1–xSnxI3, the band gap exhibits gradual narrowing with increasing Sn content.103,104 Interestingly, in some cases, the mixed lead–tin perovskites possess narrower band gaps compared to the end members, making these mixed system very attractive for photovoltaic applications. This behavior (band gap bowing) was reported for MAPb1–xSnxI3 and FAPb1–xSnxI3 solid solutions, for which the intermediate compounds showed corresponding band gaps of 1.17 eV and near 1.3 eV, whereas the band gap of MAPbI3, MASnI3, FAPbI3, and FASnI3 was 1.55, 1.30, 1.55, and 1.39 eV, respectively.89,94 It was argued that this anomalous behavior emerges from the nonlinear mixing of Pb and Sn orbitals in the band edges.105 In addition to tuning of the optoelectronic properties, the B-site doping is also highly desirable due to high lead toxicity.94,98
It is important to note that, although the photoactive black phases of FAPbI3 and CsPbI3 have great application potential, they are thermodynamically unstable and transform into the nonphotoactive yellow phases in an ambient humid atmosphere.106,107 The black phases can be stabilized by preparing FA/Cs, FA/MA, or FA/Rb solid solutions,76,108−110 by mixing FAPbI3 with large organic cations,72,79 or by mixing both the A- and X-sites.95,111−115 In the latter case, Br– anions are usually used at the X-sites.111−113 Mixing at this site is also widely used to tune optical band gaps and improve photovoltaic properties.9,116−118 The effect of the halide exchange on the band gap is much more pronounced compared to the A- or B-site mixing. In general, the band gap in Br/Cl and Br/I systems strongly narrows with increasing Br and I content, respectively, due to the increased Pb–X bond length.85,119 As a result, the PL color is altered from UV for chlorides to near-IR for iodides.9,116−119 Mixing at the X-site also tunes the NLO and dielectric properties.118,120,121
Another very important aspect of mixing in MAPbI3 is the lattice symmetrization and cubic phase stabilization achieved by suppression of the structural phase transitions. In addition to lower defect density, longer carrier lifetime, larger wide absorbance range, and enhanced stability,68,78,122 stabilization of the cubic phase below room temperature is desirable, as it prevents the phase-transition-induced accumulation of the lattice fatigue and strain, which are harmful for device operation.123
In the case of lower-dimensional perovskites, a partial mixing was reported only at the B- and X-sites and mainly for the A′2PbX4 and A″PbX4 layered perovskites. Similarly to the 3D analogues, the mixing at the X-sites strongly modifies the band gap, and therefore this mixing was chosen as a strategy to tune the PL color and to obtain materials exhibiting white-light emission.36−41 The X-site mixing may also lead to tunable dielectric properties.124 Mixing at the B-site is less common, but it was shown that substitution of lead with tin narrows the band gap and induces broadband PL.125−127 Replacing of lead with other metal cations like Mn2+ or Sb3+ is also a way to tailor the color and PL quantum yield.128,129 Although partial mixing at the A′- and A″-sites of A′2PbX4 and A″PbX4 perovskites (RP and DJ phases with n = 1) has not been reported, there are many reports on the formation of lead halide perovskites comprising two organic cations. For the ACI perovskites, employment of two different organic cations offers a potential way to tailor the spin-based and PL properties.42−44 Employment of two organic cations in the DJ and RP phases has many benefits. First of all, it allows controlling the optoelectronic and PL properties,23,46,47 but it may also induce other functional responses like ferroelectricity,130−132 antiferroelectricity,133 pyroelectricity,134 or NLO properties.133,135 Properties of DJ and RP phases with n ≥ 2 can be further fine-tuned by mixing at the A′-,136,137 A″-,138,139 A-,138,140,141 and B-sites.142,143 Although many works reported preparation of these phases with different halide anions,144 recent studies showed that this is not a real mixing since halide anions preferentially occupy different sites.145
Despite a vast number of works concentrating on the exploitation of the discussed improvements for device fabrication, mixing effects on the structural and dynamic properties of lead halide perovskites are significantly less studied and understood. It is reasonable to assume that mixing in these compounds would result in a similar behavior as observed in classical inorganic solid solutions, where the substitution effects are significantly better comprehended.
1.2. Mixing Effects in Classical Inorganic Compounds
It is well established that compositional disorder may significantly change the structural and dynamic properties of classical inorganic ferroelectrics and related materials.58−62 A ferroelectric compound exhibits spontaneous electric polarization, which can be flipped (switched) by the external electric field, as evident from a ferroelectric hysteresis loop measurements (see inset in Figure 2a). Typically, these compounds also exhibit structural phase transitions from the paraelectric (disordered in order–disorder ferroelectrics) to the ferroelectric (ordered) state with changing temperature or pressure.146
Figure 2.
Temperature dependence of the complex dielectric permittivity of (a) PbTiO3 perovskite ferroelectric, (b) canonical relaxor PbMg1/3Nb2/3O3, and (c) dipolar glass (betaine–phosphate)0.15(betaine–phosphite)0.85. A typical ferroelectric hysteresis loop is presented in the inset in (a). Inset in (b) shows a classical example of dipole frustration in the antiferrommagnetic Ising model on a 2D triangular lattice. (a) Adapted with permission from ref (154). Copyright 1971 Taylor & Francis. (b) Adapted with permission from ref (62). Copyright 2006 Springer Nature. (c) Adapted with permission from ref (153). Copyright 1996 IOP Publishing.
In some of ferroelectric and related compounds, ion substitution introduces competing interactions, which cause a complete suppression of the structural phase transitions and frustration of electric dipoles or orientational degrees of freedom (inset in Figure 2b). As a result, the long-range dipole order is replaced by the frustrated phases such as dipolar or orientational glass (e.g., (KBr)1–x(KCN)x),59−61,147,148 strain glass (e.g., Ti50–xNi50+x),149 or relaxor (e.g., PbMg1/3Nb2/3O3).58,62,150,151 The dipolar glass phase does not exhibit structural phase transitions and instead is characterized by freezing of electric dipoles at low temperature into structures, which lack long-range order.59 The relaxors also do not show structural phase transitions, though they are more related to ferroelectrics, as in relaxor compounds, the electric polarization may be induced by application of a sufficiently strong external electric field.62 The relaxor phenomena occurs due to random charge or bond disorder in the system, which also leads to formation of polar nanoregions, which grow and freeze with decreasing temperature, as confirmed by neutron diffraction.152
The frustration of electric dipoles in mixed compounds can significantly alter their dielectric response, which is typically described by the complex dielectric permittivity ε* = ε′ – iε″. Here, the real part ε′ corresponds to the dielectric permittivity (dielectric constant), while the dielectric losses (dissipation) are described by the imaginary part ε″. The dielectric response associated with dipolar glass and relaxor phases is usually very broad with respect to both probing frequency and temperature,58,60 which is caused by a broad distribution of the dipolar relaxation times. This behavior is in sharp contrast to compounds exhibiting structural phase transitions, which are typically accompanied by sharp anomalies in the material properties (Figure 2a). The temperature dependence of ε* for a canonical relaxor PbMg1/3Nb2/3O3 (PMN) is presented in Figure 2b, showing a highly frequency-dependent peak in both real ε′ and imaginary ε″ parts of the dielectric permittivity.62 As the frequency decreases, the ε′ peak becomes higher, and its position shifts toward a lower temperature. This dispersion of the dielectric permittivity covers a very wide range of frequency from mHz to GHz.58 A dielectric response of a typical dipolar glass (betaine–phosphate)0.15(betaine–phosphite)0.85 (BP0.15PBI0.85) is presented in Figure 2c also exhibiting a broad dispersion of ε* with a less pronounced maximum of ε′.153 Note that the dielectric responses of ferroelectrics, relaxors, and dipolar glasses exhibit the Curie–Weiss law (ε′ ∼ 1/T) at high temperatures away from the anomalies.62,153 Hence, to distinguish these compounds, the temperature-dependent dielectric measurements in a wide range of temperature and probing frequency are essential.
Another signature of dipolar glass and relaxor phases is divergence of the dipolar relaxation time due to freezing of electric dipoles or dipolar clusters. In such a case, the dipolar dynamics are described by the Vogel–Fulcher law60,62,155,156 instead of the typical Arrhenius behavior. In addition, due to the frustrated nature, these phases exhibit higher entropy at low temperature, which results in excess heat capacity Cp best visible as a maximum in the Cp/T3 representation.148,157 In analogy with the classical inorganic compounds, similar substitution effects can be expected in the mixed lead halide perovskites and related compounds.
In addition to the fundamental interest, the frustrated phases of classical inorganic compounds have a wide range of applications.58 For example, due to large value, slow variation with temperature, and low-loss dielectric permittivity, relaxors are used as dielectric media in ceramic capacitors and tunable microwave devices.158 Some solid solutions of relaxors with ferroelectrics also exhibit a very large electromechanical coupling and piezoelectric response making them widely applicable in piezoelectric devices in transducers and actuators.159 Some relaxor materials also exhibit a large electrocaloric effect that is promising for novel cooling applications.160
1.3. Methods to Study Mixing Effects in Lead Halide Perovskites
Here, we briefly discuss the main techniques used to study how mixing affects the structural properties and dynamics of lead halide perovskites and related compounds. Some of these techniques are summarized in detail in our recent review on molecular spectroscopy of hybrid perovskites.161
1.3.1. Diffraction Techniques
The conventional diffraction-based methods provide information on the long-range crystal structure, evolution of the crystal symmetry, and lattice parameters upon temperature change and mixing and thus are essential for construction of the concentration–temperature phase diagrams (see e.g. refs (162−166)). The presence of complex twinning in the low-temperature phases of lead halide perovskites makes it challenging to analyze partially or completely overlapped patterns obtained from single-crystal X-ray diffraction (SCXRD) methods, resulting in a multidomain diffraction. Furthermore, both single-crystal and powder XRD (PXRD) intensities are primarily dominated by the anionic sublattice due to the large scattering factors of lead and halide atoms in comparison to carbon and nitrogen, which makes it difficult to determine the positions of organic cations. Moreover, even in the case of pure lead halide perovskites, both methods frequently lack the requisite sensitivity to identify subtle changes in crystal symmetry, leading to a narrow separation of the diffracted peaks. The introduction of mixing introduces an additional layer of complexity. Therefore, significantly more advanced diffraction techniques that rely on highly monochromatic, bright synchrotron, and neutron radiation sources are often utilized to obtain the structural information.74,167−172
In general, disordered systems pose a significant challenge to crystal structure solution.173,174 In mixed organic–inorganic materials, frustration can affect either the molecular or inorganic sublattices or both. The organic components in lead halide hybrids are anisotropic and display strong temperature-dependent dynamics. Additionally, they tend to be anchored in the inorganic framework through H-bond interactions at low temperatures experiencing off-center displacements. In these instances, the ordering of the molecular part induces distortion in the inorganic component. Similarly to dipolar glasses, the freezing of molecular motions can result in the random distribution of molecular dipole moments. It may also lead to correlated arrangements and formation of ordered nanodomains, which are somewhat similar to the relaxor behavior. Thus, another important aspect of structural analysis of hybrid lead halide perovskites concerns local order and short-range interactions.
The XRD methods that rely on the periodicity of the structure give only the time-averaged positions of molecular cations and inorganic framework over the measured sample volume. The elastic scattering results in sharp Bragg peaks, whereas static or dynamic deviations from the average long-range structure manifest as additional intensities in-between or around the main peaks and are called diffuse scattering. The diffuse scattering may originate both from static local atomic correlations or from dynamic displacements due to the lattice dynamics (thermal diffuse scattering). Insight into short-range interactions and local structural distortions can be probed via total X-ray scattering, which has recently become more viable due to the advancement of high-flux, single-crystal X-ray diffuse scattering, synchrotron wide-angle X-ray total scattering, and neutron diffuse scattering from single crystals and powders.175−177
The importance of the advanced scattering techniques is evident from the recent diffuse scattering experiments probing short-range order in the average high-temperature cubic phase of MAPbI3, where the local structure was found to be of lower symmetry.178 Dynamical instabilities in MAPbI3 were also evidenced by high energy resolution inelastic X-ray (HERIX) scattering,179 while total X-ray scattering revealed dynamic nanodomains of noncentrosymmetric local structure. In CsPbI3 nanoparticles, the combined Debye scattering equation and pair distribution function approach based on total X-ray scattering evidenced the formation of orthorhombic subdomains in the room- and high-temperature phases,180 whereas two-dimensional dynamical correlations in real space arising from the peculiar CsPbBr3 lattice dynamics were found in the high temperature cubic structure of bulk single crystals.181
1.3.2. Calorimetric Techniques
Calorimetric methods such as differential scanning calorimetry (DSC) and heat capacity Cp measurements are frequently employed to detect the thermodynamic anomalies occurring due to the structural phase transitions in pure and mixed lead halide perovskites.164−166,182−186 These experiments also yield the entropy change associated with the phase transition, which is used to characterize the degree of ordering and character (order–disorder, displacive) of the transition. In addition, the low-temperature Cp measurements of the mixed compounds may provide information on the higher entropy states originating from the dipolar glass or relaxor phase, as revealed in the Cp/T3 representation.148,157,163,164
1.3.3. Spectroscopic Techniques
Various spectroscopic techniques probing vastly different time scales can be employed to study the dynamics and structural properties of mixed lead halide perovskites.161
The broadband dielectric spectroscopy (DS) is a method of choice to study the dynamics and ordering of electric dipoles, phase transitions, and correlated ordered (e.g., ferroelectric) and disordered (dipolar glass) phases.59,62,148,150 These phenomena are typically imprinted in the measured frequency and temperature responses of the complex dielectric permittivity ε* = ε′ – iε″ of the studied material (see e.g. Figure 2). The strength of the DS stems from the ability to probe various processes, which involve change of electric polarization. These include charge transport and accumulation, polar (e.g., ferroelectric) domain dynamics, molecular rotation, dipole relaxation, lattice (phonon) dynamics, and electronic polarization.187 As these processes occur on vastly different time scales ranging from minutes to fs, the DS experiments typically require measurements using a broad range of probing frequency of the applied electric field. In a typical frequency spectrum for a dipolar relaxation, the real value of ε* exhibits a decrease, while the imaginary part shows a peak, as the probing frequency crosses the frequency associated with the studied dipolar dynamics.
To probe different dipolar processes occurring in lead halide perovskites, ultra broadband dielectric experiments extending to the GHz frequency range may be necessary because of the fast reorientation of organic cations on the ns and ps time scales. This poses significant experimental challenges, as simple capacitor-type measurements are no longer suitable, and instead conceptually different and more complicated coaxial and waveguide setups must be employed.188,189 In addition to the information on the phase transitions and dipole dynamics, the DS provides the dielectric permittivity value, which directly affects the exciton binding energy and consequently the performance of the photovoltaic devices.53,55,56,183,190−192
Raman and infrared (IR) spectroscopy is another versatile tool well-suited for investigation of the mixing effect in lead halide perovskites and related compounds.69,88,118,165,193 These techniques can probe the local molecular environment, degree of disorder, symmetry, molecular vibrations, and lattice phonons,194−199 all of which to some extent can be affected by mixing.69,88,118,165,193,200 IR spectroscopy can also provide information on the frequency-dependent dielectric permittivity, electron–phonon coupling strength, and polaron masses.196,201 The temperature-dependent Raman and IR measurements of the molecular cation and lattice modes typically are highly sensitive to the structural phase transitions,88,165,185,194−199,202 complementing other techniques for construction of the phase diagrams.
Nuclear magnetic resonance (NMR) is another well-established spectroscopic technique used to probe the dynamics and phase transition in (mixed) lead halide perovskites and related compounds.75,102,172,203−206 It is also frequently employed to determine the stoichiometry and solubility limits of the mixed compounds.166,207,208 The basis of this method relies on the nuclear Zeeman interaction, which results in a splitting of the nuclear spin energy levels in the external magnetic field. The splitting is also affected by the local nuclear spin environment, which may reflect the local symmetry and degree of disorder introduced by mixing. Information about the molecular cation dynamics on a time scale ranging from ps to hours can be obtained from the spectra of the quadrupolar nuclei and relaxation time measurements.102,204,209,210 A related technique used to study mixing effects in lead halide perovskites is a nuclear quadrupole resonance, which is performed in the absence of the external magnetic field.172,203 These experiments probe the interaction of the electric field gradient with the nuclear quadrupole moment, which is very sensitivity to the changes of the local symmetry and environment of the studied nucleus, providing information on the local dynamics and structural phase transitions.
Molecular dynamics can be also studied using a quasielastic neutron scattering (QENS) technique. QENS is a limiting case of inelastic neutron scattering, in which broadening of the elastic line is due to a small energy transfer between the neutron and the atoms in the sample. QENS probes the energetically broadened signal around the elastic line and measures energy transfer down to the sub-μeV regime. This allows access to low energy collective motions, diffusional motions, relaxation processes, and molecular reorientations.211,212 Thus, QENS is often used to provide information about the dynamics (vibrational, rotational, and translational diffusive processes at the molecular scale) on time scales from ps to μs and on length scales from 1 to 500 Å.211−213 This technique finds application in many scientific topics, including polymers, dynamics of proteins, hydrogen dynamics, etc.211,212 QENS proved to be a very powerful method to determine the modes, characteristic correlation times, and activation energies of the molecular cation motion in lead halide perovskites.213−217
1.3.4. Computational Techniques
Computational techniques such as the density functional theory (DFT), Monte Carlo (MC), and molecular dynamics (MD) simulations provide information on the microscopic origins of the molecular cation and lattice dynamics, structural phase transitions, and mixing effects.218 The DFT calculations can yield the atomistic picture of the ordered phases164,166,219−221 but are constrained to small system sizes and poor ability to probe entropic effects. The temperature effects and dynamics can be probed using MD,222−226 but such simulations may also suffer from the pronounced finite-size effects, especially if ab initio MD is used. The MC simulations are capable of simulating large many-particle systems and temperature effects, but usually the atomistic picture must be simplified to a coarse-grained model based on the effective model Hamiltonian formalism.164,213,227−229 Recently, significant attention was also concentrated on the machine-learning-augmented calculations, which are capable of predicting new structures, phase transitions, and properties of (mixed) lead halide perovskites, while overcoming the need for large-scale simulations.205,206,230−237
2. Structural Phase Transitions in Nonmixed 3D Lead Halide Perovskites
Here, we discuss the structural phase transitions and phase symmetries of MAPbX3, FAPbX3, CsPbX3, and MHyPbX3 lead halide perovskites forming 3D structures. The commonly accepted phase diagrams of these compounds are summarized in Figure 3.
Figure 3.

Summary of the structural phase transitions and symmetries of different 3D lead halide perovskites. Color coding: green - cubic, pink - tetragonal, blue - orthorhombic, yellow - trigonal, violet - monoclinic, gray - unknown.
2.1. MAPbI3
MAPbI3 exhibits two structural phase transitions
at 327 and 162 K associated with the symmetry lowering from the cubic
phase to tetragonal and orthorhombic symmetries.
The former transition is similar to inorganic SrTiO3238 and is being driven by out of phase octahedral
rotation around the cubic c-axis (Glazer notation:239a0a0c–) followed by a
partial MA cation ordering and transformation of the unit cell to
body centered b – a, a + b, 2c superstructure. The second phase
transition results in a complete ordering and orthorhombic phase with
more complex tilting and distortion of the PbI6 octahedra.
Consensus has been reached in the literature regarding the change
of the crystallographic system of MAPbI3, yet the assignment
of the tetragonal and orthorhombic space group remains a topic of
debate. Initially, the orthorhombic phase was modeled in the Pna21 polar space group based on low-temperature
Guinier data.240 However, more recent powder
neutron diffraction experiments have led to its reclassification to
centrosymmetric Pnma.167,241 Similarly,
the tetragonal phase was originally discovered in the polar I4cm space group93 and later recalculated in the centrosymmetric I4/mcm space group.242,243 The commonly
accepted centrosymmetric models of the room-temperature tetragonal
and low-temperature orthorhombic phases have been called into question
by numerous studies claiming ferroelectric behavior of MAPbI3,53,244−248 although no solid proof of ferroelectricity
(and thus absence of centrosymmetry) was observed. Contrary, many
studies claim nonferroelectric ordering in this compound,183,227,249−252 supporting the centrosymmetric symmetry.
Several studies also revealed the presence of octahedral rotational instabilities in the cubic phase of MAPbI3, which correlate with the MA cation motion and cause a local dynamic symmetry breaking.178,179,253−258 Beecher et al.179 estimated that the phonon lifetimes associated with such instabilities is on the ps time scale, and the size of such correlated domains is a few nm. Recently, Weadock et al.178 observed that such local structures are two-dimensional circular regions of dynamically tilted PbI6 octahedra that induce longer-range intermolecular MA correlations. Such lattice fluctuations cannot be observed using standard XRD techniques, which are sensitive to long-range order and thus provide an averaged-out cubic symmetry. The presence of such low-symmetry correlated octahedral distortions on the nanoscale may be related to the observed transient ferroelectric-like effects in this compound.245−248 Note, however, that such a dynamic picture of the tetragonal domains in the cubic phase of MAPbI3 was questioned in a high-resolution neutron inelastic scattering study, which instead found that the phonon scattering is consistent with the central peak phenomenon observed in classical inorganic perovskites.259
2.2. MAPbBr3
The bromide analogue MAPbBr3 exhibits an even more intricate phase situation than MAPbI3. Poglitsch and Weber240 reported four different phases, starting from the cubic Pm3̅m, tetragonal, I4/mcm, at 234 K, tetragonal, P4/mmm, at 155 K and finally orthorhombic with polar Pna21 space group below 145 K. This sequence was further confirmed by the heat capacity studies.182 Other studies have suggested an incommensurate intermediate phase between the tetragonal and orthorhombic structures,260 or that only two phase transitions occur, with a phase sequence of Pm3̅m - I4/mcm - Pnma analogous to MAPbI3.168 The assignment of the centrosymmetric space groups is supported by the absence of a clear proof of ferroelectricity in this compound.
2.3. MAPbCl3
Hybrid MAPbCl3 also displays a diverse range of phase transitions. At room temperature, as iodide and bromide analogues, it adopts a simple cubic Pm3̅m structure, which transforms into a tetragonally distorted unit cell of P4/mmm symmetry, which exists within a narrow temperature range between 173 and 179 K. Below 173 K, the orthorhombic phase is stabilized. The P2221 space group was proposed for this phase, with a doubled c lattice parameter,240 while a P212121 symmetry was observed in a 2a, 2b, 2c modulated superstructure in ref (261). More recently, the Pnma space group was reported based on the synchrotron radiation data indicating heavily distorted PbCl6 octahedra.262 The low-temperature structure of MAPbCl3 appears to adopt the a0b+c– tilt system.
2.4. FAPbI3
An even more complex behavior of the structural phase transitions
is observed in the FA-based hybrid perovskites. In contrast to MAPbI3, the FAPbI3 system crystallizes in the hexagonal
photoinactive yellow phase (δ-FAPbI3), which can
be transformed into the desirable metastable photoactive black phase
(α-FAPbI3) by thermal annealing above 150◦ C.222,263 When growing from solution media, the black
phase is favored above 60 °C.93 Chen
et al.264 showed that the black to yellow
phase transition is a thermally activated process between the two
structures separated by an energy barrier. As a result, after quenching
from 400 to 200 K over 80 min, FAPbI3 was kinetically trapped
in the black cubic phase and remained in this phase upon further cooling
to low temperatures. Neutron and X-ray powder diffraction experiments265 conducted on both N-deuterated and hydrogenous
α-FAPbI3 revealed that this material exhibits two
phase transitions. At 285 K, the systems undergo a phase transition
from the cubic
to the intermediate tetragonal centrosymmetric
β-phase with a primitive tetragonal b – a, a + b, c unit cell,
characterized by in-phase rotations (a0a0c+) similar
to those found in NaTaO3 at high temperature.266 Below 140 K, a low-temperature tetragonal γ-phase
of polar P4bm symmetry was claimed.
SCXRD confirmed the P4/mbm tetragonal
symmetry of the β-phase at 200 K.267 The γ-phase was also reported to have the tetragonal P4/mbm space group similar to the β-phase—on
the grounds of the high-resolution synchrotron diffraction data, testifying
the isostructural phase transition from the β- to γ-phase.268 Recent neutron powder diffraction experiments222 have indicated that the γ-phase is locally
disordered and that there is no long-range ordering of the FA cations.
The disordered γ-phase is characterized by the geometrical frustration
of the molecule-cage interaction, which seems to cause a glassy behavior
at low temperatures.209
2.5. FAPbBr3
The bromide analogue FAPbBr3 exhibits less complex phase behavior with the high-temperature cubic Pm3̅m (above 262 K), intermediate tetragonal P4/mbm and low-temperature orthorhombic Pnma structure (below 153 K) determined through combined neutron powder diffraction and synchrotron X-ray diffraction.269 These phase transitions are associated with partial and complete ordering of FA cations, which introduce lattice deformations. In addition, three additional isosymmetric phase transitions, which are not resolved crystallographically, were also observed in this system at 182, 162, and 118 K.165,270 According to QENS and Raman scattering, the isosymmetric phase transitions relate to changes in the FA cation dynamics, which may affect the optoelectronic properties of this material.270,271 Note that isosymmetric transitions are also observed in CsPbBr3191 and GAPbI3,272 while in MAPbBr3 it can be induced by applied pressure.273
2.6. FAPbCl3
The FAPbCl3 analogue has two first-order phase
transitions at about 258 and 271 K. The first transition is associated
with the cubic
to tetragonal symmetry lowering, while
the low temperature phase was solved in the orthorhombic centrosymmetric Cmcm symmetry based on SCXRD data.274 The recent neutron diffraction studies show that in the low-temperature
orthorhombic phase, FA cations undergo a 2-fold jump reorientation
about the C–H axis, which changes to an isotropic rotation
in the higher temperature tetragonal and cubic phases.275
2.7. CsPbI3
The first reports on CsPbI3 crystal structure
appeared in the late fifties;276 however,
the complete picture on the temperature phase behavior of CsPbI3 has been documented quite recently by synchrotron powder
diffraction.277,278 The results are consistent with
the formation of two low-temperature phases related to the high-temperature
cubic aristo phase of α-CsPbI3
. The first transition to a primitive tetragonal
cell of P4/mbm occurs at about 554
K. On further cooling below 457 K, the orthorhombic CsPbI3 phase of Pnma symmetry is stabilized. The crystal
structure of CsPbI3 has been also reported in refs (180, 279−281). It is noteworthy that the photoactive black phase of CsPbI3 is formed at high temperature through the reconstructive
phase transition from a nonperovskite yellow δ-phase, which
takes place slightly above 300 °C.276 The reconversion of the black phase to the initial yellow phase
is catalyzed by humidity, and therefore, when the black phase is not
exposed to humidity, it remains stable at room temperature for a matter
of weeks, while in a humid environment, it transforms to the yellow
phase within minutes.107
Similarly to hybrid lead halide perovskites, multiple studies have found that inorganic CsPbX3 perovskites also exhibit short-ranged dynamic fluctuations of the PbI6 octahedra showing that this behavior is intrinsic to the general lead-halide perovskite structure and not unique to the dipolar organic cation.181,257,258,282 Lanigan-Atkins et al.181 found that such correlated dynamics occur over a ps time scale and involve liquid-like damping of low-energy Br-dominated phonons, which correspond to 2D sheets of correlated octahedral rotations. In addition, a recent study283 revealed that the transformation from the yellow to the black phase of CsPbI3 is driven by the harmonic phonon entropy contribution to the Gibbs free energy, which is substantially higher in the softer black phase of CsPbI3 compared to the stiffer yellow phase. The difference in the vibrational entropy arises from the differently stacked configurations of PbI6 octahedra in both phases, which permit larger amplitudes of atomic vibrations in the black phase. Note that similar entropy arguments were used to explain the yellow to black phase transformation in the hybrid FAPbI3 system.264
2.8. CsPbBr3
A similar phase transition pathway is present in CsPbBr3, where the cubic aristo phase initially distorts to the tetragonal phase at 403 K, followed by a further symmetry decrease to the orthorhombic phase at 361 K.276,284−288
2.9. CsPbCl3
CsPbCl3 also has the cubic perovskite structure above 320 K289 and undergoes three successive phase transitions in a narrow, 10 K, temperature range.290 The prototypic Pm3̅m phase I transforms at 320 K to the tetragonal phase II (P4/mbm) and further at 315 and 311 K to the orthorhombic phases III (Cmcm) and IV (Pnma).291−293 The crystal structure of the Pnma phase has been confirmed among others by synchrotron powder diffraction288,294 and SCXRD at 200 MPa.295 It is worth adding that, contrary to CsPbCl3, the corresponding tin analogue CsSnCl3 can be trapped in the high-temperature cubic phase at room temperature by heating the sample to 100 °C and avoiding exposure to humidity.296,297
2.10. MHyPbX3
In contrast to lead halides templated by MA and FA cations,
perovskites based on MHy crystallize into well-ordered, low-symmetry
structures. MHyPbBr3, for instance, adopts the polar monoclinic P21 space group, which transforms to a disordered
cubic structure
above 418 K.185 Similarly, the chloride analogue crystallizes in the same monoclinic P21 symmetry, which evolves to the orthorhombic
polar structure, Pb21m, at 342 K.184 Notably, the high-temperature
phase is still ordered, and the contribution to spontaneous polarization
from MHy dipoles is even larger than at room temperature, leading
to stronger second-harmonic generation (SHG) in the high-temperature
phase. It is worth highlighting that the MHy-based 3D perovskite structures
have a tolerance factor greater than 1. This unique feature results
from a significant distortion of the PbX6 octahedra by
MHy, which enters the coordination sphere of lead and forms Pb–N
covalent coordinate bonds. Note, however, that despite possessing
a similar tolerance factor, no one has succeeded in growing a 3D MHyPbI3 analogue to date.
2.11. AZRPbX3
Recently, the family of 3D hybrid perovskites expanded, as new members based on cyclic aziridinium cations were synthesized.298 Similarly to MA and FA analogues, it was found that AZRPbX3 compounds exhibit a cubic Pm3̅m symmetry at room temperature. A subsequent phase transition study reported by us237 revealed that AZRPbI3 exhibits a single structural phase transition at 132 K to orthorhombic Pnma symmetry similar to the orthorhombic polymorph of MAPbI3. A single phase transition was also observed for AZRPbBr3 at 145 K, but instead of the orthorhombic symmetry, an unusual trigonal R3̅c space group was obtained. The chloride analogue was found to exhibit two closely spaced phase transitions at 162 and 154 K, but the low-temperature symmetries were not determined due to complex twinning of the crystal structure.
3. Structural Phase Transitions in Lead Halide Perovskites Comprising Two Cations
Here, we briefly discuss the structural phase transitions in pure RP, DJ, and ACI lead halides, which contain only a single cation at the A-, A′-, and A″-sites. In contrast to 3D APbX3 compounds, many of these structures exhibit clear ferroelectric or antiferroelectric ordering evident by the appearance of spontaneous electric polarization.
3.1. Ruddlesden–Popper Phases (n ≥ 2)
Temperature-dependent structural phase transitions were reported for many RP lead halide phases of the general formula A′2An-1PbnX3n+1. Here, A′ denotes a spacer cation such benzylammonium (BZA), isobutylammonium (i-BA), EA, amylammonium (AM), isoamylammonium (i-AM), propylammonium (PA), isopropylammonium (i-PA), 3-bromopropylammonium (BPA), 3-chloropropylammonium (CPA), 4-aminomethyl-1-cyclohexanecarboxylate (t-ACH), allylammonium (AA), cyclohexanemethylammonium (MACH), 4-iodobutylammonium (4IBA), and CH3(CH2)nNH3+ (n = 6, 7, 8, 11, 13, 15, 17), while A is a “perovskitizer” cation such as MA, Cs, FA, EA, DMA, GA, MHy, and FA (Figure 1b).47,130,133,134,299−302,302−330 In contrast to 3D APbX3 lead halides,183,227 many of the discovered RP compounds, especially bromides, exhibit ferroelectric or antiferroelectric ordering.130,133,134,300,301,303−316,320−326,328−331Table 2 lists RP lead halides, for which the ferroelectric or antiferroelectric phase was confirmed by observation of hysteresis loops or a pyrocurrent. The electric dipole order is usually related to ordering of the bulky A′ organic cations301,313,322,326,329,331 or to the synergic effect of both A′ and A cation ordering and distortion of the framework.130,133,300,303−312,314,315,320,321,323−325,328,330 As an illustration of this process, we briefly discuss the ferroelectric phase transition in BA2MAPb2Br7. Li et al. showed that, above 352 K, this compound crystallizes in the centrosymmetric structure (space group Cmca).300 In this phase, the organic MA and BA cations are highly disordered, which causes cancelation of the molecular dipole moments (Figure 4a). When the temperature decreases, the organic cations order, and the crystal symmetry changes to Cmc21. The electric polarization appears along the c-direction due to a specific reorientation of the organic cations (Figure 4b).300
Table 2. List of Ferroelectric (FE) and Antiferroelectric (AFE) RP Phases.
| compound | ordering, transition temperature (K) | symmetry change | Reference |
|---|---|---|---|
| BA2MAPb2Br7 | FE, 352 | Cmc21 → Cmca | (300) |
| BA2MA2Pb3Br10 | FE, 315 | Cmc21 → Cmca | (303) |
| BA2FAPb2Br7 | FE, 322 | Cmc21 → Cmcm | (301) |
| BA2FAPb2I7 | FE, 304 | Fmm2 → I4/mmm | (314) |
| BA2CsPb2Br7 | FE, 412 | Cmc21 → Cmca | (331) |
| BA2EA2Pb3Br10 | FE, 380 | Cmc21 → I4/mmm | (305) |
| BA2EA2Pb3I10 | AFE, 363; FE, 322 | Cmc21 → Pbca → I4/mmm | (133), (134) |
| i-BA2CsPb2Br7 | AFE, 353 | Pmnb → Cmca | (304) |
| i-BA2EAPb2Br7 | FE, 326 | Cc → I4/mmm | (313) |
| i-BA2EA2Pb3Br10 | FE, 370 | Cmc21 → I4/mmm | (313) |
| i-BA2FAPb2Br7 | AFE, 303.2 | Pnma → I4/mmm | (315) |
| i-BA2MAPb2Cl7 | FE, 316 | Fmm2 → I4/mmm | (320) |
| EA4Pb3Cl10 | FE, 415 | Cmc21 → I4/mmm | (310) |
| EA4Pb3Br10 | FE, 384 | C2cb → I4/mmm | (306) |
| EA2MA2Pb3Br10 | FE, 375 | Cmc21 → I4/mmm | (311) |
| EA2MA2Pb3Cl10 | FE, 390 | Cmc21 → I4/mmm | (321) |
| AM2CsPb2Br7 | FE, 313 and 349 | F222 → Cmc21 → I4/mmm | (322) |
| AM2EA2Pb3I10 | FE, 361 and 398 | Pmc21 → Cmce → I4/mmm | (330) |
| i-AM2CsPb2Br7 | FE, 323 and 349 | Cmc21 → Pbcm → I4/mmm | (329) |
| i-AM2Cs3Pb4Br13 | FE, 351 and 365 | Cmc21 → I4/m → I4/mmm | (326) |
| i-AM2EA2Pb3Br10 | FE, 371 | Cmc21 → I4/mmm | (307) |
| i-AM2EA2Pb3Cl10 | FE, 392 | Fmm2 → I4/mmm | (324) |
| -AM2EA2Pb3I10 | FE, 313; AFE, 340 | Cmc21 → Pmcn → I4/mmm | (309) |
| i-AM2MA2Pb3Br10 | FE, 305 | Cmc21 → I4/mmm | (323) |
| i-AM2MA2Pb3Cl10 | FE, 343 K | Pmc21 → I4/mmm | (325) |
| PA2FAPb2Br7 | FE, 263.3 | Cmc21 → Cmcm | (130) |
| BPA2FAPb2Br7 | FE, 348.5 | Cmc21 → Cmcm | (130) |
| CPA2FAPb2Br7 | FE, 335 K | Cmc21 → Fmmm | (328) |
| t-ACH2EA2Pb3Br10 | FE, 355 and 380 | Pm → Pmmn → Pmmn | (308) |
| AA2EA2Pb3Br10 | FE, 378 | Cmc21 → I4/mmm | (312) |
| C62Pb2Br7 | FE, 248, 306 | ? → Cc → C2/c | (316) |
| C72Pb2Br7 | FE, 260 | ? → Cc | (316) |
| C72Pb2Br7 | FE, 261, 275 | ? → ? → Cc | (316) |
Figure 4.

(a) Low-temperature (ferroelectric) structure of BA2MAPb2Br7 viewed along the b-axis. Red arrow denotes the direction of electric polarization. (b) High-temperature (paraelectric) structure of BA2MAPb2Br7. Reprinted with permission from ref (300). Copyright 2019 American Chemical Society.
3.2. Dion-Jacobson Phases (n ≥ 2)
In contrast to the RP compounds, temperature-dependent structural phase transitions were reported only for two DJ lead halide perovskites: (AMP)(MA)Pb2I7, (where AMP = 4-(aminomethyl)piperidinium), which undergoes two order–disorder phase transitions at 367 and 449 K,131 and (BDA)(EA)Pb3Br10 (BDA = 1,4-butyldiammonium), which also shows the presence of two phase transitions at 363 and 397 K.332 (AMP)(MA)Pb2I7 exhibits ferroelectric properties, and the polar order is preserved up to the decomposition temperature,131 while (BDA)(EA)Pb3Br10 undergoes ferroelectric-ferroelectric and ferroelectric-paraelectric phase transitions associated with the Pmn21 → Ama2 → Cmcm symmetry change.332 Similarly to the RP phases, the electric order in these compounds appears due to the ordering of organic cations.
3.3. ACI Perovskites
In this family of compounds, structural phase transitions were reported for (TR)(GA)PbBr4 (P21/c → P2/c)44 and (IM)(MHy)PbX4 (P21/c → P2/c, X = Br, Cl).333 In both cases, the phase transitions were found to be of the order–disorder type. However, in the case of (TR)(GA)PbBr4, the change of the space group from P2/c to P21/c occurs due to the ordering of the intralayer TR cations, whereas in (IM)(MHy)PbBr4 and (IM)(MHy)PbCl4 both organic cations are disordered in the high-temperature phases and well-ordered at low temperature.
4. A-Site Mixing in 3D Perovskites
We now will start discussing the mixing effects in 3D lead halide perovskites by considering the most popular mixing recipe, which involves cation substitution at the A-site. We discuss how this mixing affects the structural and dynamic properties of these materials. The cubic 3D perovskite structure is predicted to be most stable for the tolerance factor close to 1, while for values significantly lower than 1 a lower symmetry can be expected at room temperature. For instance, MAPbI3 with a tolerance factor of 0.912334 crystallizes in the tetragonal I4/mcm symmetry at room temperature. An effective way to increase the tolerance factor and stabilize the cubic phase is A-site mixing with larger cations. However, the extent of mixing highly depends on the size of the guest cation—a significant mismatch in the cation size results in a solubility limit, beyond which the phase separation typically occurs.
4.1. MA1–xFAxPbX3
Mixing of MA and FA cations is among the most popular recipes to enhance the performance and stability of lead halide perovskites.115,335 Similar radii of the cations (217 and 253 pm, Table 1)15 and the fact that the end members have perovskite architectures result in the 3D perovskite structures for all mixing concentrations of MA1–xFAxPbX3 (no solubility limit) providing an excellent system for a complete mapping of the phase diagram.
4.1.1. MA1–xFAxPbI3
The first work studying mixing effects on the phase behavior in the MA1–xFAxPbI3 system was reported by Weber et al.,162 where the high-temperature phase transition was investigated using SCXRD and PXRD methods. In the low x region, a substantial lowering of the phase transition temperature was observed upon mixing, reaching a minimum value of about 260 K for x = 0.3 (Figure 5a). For higher values of x, the phase transition temperature recovered, as the system evolved from the MA-rich state to the FA-rich composition. The space group of the cubic phase was assigned to Pm3̅m, which is in agreement with parent MAPbI3 and α-FAPbI3 compounds (Figure 3), while the XRD data below the phase transition point were indexed using either doubled cubic or tetragonal unit cell.
Figure 5.
(a–c) Temperature–composition phase diagrams of a mixed MA1–xFAxPbI3 system obtained in different studies. Symmetry notation: C - cubic, LC - large-cell cubic, T - tetragonal, O - orthorhombic. (a) Adapted with permission from ref (162). Copyright 2016 Royal Society of Chemistry. (b) Adapted with permission from ref (336). Copyright 2019 American Chemical Society. (c) Adapted with permission from ref (88). Copyright 2020 American Chemical Society. (d) Comparison of the phase boundaries presented in (a–c) together with the added data points from refs (102) and (337).
A following study from the Sarma group336 reported a detailed XRD characterization and dielectric properties of the MA1–xFAxPbI3 system in the MA-rich region of the phase diagram (x ≤ 0.4) and a broad temperature range (Figure 5b). For very low values of mixing (x < 0.07), the authors observed a typical cubic-tetragonal-orthorhombic symmetry lowering inherited from MAPbI3 perovskite. The mixing also caused the decrease of the phase transition temperatures compared to the MAPbI3 system in agreement with the aforementioned study by Weber et al.162 At the intermediate values of the studied compositions (0.07 < x < 0.2), the authors claim an observation of a cubic super cell (Im3̅ space group) at low temperature instead of the orthorhombic phase, which demonstrated an unusual reentrant cubic-tetragonal-cubic behavior. For x > 0.2, the tetragonal phase also becomes completely suppressed at the expense of the growing boundary of this large-cell cubic phase. Eventually, only the transition between the cubic and large-cell cubic phases is observed. The authors also note that the XRD data indicate narrow phase coexistence regions at the phase boundaries as depicted in Figure 5b.
The whole FA concentration range in the MA1–xFAxPbI3 system was recently explored by Francisco-López et al.88 using the temperature-dependent Raman and PL spectroscopies. The authors relied on the available structural information on MAPbI3 and α-FAPbI3 perovskites to construct a complete phase diagram of the mixed system (Figure 5c). In the FA-rich region (x > 0.5), the authors show the presence of the tetragonal P4/mbm phase inherited from α-FAPbI3. In addition, it is claimed that a tetragonal-tetragonal phase transition exhibits a critical point at x = 0.7, as the evolution of the PL peak shows no discontinuity with decreasing temperature. However, this assignment is debatable, as the sharp transition anomaly is expected to become diffused in the highly mixed region. A comparison with the study by Weber et al.162 reveals an almost identical phase boundary between the cubic and tetragonal phases in the whole range of mixing (summarized in Figure 5d). However, a comparison with the phase diagram provided by the Sarma group336 shows that the phase boundary of the orthorhombic phase is substantially shifted to higher values of x (Figure 5d). This may suggest different incorporation yields of FA cations in both works, although agreement of the phase boundaries in the high temperature region is satisfactory. No conclusions regarding the discrepancy between the phase symmetries can be made, as no diffraction experiments were performed.
The tuning of the phase transition temperature by mixing was also observed in the nanostructures of MA1–xFAxPbI3 using temperature-dependent PL measurements.337 A small disagreement with other studies in transition temperatures can be observed (see Figure 5d), which may originate from a small size of the nanostructures and broadening of the anomalies in the highly mixed region.338
Ahmadi et al. used piezoresponse and Kelvin probe force microscopies and SHG to study the long- and short-range electric dipole and charge dynamics in MA0.15FA0.85PbI3 composition.339 Surprisingly, the authors reported indications of ferroelectric domains in this mixed compound; however, the domain dynamics were found to be suppressed by fast ion motion.
In addition to the structural data, the aforementioned study from the Sarma group336 reported the temperature dependence of the real part of the complex dielectric permittivity of the MA1–xFAxPbI3 (x ≤ 0.4) pellet samples in a narrow frequency range (1 kHz - 1 MHz) (see Figure 6 for selected compositions). As pointed out by the authors, the dielectric response in this mixed system mainly originates from MA cations due to a much lower electric dipole moment of FA (2.3 D vs 0.2 D, Table 1). At a low FA concentration (Figure 6a), the main feature of the dielectric response is a sharp decrease at the tetragonal-orthorhombic phase transition point caused by the MA cation ordering, while the cubic-tetragonal phase transition is poorly visible. This behavior is in a good agreement with a number of previous DS works on pure MAPbX3 perovskites.183,340,341 Interestingly, the authors claim that the anomaly at higher values of x is related to the transition from the cubic to large-cell cubic phase (Figure 6b). The most intriguing feature of the dielectric response is the broad dipolar relaxation at low temperature, which appears for a higher FA content (Figure 6b,c). This relaxation extends from about 120 to 25 K and is assigned by the authors to the dipolar (orientational) glassy state, which indeed has some similarities with the classical dipolar glasses (see Figure 2c), although broadband DS experiments are necessary to make an unambiguous assignment.
Figure 6.
Temperature dependence of the real part of the complex dielectric permittivity ε′ of the mixed MA1–xFAxPbI3 pellet samples for x = (a) 0.05, (b) 0.2, and (c) 0.4. Reprinted with permission from ref (336). Copyright 2019 American Chemical Society.
A subsequent study from the same group used QENS experiments to investigate the dipolar glass phase and molecular cation dynamics of the mixed MA0.875FA0.125PbI3 compound in more detail.217 The temperature-dependent experiments allowed probing the MA and FA cation dynamics in different structural phases including the proposed large-cell cubic phase. The QENS data were best modeled using a 4-fold (C4) reorientation for the MA cation motion in the tetragonal and cubic phases with the activation energy Ea of 83 meV in a good agreement with 82 meV observed for pure MAPbI3. At a lower temperature, where the 3-fold rotation of MA around the C–N axis dominates, the mixed composition displayed a distribution of activation energies in contrast to pure MAPbI3, which showed Ea = 53 meV. This result further supports the appearance of the dipolar glass phase due to different MA cation environments caused by FA incorporation. Interestingly, the analysis of the elastic incoherent structure factor revealed that FA cations are frozen even up to 340 K, indicating they exhibit a much higher activation energy, and thus the A-site dynamics are governed by MA cations. The fraction of the mobile MA cations was found to be slightly higher in the mixed composition compared to pure MAPbI3, which was explained by an increase of the unit cell volume upon introduction of bigger guest cations. A fractional isotropic rotational model was used to estimate the characteristic time scales of the MA rotation at different temperatures, revealing that MA cation motion is faster in the mixed compound (e.g., 84 ps vs 188 ps at 110 K).
Similar observations and signatures of the orientational glass state were presented by Drużbicki et al.,225 where higher mixing compositions were investigated (x = 0.6 and 0.9) by the inelastic neutron scattering experiments assisted by the harmonic lattice dynamics and ab initio MD simulations. In this study, a highly disordered local environment of MA cations was observed, which was assigned to weakening of the H-bonds and framework expansion induced by mixing. On the other hand, an opposite behavior was found for FA cations, as they form stronger H-bonds with the distorted inorganic cage caused by mixing.
Another interesting NMR and state-of-the-art MD study was reported by Grüninger et al.,206 where the authors found indications of a clustering of the MA and FA cations within the mixed x = 0.25, 0.5, and 0.75 compounds. Such a local heterogeneous distribution would inevitably cause a local variation of the perovskite lattice and is likely related to the observation of the glassy phase in these compounds.
The MA and FA cation dynamics in the mixed deuterated MA0.33FA0.67PbI3 system were investigated by Kubicki et al.102 using solid-state 2H and 14N magic angle spinning (MAS) NMR spectroscopy, which allowed a simultaneous probing of both cations. Temperature-dependent measurements of the quadrupolar coupling constants of 2H and 14N nuclei demonstrated that the reorientation of the C–N bond in MA cations at 327 K is not fully isotropic in the mixed compound, which is in contrast to pure MAPbI3 perovskite. The authors also observed close to isotropic motion of FA cations in the cubic phase. As temperature decreased, this motion slowed down and froze, while a substantial rotation of the N–D3 and N–D2 groups around the C–N bonds of FA remained even at 100 K. At room temperature, it was found that FA rotates much faster than MA (12 ps vs 133 ps), which was attributed differences in the H-bond strength between the cations and the framework. The authors also correlated much faster FA rotation to increased charge carrier lifetimes observed in FAPbI3 and mixed MA1–xFAxPbI3 systems compared to pure MAPbI3.102 In addition to the cation dynamics, this study also provided the phase transition temperature to the cubic phase of 285 K in a good agreement with other studies (Figure 5d). It is interesting to compare the reported results with the aforementioned QENS study.217 Both works detected that mixing alters the MA cation motion, while the observations regarding the FA cation dynamics are in sharp contrast, as the Sarma group reported that the rotational dynamics of FA cations are entirely suppressed over the entire temperature range in the x = 0.125 compound. The observed differences may be explained by different studied compositions, suggesting that level of mixing can highly influence the FA cation dynamics.
A recent NMR quadrupolar relaxation study204 from the same group indicated that the FA cation dynamics in the cubic phase of MA0.3FA0.7PbI3 is essentially the same as in FAPbI3 compound and occur on about a 1 ps time scale with activation energy close to 100 meV. On the other hand, the authors reported that the MA cation motion is about 2× faster in the mixed compound compared to MAPbI3 (1.0 ps vs 1.95 ps). Note that the reported cation correlation times are about 1–2 orders of magnitude shorter compared to their previous work,102 which was explained by difficulties in relating the measured 14N line widths to the motionally induced transverse relaxation, whereas quadrupolar relaxation measurements are much more reliable. Interestingly, the authors observed that the activation energy of the MA cation motion was not affected by mixing and remained close to 150 meV. Note that Fabini et al.209 also found very similar MA and FA cation rotation rates in nonmixed MAPbI3 and FAPbI3 compounds. The observation of fast FA cation dynamics in the x = 0.7 compound is in a sharp contrast to the aforementioned QENS results obtained by the Sarma group,217 where frozen FA cations were observed for the x = 0.125 composition. This discrepancy may originate from vastly different fractions of FA cations in the studied compounds, which belong to the opposite sides of the phase diagram.
4.1.2. MA1–xFAxPbBr3
Recently, we used a multitechnique approach (DSC, heat capacity, ultrasonic measurements, SCXRD, broadband DS, Raman spectroscopy) to thoroughly investigate the mixing effects on the structural phase transition, dielectric response, and cation dynamics in the bromide analogue of the mixed MA1–xFAxPbBr3 (0 ≤ x ≤ 1) system.165 A suite of different experimental techniques probing structural phase transitions of single crystal samples allowed us to construct the phase diagram for the whole range of FA concentrations (Figure 7a). The obtained phase diagram shows lowering of the phase transition temperatures upon mixing and thus stabilization of the desirable cubic phase.122 We also observed disappearance of the intermediate tetragonal phase upon mixing in the MA-rich region of the phase diagram. On the other limit of mixing, the three isosymmetric transitions of FAPbBr3172,270 were also fully suppressed even at the lowest studied MA content, which may paradoxically be related to a relief of the FA cation frustration by mixing (see ref (172) for details).
Figure 7.
(a) Temperature–composition phase diagram of the mixed MA1–xFAxPbBr3 system. Symmetry notation: C - cubic, T - tetragonal, O - orthorhombic. The phase boundary between the tetragonal I4/mcm and P4/mbm phases is unknown and thus not indicated in the phase diagram. FA concentration dependence of the (b) activation energy of the dipolar relaxations and (c) room-temperature and low-temperature dielectric permittivity (1 MHz). Adapted with permission from ref (165). Copyright 2021 American Chemical Society.
A comparison of the phase diagrams of the MA1–xFAxPbBr3 and MA1–xFAxPbI3 systems shows a qualitative agreement for low values of x (see Figures 5d and Figure 7a). However, the central parts of the diagrams are significantly different, as higher symmetry phases were found at low temperature for the MA1–xFAxPbI3 perovskites (see Figure 5).88,336 We note that such differences between both systems are expected due to different phase behaviors of the iodide and bromide parent compounds.
In the same study,165 we also reported the temperature-dependent broadband DS experiments (Hz-GHz frequency range) of the MA1–xFAxPbBr3 single crystal samples. The most important results comparing the dielectric responses of the parent compounds (x = 0 and 1) and the highest mixing case (x = 0.5) are presented in Figure 8. Similarly to the iodide analogue, the complex dielectric permittivity of MAPbBr3 compound exhibits an anomalous decrease at the tetragonal-orthorhombic phase transition related to the establishment of the long-range MA cation order.183,340,341 Other phase transitions are visible as small anomalies. On the other end of the composition, the dielectric response of FAPbBr3 perovskite shows several weak anomalies originating from the phase transitions (see phase diagram, Figure 7a).274 In addition, five distinct and well-resolved dielectric relaxations of unknown origin were identified in the orthorhombic phase (best visible in the ε″ data). Interestingly, these relaxations mimic the behavior of the isosymmetric transitions, as they also disappear upon mixing even for the lowest MA concentration. The most interesting dielectric response was observed for the highly mixed MA0.5FA0.5PbBr3 compound, where both phase transitions (cubic-tetragonal-orthorhombic) were visible as very weak anomalies (Figure 8c). In addition, a broad dielectric dispersion of MA cations reminiscent of a glassy phase was detected at low temperature in agreement with the iodide analogue (compare with Figure 6c). A strong proof of the dipolar glass phase would be observation of the Vogel–Fulcher freezing behavior of the relaxation time associated with the observed dielectric relaxation. However, it was found that the dynamics of this process follows the typical Arrhenius law preventing an unambiguous assignment of the dipolar glass phase.
Figure 8.
Temperature dependence of the complex dielectric permittivity of (a) MAPbBr3, (b) MA0.5FA0.5PbBr3, and (c) FAPbBr3 single crystals presented at selected frequencies. Arrows indicate phase transition anomalies and dielectric relaxations. Adapted with permission from ref (165). Copyright 2021 American Chemical Society.
The broadband DS can directly measure how mixing affects the value of the dielectric permittivity, which is a relevant parameter influencing the performance of photovoltaic devices.53,55,56,183,190−192 We observed that the MA and FA cation mixing significantly affects the room-temperature dielectric permittivity of the MA1–xFAxPbBr3 system (Figure 7c), as ε′ showed a rather strong decrease with increasing x, which was assigned to the diminishing number of the MA cations.165 Note that the permittivity value at low temperature was not affected by mixing (Figure 7c), indicating that it is originating from the inorganic part of the lattice. As discussed by Fabini et al.,342 such a uniform and rather high value of ε′ is related to the lattice polarizability invoked by the 6s2 lone-pair electrons and associated off-centering of lead ions.
In the same study, we also proved the dynamic effects in the MA1–xFAxPbBr3 system using the broadband DS.165 The relaxation times extracted from the dipolar relaxations were approximated using the Arrhenius law providing the activation energy Ea of the MA cation motion. More than a 2-fold increase (from 80 to 180 meV) of the Ea values with increasing x was observed, indicating that FA cations significantly raise the rotation barrier of the MA cations (Figure 7b). The increase of the activation energy with x may seem to be in contradiction with the aforementioned NMR study of the related MA0.22FA0.78PbI3 compound,204 where practically no change in Ea was observed upon mixing at room temperature. However, one should note that our study probed activation energies at significantly lower temperatures, where molecular cation motion was substantially hindered. Note that the dipolar dynamics observed by the DS can be extrapolated to room temperature, revealing a 2 ps dipolar relaxation time of MA in MA0.2FA0.8PbBr3, which is in a very good agreement with the MA reorientation time of 1 ps obtained in MA0.22FA0.78PbI3 using NMR.204 Interestingly, the activation energies of the five low-temperature processes of FAPbBr3 showed a very huge spread ranging from 145 to 12 meV (Figure 7b),165 which may be related to the recently proposed quadrupolar geometric frustration of the FA cations.270
The temperature-dependent Raman data reported in the same work165 provided information on the degree of the molecular cation ordering. For small values of x, the MA cation modes exhibited significant narrowing at the transition to the orthorhombic phase, indicating ordering of the cations. In contrast, no substantial narrowing was observed at the phase transitions points in the highly mixed and FA-rich compositions (x ≥ 0.5) showing the absence of the MA cation ordering in agreement with the low-temperature dielectric response. For all mixed compounds, the width of the FA band decreased with decreasing temperature, indicating slowing down of the FA cation dynamics.
In a recent study, Kalita et al. investigated structural phase transitions, framework deformation, and dynamic effects in MAPbBr3, MA0.5FA0.5PbBr3, and FAPbBr3 compositions.343 The heat capacity measurements revealed that the mixed composition exhibits two structural phase transitions at 164 K (cubic-tetragonal) and 128 K (tetragonal-orthorhombic) in a very good agreement with the phase diagram reported in ref (165). Based on the synchrotron XRD measurements, the symmetry of the orthorhombic phase for the x = 0.5 solid solution was assigned to P4/mbm. In addition, it was observed that the Pb–Br length increases upon mixing, which was assigned to a larger radius of FA cation. It was also found that the PbBr6 octahedral tilting in the tetragonal phase increases with an increase of x, while the highest distortion of the octahedra in the orthorhombic phase was observed for the MAPbBr3 compound. The temperature-dependent Raman spectroscopy was employed to study the phonon modes in the low-frequency range (50 to 500 cm–1). In agreement with the previous study,165 it was found that the modes become well resolved in the orhothorhombic phase of MAPbBr3 due to the ordering of the system. In the mixed composition, the low-frequency modes were found to be much broader, indicating disorder compared to MAPbBr3.
4.2. MA-Based Compounds
Here, we review how mixing with less compatible A-site cations affects the phase transitions and dynamic properties of the MAPbX3-based lead halide perovskites. Mixing with such cations results in a solubility limit, beyond which the phase separation (mixed phase) typically occurs.
4.2.1. MA1–xCsxPbBr3
Mozur et al.163 reported a detailed multitechnique investigation of the mixing effects in the MA1–xCsxPbBr3 system. The employed advanced synchrotron and neutron diffraction methods allowed them to construct a rich phase diagram of this system (Figure 9a). Interestingly, despite both parent compounds existing in a 3D perovskite form (MAPbBr3 and CsPbBr3), the authors observed a solubility limit of Cs+ cations in MAPbBr3 (x ≈ 0.4). This was explained by different octahedral tilting preferences of smaller Cs+ cations (Table 1).
Figure 9.
(a) Temperature–composition phase diagram of the mixed MA1–xCsxPbBr3 system. Phase separation occurs for x ≥ 0.4. (b) Neutron powder diffraction data of the x = 0.2 composition showing the reentrant phase transition (arrow). (c) Fraction of the MA cations participating in the reorientation around the C4 symmetry axis (inset) as extracted from the QENS data for different Cs concentrations. Reprinted (adapted) with permission from ref (163). Copyright 2017 American Chemical Society.
The authors also found that the x = 0.1 composition exhibits the same sequence of the phase transitions (cubic-tetragonal-orthorhombic) as in pure MAPbBr3. The situation is different for x = 0.2, where, instead of the tetragonal-orthorhombic transformation, a reentrant transition back to the same Pm3̅m cubic phase was observed (Figure 9b). Such a surprising behavior was explained using the modified Blume–Capel model with included strain coupling. At higher Cs fraction (x = 0.3), the system transitioned from the cubic symmetry to the tetragonal P4/mbm phase instead of I4/mcm observed at lower values of mixing. Note that the P4/mbm space group is also observed for pure CsPbBr3 compound (Figure 3), indicating that Cs+ ions already dictate the octahedral tilt pattern. Interestingly, the transition to the orthorhombic phase was completely suppressed for this Cs concentration. This composition also showed the maximum extent of the cubic phase reaching 200 K. The situation changed again for the x = 0.4 composition, where the orthorhombic phase recovered and showed coexistence with the tetragonal phase (Figure 9a).
The same study also used a suite of neutron diffraction and neutron scattering techniques to probe the MA cation arrangement and dynamics for different fractions of mixing.163 The neutron powder diffraction data demonstrated the absence of the preferred MA cation orientation for all values of x even at low temperature. This was also supported by the inelastic neutron scattering data, which provided evidence of the heterogeneous environments of both the inorganic framework and the MA cations. In addition, the authors observed an excess of low energy modes in the heat capacity data due to the increased disorder. All these findings were attributed to the orientational glass behavior present in both organic and inorganic sublattices and caused by the local-strain induced frustration. The dynamics of the MA cations in MA1–xCsxPbBr3 compounds were probed using QENS. The mean squared displacement of hydrogen atoms revealed decreasing amplitude with increasing x and lowering of temperature. This indicated that the motion of MA cations is more hindered in the highly mixed phases. This was also supported by the measurement of the elastic incoherent structure factor, which revealed a drastic decrease in the overall MA cation reorientation around the C4 symmetry axis upon mixing (Figure 9c). Interestingly, this observation seems to be different than detected in the MA1–xFAxPbX3 system, where for high levels of mixing the MA cations exhibited substantial dynamics even in the orthorhombic phase as is evident from the dielectric response (Figures 6 and 8).165,336
4.2.2. MA1–xCsxPbI3
In a recent study, Gallop et al. used XRD, Raman, and 2D IR spectroscopies to investigate the structure and MA cation dynamics of the MA1–xCsxPbI3 thin film samples.344 The authors concentrated on the x ≤ 0.3 compositions, as for higher Cs fractions signatures of the yellow phase were observed. Note that a similar solubility limit was obtained in the discussed MA1–xCsxPbBr3 system.163 The XRD and Raman experiments indicated shrinkage of the unit cell and a continuous tilting of the octahedral framework upon mixing due to the smaller size of the Cs+ cation. The 2D IR spectroscopy experiments performed at room temperature revealed that the introduction of Cs+ makes the MA cation dynamics faster. The characteristic time of the jumping dynamics increased from about 4 ps (x = 0) to 12 ps (x = 0.3), while the increase of the wobbling dynamics was less pronounced (0.5 ps vs 0.65 ps). The obtained results were correlated with the increased c/a distortion upon mixing, which was also confirmed by the MD simulations.
4.2.3. MA1–xDMAxPbX3(X = I, Br)
The effective radius of DMA is only slightly larger than that of FA (Table 1), making it a highly promising alternative cation for mixing in lead halide perovskites. However, pure DMAPbX3 compounds form hexagonal structures,345,346 making them only partially compatible with the 3D perovskites.
Shi et al. reported the first study68 on the DMA incorporation (x = 0.11) in MAPbI3. The authors found that mixing with DMA suppresses the tetragonal symmetry by stabilizing the cubic phase at room temperature, which provided a substantial boost in the carrier diffusion length and device stability. A similar work was reported by Shao et al.,78 where a series of compositions up to x = 0.25 were investigated and also linked to the enhanced photovoltaic performance and device stability. A stabilization of the cubic phase at room temperature for DMA concentrations higher than 6% was observed. The authors claimed that the cubic space group of the mixed compounds is noncentrosymmetric P4̅3m, which is different from that of pure MAPbI3 (centrosymmetric Pm3̅m). The absence of the centrosymmetry was not discussed or proved by the authors, making this assignment questionable. The same work also reported a rather coarse temperature dependence of the PXRD patterns of the MA0.91DMA0.09PbI3 sample showing that the cubic-tetragonal phase transition occurs at about 273 K. Surprisingly, the authors claimed the orthorhombic symmetry already below 233 K, which is a much higher transition temperature compared to pure MAPbI3 (see Figure 3).
Franssen et al.208 reported a solid-state NMR and PXRD study of the mixed MA1–xDMAxPbI3 (x ≤ 0.21) system. The authors obtained that the DMA solubility limit in MAPbI3 is 21%, above which the mixed phase was formed. In agreement with other works, stabilization of the cubic phase at room temperature was also obtained even for small levels of mixing. In addition, the authors performed 14N NMR relaxation experiments of the x = 0.21 sample and obtained similar reorientation times for both MA and DMA cations (close to 2 ps) at room temperature. This indicates that despite its bigger size, the DMA cation is still mobile in the mixed structure at room temperature.
The structural phase transitions and cation dynamics in the single crystals of mixed MA1–xDMAxPbBr3 bromide analogues were investigated by us using a suite of different experimental and theoretical techniques.164 The incorporated DMA fractions were determined to be 4, 14, and 21% by measuring the atomic C/N ratio. Crystals with higher mixing levels were not synthesized due to the finite DMA solubility limit in MAPbBr3 of about 30% reported by Anelli et al.86 The phase transition behavior in the mixed compounds was measured using temperature-dependent DSC, ultrasonic propagation, SCXRD, and broadband DS experiments. Upon introduction of a small amount of DMA, the temperatures of all three phase transitions decreased, indicating increased stability of the cubic phase as indicated in the phase diagram of this system (Figure 10a). Such a behavior is in agreement with the results obtained for the iodide analogues.68,78,208 For the x = 0.14 composition, no clear transition anomalies were observed in the DSC, ultrasonic, and DS experiments; however, the SCXRD data indicated a transition below 130 K to a phase having a similar symmetry to the orthorhombic phase observed for the x = 0 and 0.04 compositions. No symmetry lowering was observed for the highest mixing level (x = 0.21) as evident from the reciprocal space reconstruction (Figure 10b), suggesting a complete suppression of the phase transitions for this DMA concentration.
Figure 10.
(a) Tentative temperature–composition phase diagram of the mixed MA1–xDMAxPbBr3 system. Symmetry notation: C - cubic, T - tetragonal, O - orthorhombic. (b) Reciprocal space reconstruction (hk0 layer) of the x = 0.21 single crystal sample measured at 100 K indicating cubic symmetry. (c) Temperature dependence of the broadband dielectric response of the x = 0.21 single crystal sample. (d) Low-frequency dependence of ε″ at selected temperatures of the same sample. Solid curves are the best fits to the Cole–Cole relaxation model. Adapted with permission from ref (164). Copyright 2020 Springer Nature.
The same study164 reported the broadband DS experiments of the MA1–xDMAxPbBr3 single crystals in a vast frequency range covering more than 11 orders of magnitude in frequency (from mHz to GHz). For small mixing levels (x = 0.04), the temperature dependence of the dielectric response was essentially the same as for the MAPbBr3 compound with the tetragonal-orthorhombic phase transition anomaly being the dominant feature. However, a drastically different situation was observed for higher mixings (x = 0.14 and 0.21), where the transition anomaly was replaced by a very broad and highly frequency-dependent peak spanning from room temperature to 50 K at microwave and low frequencies, respectively (Figure 10c). The additional low-temperature experiments covering the mHz frequency range revealed very slow dynamics extending to even lower frequencies and temperatures (Figure 10d). These results demonstrated a gradual slowing down of the MA cation dynamics by approximately 12 orders of magnitude and implied formation of a glassy disordered phase (compare with Figure 2c). This behavior was also supported by the low-temperature heat capacity measurements, which revealed higher entropy states below 5 K only for the highly mixed compounds. In addition, the activation energy of the MA cation dynamics was determined from the frequency domain data, indicating close to 2-fold increase of Ea from 80 to 140 meV as the DMA content was increased. Note that a very similar behavior was also observed for the already discussed mixed MA1–xFAxPbBr3 system.165
In the same work, we also reported DFT and MC calculations of the mixed compositions supporting the experimental results.164 The DFT calculations revealed that the DMA cation is situated in the center of the lead–bromine cuboid cavity with two amine protons pointing to the halogens and the C–C bond directed along the <001> (or equivalent) direction (Figure 11a). The same calculations also provided rotation barriers around three orthogonal lattice directions of the DMA and two neighboring (nearest neighbor (NN) and next–next-nearest neighbor (NNNN)) MA cations (Figure 11b). A comparison was made between fully relaxed pseudotetragonal and nonrelaxed cubic (based on pure MAPbBr3, where a single MA was replaced by DMA) structures in order to clarify the effect of DMA incorporation on the neighboring MA cations. The rotation barrier of the DMA cation was found to be significantly higher than that of the MA cations in both structures, indicating that the DMA motion should be significantly hindered. Second, substantially higher rotation potentials of the MA cations were obtained in the relaxed structure demonstrating that the DMA cation perturbs their rotation via the lattice deformation. Third, the energy barrier was found to decrease with increasing distance to the DMA cation reflecting that the lattice deformation is a local effect. Based on these findings, a microscopic picture of the glassy phase formation was proposed, where a random distribution of DMA cations causes different local framework distortions resulting in a multiwell potential for MA cations. This results in cation frustration followed by the glassy phase formation and suppression of the structural phase transitions. Signatures of the orientational glass phase were also obtained using the MC simulations based on the dipole–dipole interaction model developed by the Walsh group.244 Here, the MA cation (dipole) was allowed to rotate, while the orientation of the DMA dipole was randomly fixed toward a <100> (or equivalent) lattice direction, which should be a valid approximation at low temperature. Despite a simplified nature of this mode, a long-range order emerged for low mixing levels, while no ordering was observed for x = 0.2 indicating that the MA dipoles experience severe competing interactions (Figure 11c).164
Figure 11.
(a) Lowest energy structure of the x = 0.125 supercell obtained by DFT calculations, as seen from the <100> (top), <010> (middle), and <001> (bottom) directions. For clarity, only a 2 × 2 × 1 slab containing DMA is presented. (b) DFT calculations of the energy surfaces obtained by rotating DMA and neighboring MA cations around the <100>, < 010>, and <001> directions within the nonrelaxed and relaxed x = 0.125 systems. (c) Snapshots of the MC simulations for a two-dimensional slice of a three-dimensional periodic slab representing the x = 0 and x = 0.2 compositions. The orientations of the MA and DMA dipoles are depicted by gray and red arrowheads, respectively. Reprinted with permission from ref (164). Copyright 2020 Springer Nature.
A subsequent study of the mixed MA1–xDMAxPbBr3 system was published by Ray et al.,74 where a substantially higher DMA incorporation fraction of 44% (supported by solution 1H NMR) was reported using solvent acidolysis crystallization. The temperature-dependent synchrotron radiation PXRD experiments were performed to study the phase symmetry of the x = 0.44 composition. A symmetry lowering from cubic to the tetragonal symmetry was observed at about 205 K, and no orthorhombic phase was recorded down to 80 K. This result is in contrast with the study discussed above,164 where the cubic phase was maintained at least down to 100 K for the x = 0.21 composition (Figure 10a). A likely explanation for this discrepancy is that, in contrast to high-resolution synchrotron PXRD, the conventional SCXRD was not sufficient to resolve small changes associated with this symmetry lowering. Interestingly, the authors claimed that the space group of this tetragonal phase is P4/mbm, while pure MAPbI3 shows the I4/mcm symmetry. This transformation was explained by the difference in H-bonding between the molecular cations and the inorganic framework. Note that the identical change of the tetragonal unit cell symmetry upon mixing was observed for the MA1–xCsxPbBr3 system (Figure 9a).163
4.2.4. MA1–xEAxPbI3
The ethylammonium cation has an effective radius of 274 pm (Table 1), making it a good candidate for incorporation in lead halide perovskites. Similarly to other big cations, pure EAPbI3 compound does not form a 3D perovskite and instead crystallizes into a 1D topology.166,347−349
Peng et al. reported the first structural study of the mixed MA0.83EA0.17PbI3 system,350 where the concentration of the incorporated EA cations was determined using the C/N analysis and solution NMR spectroscopy. Using PXRD and SCXRD methods, the authors observed cubic Pm3̅m symmetry at room temperature, indicating that EA cations also stabilize the desirable cubic phase.122 The temperature-dependent PXRD experiments revealed that the phase transition to the tetragonal phase occurs at about 243 K, which is almost 90 K lower compared to MAPbI3. Stabilization of the cubic symmetry at room temperature was also observed in two other studies68,351 reporting very similar levels of mixing (x = 0.15 and 0.14). Surprisingly, another study352 of MA1–xEAxPbI3 thin films reported a tetragonal symmetry at room temperature up to x = 0.3 EA fraction, above which the phase separation occurred. A DFT study by Liu et al.84 considered the whole range of mixed compositions (x = 0, 0.25, 0.5, 0.75, and 1) and found almost negligible changes of the Pb–I– Pb bond angles upon mixing with reasonable levels of EA cations (x ≤ 0.5).
A detailed multitechnique study of the mixing effects on the structural phase transitions, cations dynamics and PL of the MA1–xEAxPbI3 (x = 0, 0.09, 0.16, 0.21, 0.31, and 0.38) compounds was recently reported by us.166 The EA concentration was determined using solution 1H NMR spectroscopy, and the solubility limit of about 40% was found. A set of different experimental techniques (DSC, Cp, ultrasonic, SCXRD, broadband DS, and Raman spectroscopy) was used to construct the temperature–composition phase diagram of this system (Figure 12a). For low and intermediate levels of mixing, the temperature of the cubic-tetragonal phase transition significantly decreased with increasing x, indicating stabilization of the cubic phase. For x = 0.16, this transition occurred at 253 K in a good agreement with the temperature of 243 K reported by Peng et al. for the x = 0.17 composition.350 The temperature of the tetragonal-orthorhombic phase transition also exhibited a decrease with increasing EA content, although the effect was slightly less pronounced (Figure 12a). The situation changed substantially for higher mixing levels (x > 0.3), where this transition became completely suppressed, and the low-temperature phase was described by a primitive tetragonal or pseudotetragonal (with very weak orthorhombic deformation) unit cell. However, due to the pseudomerohedral twinning, a reliable model of this structure was not obtained. The temperature-dependent Raman experiments of this phase revealed a substantial disorder related to the molecular cations even at low temperatures. The phase transitions of the pure 1D EAPbI3 compound were also investigated providing hexagonal-orthorhombic-monoclinic symmetry lowering (also summarized in Figure 12a). Another transition was also observed at an unusually low temperature of 60 K, but the symmetry below this point was not determined.
Figure 12.
(a) Temperature–composition phase diagram of the mixed MA1–xEAxPbI3 system. Symmetry notation: C - cubic, T - tetragonal, O - orthorhombic, M - monoclinic, H - hexagonal. EA concentration dependence of the (b) activation energies and (c) ε′ (1 MHz) obtained at 290 and 10 K. (d) Temperature dependence of the complex dielectric permittivity of (d) MA0.84EA0.16PbI3 and (e) MA0.62EA0.38PbI3 single crystals presented at selected frequencies. Arrows indicate phase-transition anomalies, while dielectric relaxations are marked by τ1, τ2, and τ3. Reprinted with permission from ref (166). Copyright 2022 American Chemical Society.
The broadband DS was employed in the same work166 to probe the dipolar dynamics of the mixed MA1–xEAxPbI3 single crystal compounds. For the intermediate values of x, the tetragonal-orthorhombic transition anomaly became broader (Figure 12d) and shifted to a lower temperature. In addition to the expected dipolar relaxation of MA cations in the tetragonal phase (indicated by τ1 in Figure 12d), an additional dipolar process τ3 of undetermined origin emerged below 100 K. For x > 0.3, the transition anomaly transformed into a broad dipolar relaxation of MA cations extending to low temperatures (indicated by τ2 in Figure 12e). The activation energies of both processes associated with the MA dynamics increased with increasing level of mixing (Figure 12b). Note that this effect was also observed in the MA1–xFAxPbBr3 and MA1–xDMAxPbBr3 systems suggesting universal behavior.164,165 Upon mixing, the dielectric permittivity of the MA1–xEAxPbI3 compounds also exhibited some tuning (Figure 12c), although the effect was rather small likely due to comparable electric dipole moments of both molecular cations (Table 1).
The observed broad dielectric response for the x > 0.3 compositions is similar to other highly mixed compounds (Figures 6, 8, and 10), and, as discussed above, resembles the glassy phase (compare with Figure 2c). However, a clear Vogel–Fulcher behavior of the mean relaxation time was not detected despite broadband measurements (Hz-GHz),166 indicating that freezing of electric dipoles may occur at very low temperatures. Note that the DFT calculations of the MA0.875EA0.125PbI3 composition revealed about a 20% variation of the rotational barrier for crystallographically different MA cations, indicating that the MA dynamics are affected by the local lattice strains introduced by bigger EA cations.
4.2.5. MA1–xGAxPbI3
Guanidinium is another highly promising molecular cation used to significantly enhance the performance of the devices based on lead halide perovskites.69,73,75,75,77,80 The effective radius of this cation is 278 pm, which is practically the same as of DMA and EA (Table 1). For this reason, we may expect similar mixing effects on the structural and dynamic properties of lead halide perovskites.
Several studies reported contrasting solubility limits of these cations in MAPbI3. Jodlowski et al.69 reported a solubility limit of 25% in the perovskite film samples, and a similar value of 20% was claimed in another study.353 A significantly higher solubility limit of 40% (based on 13C MAS NMR measurements) was reported by Kubicki et al.75 for powder samples obtained by mechanosynthesis. Contrary to these studies, Gao et al.80 measured the solubility limit using NMR to be only 5% in single crystal samples. This result is also in sharp contrast to the solubility limits obtained for other similar size cations such as DMA and EA, prompting further studies to clarify this discrepancy.
Regarding structural properties, several studies reported preservation of the tetragonal symmetry at room temperature in the mixed MA1–xGAxPbI3 system.69,73,75,80,353,354 This observation is also in contrast to many other already discussed mixed-cation systems, where stabilization of the cubic phase was observed. Gao et al.80 observed almost negligible changes of the octahedral tilting and Pb–I–Pb bond lengths and angles for the mixing compositions of x = 0.05 and 0.1 showing that these systems remained tetragonal at room temperature. This discrepancy was also recognized and investigated in more detail in a recent study by Minussi et al.,355 where the phase transition and dielectric properties of the mixed MA1–xGAxPbI3 (x = 0, 0.1, 0.2) system were reported. The authors performed temperature-dependent XRD and DSC experiments, which indeed revealed stabilization of the cubic phase for the x = 0.2 compound at room temperature (Figure 13a). This removed the GA cation ambiguity and placed it in the same league with other molecular cations of similar size such as DMA, EA, and FA.
Figure 13.
(a) Variation of the lattice parameters with temperature of the MA0.8GA0.2PbI3 system as determined from the XRD data. (b–d) Temperature dependence of the real part of the complex dielectric permittivity ε′ of the MA1–xGAxPbI3 pellet samples at frequencies between 10 kHz and 1 MHz. Reprinted with permission from ref (355). Copyright 2022 Royal Society of Chemistry.
The same work also reported a temperature dependence of the real part of the complex dielectric permittivity ε′ of the MA1–xGAxPbI3 pellet samples in a narrow frequency range (10 kHz - 1 MHz)355 (Figure 13b–d). The obtained results for pure MAPbI3 compound show an expected anomalous decrease of the permittivity at the tetragonal-orthorhombic phase transition point in agreement with other studies.183,340,341 For the x = 0.1 composition, this anomaly became significantly broader, and its maximum shifted to a slightly lower temperature, suggesting that the stability of the tetragonal phase is also increased. The peak of the anomaly was not detected for the x = 0.2 compound likely due to the limited temperature range of the reported experiment. In addition, the authors observed a significant decrease of the dielectric permittivity upon mixing, which can be explained by a negligible electric dipole moment of GA (Table 1).
The dynamics of the molecular cations in the MA1–xGAxPbI3 solid solutions was investigated by the Emsley group using MAS NMR spectroscopy. Their first study75 reported the reorientation rates of both molecular cations for the x = 0.25 composition by fitting the corresponding 14N NMR spectra. Such an analysis provided 113 ps and less than 18 ps reorientations rates of MA and GA cations at room temperature, respectively. The observed almost 1 order of magnitude faster GA motion was also correlated to a spectacular increase of the charge carrier lifetimes in this mixed compound. However, a recent study from the same group204 reevaluated these findings using more reliable quadrupolar relaxation measurements, which provided much shorter reorientation times of both cations for the same mixed composition (x = 0.25). The MA dynamics perpendicular to the C3 molecular axis were found to occur on the 1.3 ps time scale with the activation energy of 168 meV at room temperature. A substantially faster rate of 0.3 ps was measured around the C3 axis with a negligible energy barrier. Note that MA cations in pure MAPbI3 showed very similar reorientation rates, indicating a weak mixing effect on the MA cation motion at room temperature. The dynamics of the GA cation was also characterized by two distinct principal axes owing to its C3 symmetry. Here, the motion perpendicular to this axis was obtained to be much faster (1.3 ps, Ea = 121 meV) compared to the parallel reorientation (14 ps) in line with different moments of inertia.
4.2.6. MA1–x–yGAxFAyPbI3
Recently, the Araújo group reported several studies of triple-cation MA1–x–yGAxFAyPbI3 solid solutions obtained by mechanosynthesis. In the first work,356 the authors used PXRD, DSC, and IR spectroscopy to study the structural phase transitions and cation dynamics for the case, where the GA and FA content was equal (i.e., x = y) and did not exceed 30% of the total MA substitution. Upon increase of x and y, it was found that the cubic-tetragonal phase transition temperature is substantially lowered to about 270 K at the maximum mixing level (x = y = 0.15), indicating increased stability of the cubic phase. The IR spectroscopy measurements revealed a gradual shift of the C–N stretching frequency for all organic cations with mixing, which was assigned to the increase of the bond strength. The intensity of the IR bands also exhibited dependence on the mixing level especially when crossing the cubic-tetragonal phase transition. This behavior was related to the degree of cation freedom within differently deformed inorganic framework in the cubic and tetragonal phases.
Subsequent XRD, DSC, and DS studies from the same group357,358 reported a thorough investigation of the MA1–x–yGAxFAyPbI3 solid solutions for 0 ≤ x ≤ 0.3 and 0 ≤ y ≤ 0.3 with x + y ≤ 0.3. The reported composition dependence of the cubic-tetragonal phase transition temperature is presented in Figure 14a, revealing that FA cations are more effective in stabilizing the cubic phase compared to GA.357 Interestingly, the opposite behavior is observed for the tetragonal-orthorhombic phase transition, where the GA-rich compositions have significantly lower transition temperatures (Figure 14b). Both observations clearly indicate that GA cations are more efficient in stabilization of the tetragonal phase compared to FA cations.
Figure 14.
Phase diagrams for the (left) cubic-tetragonal and (middle) tetragonal-orthorhombic phase transitions of the MA1–x–yGAxFAyPbI3 system. (Right) Composition map for of the real part of the complex dielectric permittivity ε′ obtained at room temperature and 1 MHz probing frequency. Reprinted with permission from ref (357). Copyright 2023 Wiley-VCH.
In addition to phase transition studies, the same study reported the DS experiments of the MA1–x–yGAxFAyPbI3 pellet samples.357 The measured values of ε′ obtained at room temperature and 1 MHz probing frequency are summarized in Figure 14c, indicating that GA incorporation results in a significantly more pronounced drop of the dielectric permittivity compared to FA cations. This result is in agreement with the aforementioned DS studies of the MA1–xFAxPbBr3165 and MA1–xGAxPbI3355 solid solutions. Such a contrasting FA and GA cation effect on the dielectric response might be related to the absence of the electric dipole moment of GA, although the dipole moment of FA is also almost negligible compared to MA (see Table 1). As also pointed out by the authors, this behavior may also indicate that both cations exert a different hindering effect on the neighboring MA cations. However, this interpretation seems to contradict the recent NMR study by Mishra et al.,204 where the effect of the incorporated GA and FA cations on the MA cation motion was found to be small in MA1–xFAxPbI3 and MA1–xGAxPbI3 compounds, suggesting a more intricate mechanism. We also note that the reported DS experiments were performed on the pressed pellet samples, which always introduces a degree of uncertainty in the determination of the dielectric permittivity.
4.3. FA-Based Compounds
The current best performing hybrid perovskites for photovoltaic applications are based on the mixed compositions with α-FAPbI3.8,9,11 Despite this significance, the mixing effects on the structural phase transitions and dynamics in these compounds are substantially less investigated compared to the MA-based perovskites. In this section, we review the available studies, which mostly concentrate on mixing with Cs+ and GA cations, as mixing with MA is covered above. Note that many works report stabilization of the photoactive black α-phase of FAPbI3 upon mixing (e.g., refs (11, 72, 76, 110, 359).). However, the majority of these studies fall out of scope of this review, as they concentrate on utilization of the stabilized α-FAPbI3 phase for device fabrication and provide no or very little additional information about the structural and dynamic effects.
4.3.1. FA1–xCsxPbX3(X = I, Br)
Charles et al.171 used a suite of diffraction techniques (neutron powder diffraction, SCXRD, PXRD) to thoroughly probe the phase transition behavior in the mixed FA1–xCsxPbI3 system. A solubility limit of about 15% was determined using PXRD in agreement with the magnetic resonance studies,207,360 while a substantially higher limit close to 30% was reported for thin films.95 To avoid phase separation into black and nonperovskite yellow components observed for higher mixing levels, the authors selected x = 0.1 concentration for a more detailed investigation. The room-temperature diffraction measurements revealed that the time averaged structure of this compound has a cubic Pm3̅m symmetry in agreement with pure FAPbI3 perovskite (Figure 3). It is also noted that the local structure may contain pseudotetragonal domains due to tilted octahedra switching rapidly between preferred directions. The determined Pb–I–Pb angle in the mixed composition was found to be 162.8°, which is almost identical to the nonmixed FAPbI3 compound (163.7°).222 Interestingly, upon introduction of Cs+, the phase transition temperature to the tetragonal P4/mbm phase slightly increased (290 K), which likely occurs due to the significantly higher phase transition temperatures of CsPbI3. Below 180 K, the diffraction data were indexed to either crystallographic twins in the tetragonal space group P4/mbm or orthorhombic Pnma. A convergence to the orthorhombic Pnma symmetry was obtained below 125 K with additional weak reflections attributed to glass-like disorder, which was also proposed for pure FAPbI3.209 The transition temperature to this phase was lowered upon mixing. All these results are summarized in a crude phase diagram presented in Figure 15a. Note that signatures of the glassy phase formation at low temperature in the FA1–xCsxPbI3 system were also observed in a recent MC study.228
Figure 15.
(a) Tentative temperature–composition phase diagram of the mixed FA1–xCsxPbI3 system. Symmetry notation: C - cubic, T - tetragonal, O - orthorhombic, disT - disordered tetragonal, disO - disordered orthorhombic. Adapted with permission from ref (171). Copyright 2020 American Chemical Society. (b) Temperature dependence of the heat capacity of the x = 0 and x = 0.1 compounds. Reprinted with permission from ref (186). Copyright 2019 American Chemical Society. (c) Vector autocorrelation function of FA cation in the x = 0 and x = 0.1 compositions showing the probability of the cation remaining in its initial orientation over time. Molecular vector is indicated in the inset. Adapted with permission from ref (226). Copyright 2017 American Chemical Society. (d) Anisotropic rotational correlation time (298 K) of FA cations in the x = 0 and x = 0.05 compounds determined using NMR. Reprinted with permission from ref (204). Copyright 2023 American Chemical Society.
Kawachi et al.186 reported heat capacity measurements of the high quality FA1–xCsxPbI3 (x = 0 and 0.1) single crystals. For pure FAPbI3 compound, a second-order transition from the cubic to tetragonal symmetry was observed at 280 K (Figure 15b) in agreement with other works (Figure 3). Interestingly, two sequential first-order transitions at 141 and 130 K were observed, as the system evolved to the low-temperature phase. The origin of this intermediate phase was not investigated. An additional weak anomaly was detected at about 80 K. The temperature dependence of the heat capacity of the mixed (x = 0.1) crystal revealed that the cubic-tetragonal phase transition shifted to higher temperatures (300 K) (Figure 15b). In addition, mixing resulted in a complete suppression of the intermediate phase, as a single phase transition was detected at 149 K, while the low-temperature anomaly at about 80 K was unaltered. Here we note that mixing also suppressed the intermediate phase of MAPbBr3.165 The measured transition temperature of 149 K is not in line with the results reported by Charles et al.171 for the same mixed composition, where transitions were observed at 125 and 180 K (see Figure 15a for comparison). Further studies are needed to clarify this discrepancy.
Ghosh et al.226 used ab initio MD simulations to study the microscopic dynamics in the mixed FA0.9Cs0.1PbI3 and FA0.9Rb0.1PbI3 perovskites. The authors reported that the introduction of smaller cations such as Cs+ and Rb+ (Table 1) results in a greater tilt angle of the octahedra compared to pure FAPbI3. In addition, this tilting is locked, and thus the octahedra undergo much more restricted rocking dynamics. This also affects the neighboring molecular cations, which form stronger intermolecular N–H···I H-bonds. As a result, it was observed that the tumbling time scale of the FA cations slowed down more than twice from 2 ps to almost 5 ps (Figure 15c) upon introduction of 10% of Cs. For comparison, the anisotropic rotational correlation time of FA cations in the x = 0 and 0.05 compounds was recently measured using quadrupolar relaxation NMR experiments.204 Slightly slower dynamics were indeed observed in the mixed system, especially around the molecular x-axis (see Figure 15d), although the uncertainties of the determined values were rather high. Note that a later study found that, in contrast to Cs+, Rb+ ions have no capacity to be incorporated in FAPbI3.207
Mundt et al.361 studied FA0.83Cs0.17PbI3 thin films to demonstrate that nanoscale compositional heterogeneity can serve as initiation sites for more macroscale irreversible phase segregation. The degree of nanoscale heterogeneity is kinetically controlled and the well-mixed system can be reached when the sample is annealed either long enough or at sufficiently high temperature. In particular, the authors showed that the well-mixed sample, which can be achieved by annealing at 140 °C for 30 min, exhibits a sharp cubic-tetragonal phase transition at 50 °C, while the poorly mixed sample, prepared by annealing at 140 °C for 15 min, showed a smeared out phase transition, reaching the cubic phase close to 130 °C. The smeared phase transition was attributed to a higher level of nanoscale heterogeneity, since various local compositions exhibited phase transitions at various temperatures. This paper emphasized the importance of kinetics on formation of perovskite phases and consequently on the properties of the synthesized thin films.
The mixed FA1–xCsxPbBr3 bromide analogues were thoroughly investigated by Mozur et al.172 using XRD, neutron scattering, and NMR techniques. The solubility limit of Cs in FAPbBr3 was found to be about 40%, beyond which a phase separation occurred. Note that a very similar limit was also observed for the related MA1–xCsxPbBr3 and FA1–xCsxPbI3 systems.110,163 The authors used temperature-dependent high-resolution synchrotron XRD to map the phase diagram of FA1–xCsxPbBr3 (Figure 16a). For all studied concentrations, a cubic-tetragonal phase transition was detected with its temperature shifting to higher values upon mixing. This indicates reduced stability of the cubic phase in agreement with the phase diagram of the FA1–xCsxPbI3 system (Figure 15a). More importantly, a complete suppression of all other (tetragonal-orthorhombic and three isosymmetric270) phase transitions was observed even for the lowest mixing level (x = 0.05). However, as stated by the authors, the orthorhombic phase of pure FAPbBr3 already exhibits features of low intensity, which may be even more difficult to resolve in the mixed compositions. Note that a complete disappearance of the isosymmetric transformations of FAPbBr3 was also observed upon incorporation of a small amount of MA cations, indicating that these transitions are very sensitive to small perturbations.165
Figure 16.
(a) Temperature–composition phase diagram of the mixed FA1–xCsxPbBr3 system. Symmetry notation: C - cubic, T - tetragonal, O - orthorhombic. (b) Room-temperature 14N MAS NMR spectra of the FA1–xCsxPbBr3 compounds. (c) Schematic cartoon illustrating the FA cation arrangement in a pure FAPbBr3 system. Such a “T”-type (red) long-range quadrupolar order is only possible in a two-dimensional plane. (d) Incorporated Cs+ ion disrupts the “T”-type arrangement, as neighboring FA cations point to the positive charge of Cs+ causing disordering of the next-nearest neighbor quadrupoles (black arrows). Reprinted (adapted) with permission from ref (172). Copyright 2020 American Chemical Society.
In the same study, the mechanism behind the suppression of the transitions was investigated in more detail.172 The room-temperature 14N NMR spectra of different compositions indicated increased disorder, i.e., the presence of a larger distribution of local FA environments upon mixing (Figure 16b). The same behavior was also observed for the inorganic framework using the 79Br nuclear quadrupole resonance spectroscopy. However, inelastic neutron scattering data revealed that mixing has only minimal influence on the internal vibrational modes of the FA cations. The 1H NMR relaxation time measurements supported this result by showing that the cation reorientation rates and activation energies are also relatively weakly affected by mixing. Based on these observations, the authors concluded that the incorporated Cs+ disrupts the FA cation ordering on the longer range, while their local degrees of freedom are retained. In the previous work,270 the same group proposed that the crystallographically unresolvable isosymmetric phase transitions of FAPbBr3 are caused by the long-range dynamics of the FA cations. This behavior is enabled by the electrostatic quadrupolar interactions between the cations (quadrupole moment |Q11| = 18.3 DÅ), which highly prefer the “T”-type (perpendicular) arrangement (Figure 16c). However, such long-range dynamics do not lead to a long-range order in the system, as it is not possible to form an ordered “T”-type structure in three dimensions, meaning that FA cations are already geometrically frustrated in pure FAPbBr3. Incorporation of Cs+ ions causes local alignment of the neighboring quadrupole moments disrupting the local “T”-type ordering (Figure 16d) and interrupting the cooperative dynamics associated with these transitions.
4.3.2. FA1–xGAxPbI3
Mixing of GA in the structure of FAPbI3 was briefly studied by Kubicki et al.75 using room-temperature 13C and 14N MAS NMR spectroscopy of FA cations. The authors showed that incorporation of GA (x = 0.25) causes a significant broadening of the FA spectral envelope compared to pure α-FAPbI3. This indicates decreased reorientation dynamics of FA cations upon mixing with GA cations. Interestingly, it was also observed that the mixed FA1–xGAxPbI3 compounds are thermodynamically unstable and become yellow within hours after annealing, meaning that GA cations do not fully stabilize the desirable black phase.
4.4. Summary
Several general observations can be made regarding the A-site mixing effects on the structural and dynamic properties of MAPbX3 and FAPbX3 systems. As expected from the tolerance factor arguments, the size of the guest cation is related to the solubility limit, which ranges from 100% for compatible cations (MA/FA systems) to 20–40% for substantially bigger or smaller guests (DMA, EA, GA, Cs+). The mismatch in the cation size also greatly affects the structural and dynamic properties of the mixed compounds.
The A-site mixing with molecular cations is an effective recipe to suppress the structural phase transitions and thus symmetrize the lattice by stabilizing the desirable cubic phases.122 This does not hold for mixing with smaller Cs+ ions, which tend to stabilize the tetragonal phase likely inherited from pure CsPbX3 perovskites. The extent of transition suppression also depends on the cation size - mixing of relatively well compatible MA and FA cations results in a weak suppression, while introduction of big cations (low solubility limit) can cause a complete disappearance of the transitions. This behavior originates from a significant lattice deformation introduced by more bulky molecular cations, which disrupt the H-bonding and octahedral tilt patterns and thus prevent the cooperative long-range ordering.
Similarly to classical inorganic systems, the highly mixed lead halide perovskites also show very broad dielectric responses both in temperature and frequency domains. A comparison of different mixed systems (MA0.5FA0.5PbBr3, MA0.79DMA0.21PbBr3, and MA0.62EA0.38PbI3) is presented in Figure 17 revealing a broad dipolar dispersion of MA cations ranging from almost room temperature at the GHz band to very low temperatures at mHz frequencies (see also Figure 10d). Note that this dispersion is weakly affected by the incompletely suppressed structural phase transitions (e.g., in MA0.5FA0.5PbBr3). The main features of the observed dielectric response resemble the dipolar (orientational) glass (Figure 2c). A likely origin of this phase is frustrated interactions between the molecular cations, which are mediated by the lattice deformations introduced by mixing. It should also be noted that dipolar glass should exhibit freezing dynamics following the Vogel–Fulcher law, but no such behavior was detected for the mixed lead halide perovskites. A possible explanation for this discrepancy is that freezing actually happens, but at low temperatures, where the dipolar relaxation is already shifted to very low frequencies (sub mHz) making the DS experiments infeasible during a reasonable amount of time.
Figure 17.
Comparison of the broadband DS of the highly mixed (a) MA0.5FA0.5PbBr3, (b) MA0.79DMA0.21PbBr3, and (c) MA0.62EA0.38PbI3 systems. (a) Reprinted (adapted) with permission from ref (165). Copyright 2021 American Chemical Society. (b) Reprinted (adapted) with permission from ref (164). Copyright 2020 Springer Nature. (c) Reprinted (adapted) with permission from ref (166). Copyright 2022 American Chemical Society.
The activation energy of the MA motion in different mixed systems determined by the broadband DS is summarized in Figure 18a.164−166 An increase of Ea with increasing x is evident, indicating that guest cations raise the rotational barrier for MA cations. In contrast to DS, the activation energy of the MA reorientation at room temperature determined by NMR shows a negligible change upon mixing with FA and GA (Figure 18a), which is likely related to different temperature ranges probed by both techniques.
Figure 18.
(a) Activation energy of MA cation motion vs guest cation concentration for different MA-based mixed systems determined using the dielectric (solid) and NMR (open points) spectroscopies. Solid curves are a guide for the eye. Data taken from refs (164−166, 204). (b) Rotational correlation times at room temperature of FA, MA, and GA molecular cations in different mixed perovskites obtained by NMR. Reprinted with permission from ref (204). Copyright 2023 American Chemical Society.
Here, we also show a summary of the room-temperature rotational correlation times of MA, FA, and GA cations in different mixed compounds determined in a recent NMR study (Figure 18b).204 Interestingly, all cations exhibit dynamics on a similar few ps time scale. For example, MA and FA reorientation times seem to be only weakly affected by mixing. Note that a very similar time scale in the range of few ps can be obtained by extrapolating the MA relaxation times obtained from the broadband DS (Figure 17). This shows that both techniques essentially probe the same type of dynamics.
5. X-Site Mixing in 3D Perovskites
Here, we survey how halide mixing at the X-site affects the structural, phase transition, and dynamic properties of the 3D lead halide perovskites. The majority of works, which fall within the scope of this review, describe the MA- and Cs-based compounds with an exception of MHy-based compositions.
5.1. MA-Based Compounds
5.1.1. MAPb(I1–xBrx)3
Noh et al. claimed a complete solubility of I– and Br– anions (effective radii 220 pm vs 196 pm362) at the X-site in the mixed MAPb(I1–xBrx)3 system.363 The authors also observed that the room-temperature crystal symmetry changes from tetragonal to cubic for bromine concentration x ≥ 0.2. A solubility limit was also not considered in more recent works, where the x = 1/3 and 2/3 compositions were investigated.364,365 However, as demonstrated by the Walsh group using DFT and free energy calculations, the MAPb(I1–xBrx)3 solid solution should exhibit a very wide miscibility gap (0.19 ≤ x ≤ 0.68) at room temperature, where the phase separation occurs.366
The latter result was also confirmed experimentally by Lehmann et al.367 using synchrotron PXRD, where an even wider miscibility gap of 0.29 ≤ x ≤ 0.92 was determined. In the iodine rich phase, a partial solubility of bromine was observed as revealed by a continuous shift of lattice parameters. In contrast, iodine formed cluster-like regions in the bromine-rich phase showing a very poor solubility.
Based on the temperature-dependent PXRD experiments, the authors also mapped the temperature–composition phase diagram of the MAPb(I1–xBrx)3 system (Figure 19a).367 A clear decrease of both phase transition temperatures was observed upon mixing up to the miscibility gap in the iodine-rich region, indicating stabilization of the cubic phase. This behavior is analogous to mixing of molecular cations at the A-site. On the other end of the phase diagram, the cubic-tetragonal phase transition was found to be barely affected by mixing. However, the intermediate tetragonal phase of MAPbBr3 was suppressed indicating that, despite low solubility and tendency to cluster, iodine still perturbs the long-range order in the system.
Figure 19.
(a) Temperature–composition phase diagram of the MAPb(I1–xBrx)3 system. Symmetry notation: C - cubic, T - tetragonal, O - orthorhombic. Reprinted (adapted) with permission from ref (367). Copyright 2019 The Royal Society of Chemistry. Green square data point taken from refs (363, 368). (b) PXRD patterns of MAPbI2Br at different temperatures. (c) Temperature dependence of the real part of the dielectric permittivity of the x = 0, 1/3, 2/3, and 1 compounds measured 1 MHz frequency. Reprinted (adapted) with permission from ref (365). Copyright 2022 Wiley-VCH. (d) Transient anisotropy dynamics of MA cations in the x = 0, 0.5, and 1 compounds. Solid curves show the fits using the wobbling-in-a-cone relaxation model (inset). Adapted with permission from ref (369). Copyright 2017 American Chemical Society.
Interestingly, Tang et al. found that the miscibility gap in the MAPb(I1–xBrx)3 system could be avoided using the mechanochemical synthesis instead of a typical solvent synthesis.368 In such a way, the authors synthesized the whole series of compositions and observed no signatures of phase separation. In addition, it was found that the system changes room-temperature symmetry from tetragonal to cubic at x ≈ 0.2 in agreement with other works.363,367 For comparison, the authors also prepared MAPb(I1–xBrx)3 films using the wet chemical route and observed a wide miscibility gap (0.1 < x < 0.8) in line with other studies.366,367 No phase separation was also observed in a recent NMR and MD work by Fykouras et al.,205 where mixed compositions with x = 1/3, 1/2, and 2/3 were synthesized by mechanochemical synthesis. In this study, the authors also demonstrated that iodine and bromine ions are randomly dispersed and that reorientation of the MA cations within the inorganic cage highly depends on a local halide distribution. Note that Askar et al. also found no miscibility gap and a homogeneous halide distribution in the related mixed FAPb(I1–xBrx)3 and FAPb(Br1–xClx)3 systems using 207Pb NMR.370
A recent study by Shahrokhi et al.365 reported a more detailed multitechnique investigation of the phase transitions in MAPbI2Br (x = 1/3) and MAPbIBr2 (x = 2/3) single crystal compounds obtained using solvent synthesis. As discussed above, while the former compound could be considered to be on the edge of the miscibility gap, the latter composition should result in phase separation and iodide clustering. However, this issue was not addressed, as full solubility was assumed. The authors claimed no anomalies in the DSC and PXRD data for both compounds, indicating a complete suppression of the phase transitions and stabilization of the cubic phase. This result is also not in line with the phase diagram provided by Lehmann et al.,367 where transitions at about 270 and 115 K can be expected at the lower edge of the miscibility gap x ≈ 0.3 (Figure 19a). Indeed, a closer look at the PXRD data reported by Shahrokhi et al.365 seems to show some change of the diffraction patterns at 275 K (Figure 19b), suggesting that the phase transitions may actually be not fully suppressed. Interestingly, a substantial contribution of unusual monoclinic phase was also employed to better fit the PXRD data of both compounds, which may again be related to the unaccounted phase separation at these mixed compositions. Note that López et al.371 also did not observe any symmetry lowering in the synchrotron XRD data down to 120 K for the x = 1/3 sample, which was obtained using mechanochemical synthesis.
The same work by Shahrokhi et al.365 also reported the dielectric properties of the x = 0, 1/3, 2/3, and 1 compounds (Figure 19c). As expected, pure MAPbI3 and MAPbBr3 perovskites demonstrated a sudden decrease of the dielectric permittivity related to the MA cation ordering, as the materials transitioned into the orthorhombic phase.183,340,341 In contrast, the temperature dependence of the dielectric permittivity of the mixed compositions revealed a broad featureless peak extending below 100 K (Figure 19c). Note that a very similar dielectric response was also obtained for the A-site mixing (see Figure 17), indicating that both approaches affect the MA cation dynamics in a similar way. In addition, as pointed by the authors, such a dielectric response might be a signature of a dipolar (orientational) glass phase.
The MA cation dynamics in the mixed MAPb(I0.4Br0.6)3 film were investigated by Selig et al.369 using 2D IR spectroscopy and classical MD simulations. Note that the studied x = 0.6 composition should be deep within the miscibility gap (Figure 19a), which was not discussed by the authors. As found by Lehmann et al.,367 such a bromine concentration should result in phase clustering with partial mixing. The 2D IR spectroscopy of pure MAPbI3 and MAPbBr3 perovskites revealed wobbling and reorientation dynamics of the MA cations occurring on the 0.3 ps and 1.5–3 ps time scale, respectively. In the mixed compound, the wobbling dynamics remained the same, while the time scale of the cation reorientation increased significantly up to 15 ps. This indicates hindering of the MA cation motion in the mixed halide system likely caused by the local distortions of the inorganic framework. The experimental results were also supported by the MD simulations, which also revealed partial suppression of the MA dynamics in the mixed compound. Similar MD simulation results were also obtained for the x = 1/3 and 2/3 compositions in the aforementioned study by Shahrokhi et al.,365 while Grüninger et al.206 observed restriction in the cation mobility due to halide mixing for the MA0.15FA0.85Pb(I0.85Br0.15)3 composition. Note that the 2D IR spectroscopy of the MA1–xCsxPbI3 (x ≤ 0.3) mixture revealed a very similar change of the wobbling and jumping dynamics of the MA cations upon introduction of Cs+ ions.344
5.1.2. MAPb(I1–xClx)3
Due to an even larger difference in the ionic radii of I– and Cl– anions (220 pm vs 181 pm),362 an even wider miscibility gap is expected for the mixed MAPb(I1–xClx)3 system. Among the first studies of these compounds, Colella et al.372 reported a maximum chlorine content of x = 0.04, while somewhat higher incorporation up to x = 0.1 was stated by Unger et al.373 (both solvent synthesis). Pistor et al. reported the miscibility gap of 0.05 < x < 0.5.374 In a more recent work published by the Schorr group,375 the miscibility gap was refined using high-resolution synchrotron PXRD to be 0.03 < x < 0.99. Note that no deviation of symmetry from the parent compounds was observed in these weakly mixed compositions.
The same group also used inelastic and quasielastic neutron scattering to study the MA dynamics in the x = 0.02 compound.376 Upon introduction of chlorine, a significantly faster C3 reorientation around the C–N bond of the MA cations in the orthorhombic phase was observed compared to pure MAPbI3 (485 ps vs 1635 ps). In addition, mixing also decreased the activation energy of this motion from 41 to 22 meV. As discussed by the authors, these findings indicate that even a minute amount of chlorine is sufficient to substantially weaken the H-bonds between the MA cations and inorganic framework. In a recent study, the same group also reported that the temperature of the tetragonal-orthorhombic phase transition is lowered in the mixed composition (x = 0.02) by about 5 K compared to MAPbI3, indicating that a small amount of mixing also affects the structural phase transitions.377
5.1.3. MAPb(Br1–xClx)3
In contrast to mixing with iodine, no evidence of the miscibility gap was found for the mixed MAPb(Br1–xClx)3 perovskites due to more similar radii of Br– and Cl– anions (196 pm vs 181 pm).362364,368,369,377−382 Several studies observed cubic symmetry at room temperature for different compositions covering the whole mixing interval364,380,381,383 in agreement with the symmetry of the end members (see Figure 3).
A cubic symmetry of the x = 1/3, 1/2, 2/3 mixed compositions at room temperature was also observed by López et al. using neutron powder diffraction.379 In addition, the authors also determined the preferred MA cation orientation in the inorganic cage, which was found to change from the [110] to [111] and finally to the [100] direction, as the chlorine content increased, and the size of the unit cell shrunk. The change of preferred directions also indicates the reduction in the degrees of freedom of MA cation, as the delocalization along the [110], [111], and [100] axis involves 6, 4, and 3 possible orientations, respectively.
The phase transitions in the mixed MAPb(Br1–xClx)3 system were first investigated by Alvarez-Galván et al. using synchrotron XRD.380 The authors observed a complete suppression of the phase transitions for the studied highly mixed compositions (x = 1/2 and 2/3) and stabilization of the cubic Pm3̅m phase.
A similar behavior was also obtained in a recent multitechnique study by van de Goor,381 where a series of mixed compounds was investigated covering the whole range of compositions. A complete suppression of the phase transitions was obtained for the 0.2 < x < 0.8 compositions as evident from the PXRD and heat capacity measurements (Figure 20a,b). The authors also determined the phase diagram of this mixed system using the temperature dependence of the lattice microstrain extracted from the PXRD results (Figure 20c). Interestingly, for the highly mixed compositions, the microstrain value remained constant throughout the measured temperature range. This was attributed to the formation of the orientational glass phase caused by the halide disorder, which in turn distorts the octahedra and affects the MA cation motion. In a recent study, Naqvi et al. also observed local inhomogeneous environments of the MA cations in the mixed compositions of MAPb(Br1–xClx)3 using Raman spectroscopy.377 Note that a similar mechanism of the glassy phase formation was also proposed for the mixed A-site systems such as MA1–xDMAxPbBr3.164
Figure 20.
(a) PXRD pattern of the MAPb(Br0.5Cl0.5)3 composition showing a complete suppression of the phase transitions. (b) Temperature dependence of the heat capacity for the x = 0, 0.5, and 1 compositions. (c) Temperature–composition phase diagram of the MAPb(Br1–xClx)3 system. Symmetry notation: C - cubic, T - tetragonal, O - orthorhombic. Reprinted with permission from ref (381). Copyright 2021 American Chemical Society.
The MA cation dynamics in the mixed MAPb(Br0.4Cl0.6)3 compound was also studied in the aforementioned work by Selig et al.369 using the 2D IR spectroscopy and classical MD simulations. Similarly to the mixed MAPb(I0.4Br0.6)3 composition (Figure 19d), the MAPb(Br0.4Cl0.6)3 film sample also showed a substantial increase of the MA cation reorientation time to 5 ps in contrast to much shorter time scales observed for the end members (1.5 ps for x = 0, and 1.2 ps for x = 1).
5.2. Cs-Based Compounds
5.2.1. CsPb(I1–xBrx)3
Several studies revealed no miscibility gap in the mixed inorganic CsPb(I1–xBrx)3 perovskite system,232,384−386 although Wang et al. reported the onset of phase segregation under light illumination.387 Many works also reported that introduction of bromine suppresses the yellow phase and stabilizes the photoactive black phase of CsPbI3.232,384−386,388,389
Näsström et al.384 used in situ grazing-incidence wide-angle X-ray scattering and X-ray
fluorescence experiments to investigate structural phase transitions
and phase diagrams for a broad range of compositions of the CsPb(I1–xBrx)3 thin films obtained by inkjet printing and ordinary coating
(for Br-rich compositions). The compositions containing up to moderate
amounts of bromine (x ≤ 0.45) were found to
form the orthorhombic (Pnma) yellow δ-phase
at room temperature. This phase was converted to the cubic
black α-phase upon heating to high
temperatures (Figure 21a), and the temperature of this transition decreased significantly
with increasing x, indicating stabilization of the
cubic phase (Figure 21b). A subsequent cooling of the mixed compounds showed a contrasting
behavior compared to heating, as, instead of the yellow phase formation,
a phase transition sequence (cubic-tetragonal-orthorhombic) of the
black polymorph was observed (Figure 21c). Note that pure CsPbI3 perovskite was
an exception exhibiting an immediate conversion back to the yellow
phase (Figure 21c).
On the opposite side of the phase diagram, Br-rich compositions (x ≥ 0.85) already crystallized into the orthorhombic
(Pbnm) black γ-phase (Figure 21b) resulting in the same behavior during
heating and cooling. Interestingly, for the intermediate values of
mixing, the coexistence of the yellow and black phases was observed
at room temperature prior the heating run (Figure 21b). Note that similar phase transition temperatures
of the black polymorph were also observed for the CsPb(I2/3Br1/3)3 thin film by Breniaux et al.388 (Figure 21c). A gradual evolution of the phase transitions with
varying composition is in a sharp contrast to the MA-based systems,
where phase transitions are typically suppressed.164−166,336,381
Figure 21.
(a) Grazing-incidence wide-angle X-ray scattering pattern of the printed CsPb(I0.77Br0.23)3 thin film sample obtained on heating and then subsequent cooling runs. Red lines indicate the phase transitions. Temperature–composition phase diagram of the CsPb(I1–xBrx)3 system obtained on (b) heating and (c) subsequent cooling. Symmetry notation: C - cubic, T - tetragonal, O - orthorhombic. Green data points in (c) are taken from ref (388). Reprinted (adapted) with permission from ref (384). Copyright 2020 The Royal Society of Chemistry.
5.2.2. CsPb(Br1–xClx)3
This mixed system was investigated in the already discussed study by van de Goor et al.381 using PXRD and heat capacity measurements assisted by DFT calculations. In contrast to the hybrid MAPb(Br1–xClx)3 system (Figure 20), the authors did not observe lattice stabilization and formation of the orientational glass phase even for the highest mixing level (x = 0.5), as the system remained in the orthorhombic (Pnma) symmetry inherited from the end member compositions. The DFT calculations also revealed orthorhombic distortion and a narrow distribution of the Pb–halide–Pb angle, indicating preservation of the long-range order.
This may be related to the fact that in both CsPbBr3 and CsPbCl3, end members’ discernible octahedral distortions are evident within the orthorhombic Pnma structures, as indicated by the Pb–X–Pb angles distributed in the range of 157–165° and 160–169°, respectively.287,288 This observation stands in contrast to MAPb(Br1–xClx)3, wherein the energetically favored structural configuration is an average cubic one, characterized by a broad Pb–halide–Pb angle distribution that is also shifted toward the theoretically ideal cubic value of 180°. A contributing factor to the absence of long-range order may be attributed to the disparate degrees of distortion observed in nonmixed compositions. In MAPbBr3, for instance, all Pb–Br–Pb angles demonstrate a near-equivalence, fluctuating between 166.8° and 167.3°,269 while Pb–Cl–Pb angles in MAPbCl3 exhibit a widespread distribution in the range of 155–170°.390
5.3. MHy-Based Compounds
Unlike archetypal MAPbX3 and FAPbX3 compounds, MHyPbX3 (X = Br, Cl) 3D perovskites show exceptionally distorted and noncentrosymmetric framework consisting of alternating weakly and strongly distorted PbX6 octahedra.184,185 The MHyPbBr3 compound exhibits a single structural phase transition from the cubic to monoclinic phase at 418 K, while MHyPbCl3 also shows a single transition from the orthorhombic to monoclinic symmetry at 342 K (Figure 3). Note that the iodide analogue does not form a 3D perovskite and instead crystallizes into a 2D layered structure.391
5.3.1. MHyPb(Br1–xClx)3
Halide mixing effects in the MHy-based hybrid perovskites were recently studied by some of us using a suite of different experimental techniques including SCXRD, DSC, SHG, and DS.118 No evidence of the miscibility gap was observed in the whole range of mixing in agreement with the related MAPb(Br1–xClx)3 system.381 The SCXRD experiments of MHyPb(Br1–xClx)3 revealed that mixing does not change the polar symmetry, MHy cation arrangement, and framework distortions of the corresponding structural phases (see Figure 22a–f). The DSC experiments were used to follow the phase transition temperatures with mixing as summarized in the phase diagram presented in Figure 22g. Upon an increase of x, the stability region of the orthorhombic Pb21m phase gradually increased at the expense of the cubic Pm3̅m and monoclinic P21 phases.
Figure 22.
Crystal structure of the mixed MHyPb(Br1–xClx)3 system in the (a) cubic, (b) orthorhombic, and (c) monoclinic phases. Arrows represent electric dipole moments of MHy cations. (d) Single [100] layer of the PbX6 octahedra in the cubic phase. [100] layers of (e) less-distorted Pb(1)X6 and (f) highly distorted Pb(2)X6 octahedra in both polar orthorhombic and monoclinic phases. (g) Temperature–composition phase diagram of the MHyPb(Br1–xClx)3 system. Symmetry notation: C - cubic, O - orthorhombic, M - monoclinic. (h) Temperature dependence of the real part ε′ of the complex dielectric permittivity of the MHyPb(Br0.85Cl0.15)3 pellet sample. The dielectric response is dominated by the conductivity effects. Reprinted (adapted) with permission from ref (118). Copyright 2022 American Chemical Society.
Note that no lattice symmetrization (phase transition suppression) was observed in the mixed compositions of MHyPb(Br1–xClx)3 in contrast to the MA-based system (Figure 20c), where signatures of the glassy behavior were detected. In addition, no substantial broadening of the transition anomalies was observed in MHyPb(Br1–xClx)3 upon mixing. These results indicate that introduction of Cl– anions did not destroy the long-range order in this system, and thus lattice symmetrization and formation of the dipolar glass phase cannot be expected. This was also evident in the dielectric responses of the mixed compounds, which were dominated by the conductivity processes (Figure 22h). The absence of the lattice symmetrization may be explained by the already significantly deformed and rigid inorganic framework at room temperature. Note that a similar behavior was also observed for the related CsPb(Br1–xClx)3 system.381
5.3.2. MHyPb(Br1–xIx)3
In the same study,118 an attempt was made to incorporate iodine in the structure of MHyPbBr3 perovskite using solvent synthesis. A MHyPb(Br0.93l0.07)3 composition of 3D perovskite was obtained as revealed by energy-dispersive X-ray spectroscopy analysis. SCXRD and DSC experiments showed that this compound undergoes a structural phase transition at 390 K from the cubic Pm3̅m to monoclinic P21 phase, both inherited from pure MHyPbBr3.
5.4. Summary
Several important observations can be made from the reviewed results of mixed-halide lead halide perovskites. The mixed compositions with iodine, which has the largest ionic radius, may result in a miscibility gap, which is especially pronounced for the MA-based compounds obtained using the solvent synthesis. Interestingly, it seems that this gap can be reduced or fully suppressed using mechanochemical synthesis. There is no miscibility gap for the mixed Br/Cl compounds allowing one to fully chart the phase diagrams.
The temperature–composition phase diagrams of the mixed MAPb(I1–xBrx)3, MAPb(Br1–xClx)3, CsPb(I1–xBrx)3, and MHyPb(Br1–xClx)3 systems are summarized in Figure 23 revealing vastly different behaviors. MAPb(I1–xBrx)3 compounds obtained using solvent synthesis exhibited a wide region of phase separation, while well-mixed compositions outside this gap showed a significant lowering of the transition temperatures upon mixing367 (Figure 23a). This indicates increased stability of the cubic phase due to the disorder induced by random halide distribution, while the dielectric response obtained for the high levels of mixing showed signatures of the glassy phase formation.365 The phase diagram of the MAPb(Br1–xClx)3 system, which has no miscibility gap, revealed a complete suppression of the phase transitions and symmetrization of the lattice for the highly mixed compounds (Figure 23b).381 For these compositions, the formation of the orientational glass phase was also observed.
Figure 23.
Comparison of the temperature–composition phase diagrams of the mixed-halide (a) MAPb(I1–xBrx)3, (b) MAPb(Br1–xClx)3, (c) CsPb(I1–xBrx)3 (black phase except for x = 0), and (d) MHyPb(Br1–xClx)3 systems. Symmetry notation: C - cubic, T - tetragonal, O - orthorhombic, M - monoclinic. (a) Adapted with permission from ref (367). Copyright 2019 The Royal Society of Chemistry. (b) Adapted with permission from ref (381). Copyright 2021 American Chemical Society. (c) Adapted with permission from ref (384). Copyright 2020 The Royal Society of Chemistry. (d) Adapted with permission from ref (118). Copyright 2022 American Chemical Society.
The phase diagrams and effects of mixing are qualitatively different when mixed compositions are obtained from the end members, which exhibit distorted lower symmetry structures already at room temperature such as Cs-based (orthorhombic) and MHy-based (monoclinic) compounds. In both cases, no suppression of the phase transitions and lattice symmetrization were observed, as systems gradually evolved with halide mixing maintaining sharp phase transition anomalies (Figure 23c,d). This indicates preservation of the long-range order and absence of the glassy phase. Such a behavior may be explained by a much higher stability of the distorted phases (much higher transition temperatures) compared to the MA-based perovskites.
6. Simultaneous A- and X-Site Mixing in 3D Perovskites
Here, we discuss how simultaneous mixing at the A- and X-sites affects the structural phase transitions and cations dynamics in 3D lead halide perovskites. The reported studies concentrate on the highly promising FA1–xMAxPb(I1–yBry)3 and FA1–xCsxPb(I1–yBry)3 systems, where FA is the dominant molecular cation.
6.1. FA1–xMAxPb(I1–yBry)3
Greenland
et al. used temperature-dependent PXRD to study the structural phase
transitions in the FA0.85MA0.15Pb(I0.85Br0.15)3 thin film sample.392 The phase transition from the cubic
to tetragonal (P4/mbm) symmetry was observed at 260 K (Figure 24a), which is about 15 K lower compared to
pure FAPbI3 (Figure 3). In addition, the lowering of the phase transition temperature
seems to be more pronounced compared to the FA0.85MA0.15PbI3 composition, where mixing was done only
at the A-site (see phase diagram in Figure 5). The authors also claim observation of
the phase transition to the γ-phase, which occurred at 90 K
instead of about 150 K observed for pure FAPbI3. Interestingly,
upon heating, the PXRD peaks of this low-temperature phase persisted
up to 220 K, indicating a very broad hysteresis (also visible in Figure 24a).
Figure 24.
(a) Temperature
dependence of the reduced lattice parameter ratio
of the mixed FA0.85MA0.15Pb(I0.85Br0.15)3 system obtained
on cooling and heating. (b) Trajectories of the MA and FA cations
in the mixed FA0.875MA0.125Pb(I0.875Br0.125)3 perovskite obtained by the ab initio MD. Arrows indicate average direction of the MA
cations. (c) Simulated vector autocorrelation functions of the MA
and FA cations (insets) in the same solid solution (solid curves)
compared with simulations obtained in the parent MAPbBr3 and FAPbI3 compounds (dashed curves). (a) Adapted with
permission from ref (392). Copyright 2020 Wiley-VCH. (b, c) Reprinted (adapted) with permission
from ref (393). Copyright
2022 The Royal Society of Chemistry.
Menéndez-Proupin et al. employed ab initio MD to study molecular cation dynamics in the FA0.875MA0.125Pb(I0.875Br0.125)3 perovskite.393 The density maps of the C and N atoms revealed that the FA cations rotate freely in the lattice, while the MA cations have a unique orientation during the calculated trajectory (Figure 24b), which seems to be independent of the number of neighboring Br– anions. The authors also calculated the vector autocorrelation function for both molecular cations showing that in the highly mixed composition the dynamics of both cations are slower (Figure 24c). Note that the same behavior was obtained in the mixed FA0.9Cs0.1PbI3 (Figure 15c)226 and MAPb(I0.5Br0.5)3 systems (Figure 19d).369
Johnston et al. used QENS to study molecular cation dynamics in the highly mixed FA0.8MA0.15Cs0.05Pb(I1–yBry)3 (y = 0, 0.1, 0.15, and 0.2) system with an addition of a small amount of Cs+ cations at the A-site.394 The temperature-dependent mean-squared displacement data revealed a phase transition to the low-temperature phase at about 100 K independent of the halide mixing. This result does not agree with the phase diagram of the mixed MA1–xFAxPbI3 system (y = 0) (Figure 5), suggesting that Cs+ cations may significantly affect the phase transitions. The authors also observed the strongest suppression of the molecular cation dynamics (particularly FA) for the y = 0.15 composition.
Xie et al.395 grew high-quality single crystals of (FAPbI3)1–x(MAPbBr3)x. They observed that the best composition for a black phase without segregation is x = 0.1–0.15 and that the incorporation of MA is essential for stabilization of the black phase from the thermodynamics point of view. Namely, incorporation of MA tunes the tolerance factor toward the ideal value of 1 and lowers the Gibbs free energy via unit cell contraction and cation disorder. On the other hand, Br incorporation also stabilizes the black phase but from the kinetics point of view, because this anion controls the crystallization kinetics and reduces defect density.
6.2. FA1–xCsxPb(I1–yBry)3
Barrier et al.396 used synchrotron X-ray diffraction to probe the evolution of crystal structure across the tetragonal-cubic phase transition for FA1–xCsxPb(I1–yBry)3 thin films with x = 0.17–0.4 and y = 0.05–0.3. By analysis of the octahedral tilt angle, they observed that the cubic-tetragonal phase transition occurred over temperature range on the order of 40 °C. This behavior proved that the samples were poorly mixed and contained coexisting cubic and tetragonal regions. It has been discussed that the compositional heterogeneity can have two origins. The first one is due to the thermodynamic separation of the mixed perovskite into two or more distinct phases driven by minimization of the free energy, whereas the second one likely results from kinetics due to the solution-processing methods used to fabricate thin films. The authors have made no attempt to distinguish between these components but emphasized that it is important to take into account the presence of heterogeneity, since it may lead to chemical, structural, and electronic property heterogeneity on multiple length scales.
7. B-Site Mixing in 3D Perovskites
The number of works reporting on how the B-site mixing affects the structural and dynamic properties of hybrid perovskites is highly limited. Here, we discuss two MAPbI3 and FAPbI3 perovskite systems, where lead is partially replaced by tin.
7.1. MAPb1–xSnxI3
The room-temperature structural properties of the mixed MAPb1–xSnxI3 system were investigated by Kanatzidis group in two subsequent studies.93,94 The authors observed a full compatibility of both parent compounds, as no solubility limit in the mixed perovskites was detected. For all compositions, the room-temperature crystal structure adopted tetragonal symmetry. For the Sn-rich compositions (x ≥ 0.5), the P4mm space group was obtained, while the x < 0.5 compounds including MAPbI3 were indexed using the noncentrosymmetric I4cm space group instead of the expected centrosymmetric I4/mcm. The absence of the centrosymmetry was not discussed or proved by the authors making this assignment questionable.
7.2. FAPb1–xSnxI3
Parrott et al.89 reported a more detailed investigation of the mixing effects in the related FAPb1–xSnxI3 perovskites, where no solubility limit was also observed. The authors used temperature-dependent PL and light absorption experiments to probe how introduction of tin affects the phase transitions (Figure 25a) and optical properties such as optical band gap (Figure 25b). Clear anomalies were observed in the measured data at the low-temperature phase transition point allowing construction of the phase diagram (Figure 25c). The observed transition likely corresponds to the tetragonal-orthorhombic symmetry change based on the symmetry of the parent compounds at such temperatures (Figure 3).170
Figure 25.
(a) Temperature dependence of the PL signal of FAPb0.55Sn0.45I3 perovskite. (b) Temperature dependence of the optical band gap determined from the PL data of the same sample. Arrow marks the phase transition point. (c) Temperature–composition phase diagram of the tetragonal–orthorhombic phase transition in FAPb1–xSnxI3 system. Reprinted (adapted) with permission from ref (89). Copyright 2018 Wiley-VCH.
In the aforementioned piezoresponse force microscopy study by Ahmadi et al.,339 the authors observed that, similarly to MA0.15FA0.85PbI3, the FAPb0.85Sn0.15I3 compound also exhibits ferroelectric domain-like structures, however, much weaker than expected for classical ferroelectric domains.
Further studies with the emphasis on the structural phase transitions and dynamics are necessary to provide a more detailed picture of the B-site mixing effects.
8. Halide, Metal, and Cation (A-, A′-, A″-Site) Mixing in Lower-Dimensional Perovskites and Related Compounds
Here, we discuss how mixing affects the structural and dynamic properties of low dimensional lead halide perovskites. Note that the number of works reporting on these effects is significantly lower compared to the 3D APbX3 compounds.
8.1. CHA2Pb(Br1–xIx)4
Ye et al.397 used DSC, SHG, light absorption, dielectric and hysteresis loop measurements to probe how introduction of iodine up to x = 0.18 affects the phase transitions, band gap, ferroelectric and NLO properties of 2D layered perovskite comprising cyclohexylammonium (CHA) cations (Figure 26a). On cooling, pure CHA2PbBr4 undergoes a ferroelectric phase transition from the Cmca to the Cmc21 phase at 363 K. The authors showed that mixing increased the phase transition temperature to 378 and 380 K for the x = 0.11 and 0.18 compositions, respectively. In addition, mixing also increased intensity of the SHG signal (Figure 26b) and narrowed the band gap from 3.05 eV for CHA2PbBr4 to 2.74 eV for CHA2Pb(Br0.82I0.18)4 while maintaining the phase transition. Interestingly, the dielectric anomaly associated with the phase transitions showed no broadening upon mixing (Figure 26c) in contrast to the majority of the 3D APbX3 perovskites (see e.g. Figure 19c). In addition, the ferroelectric properties also remained in the mixed compositions (Figure 26d) showing that the CHA cation ordering is decoupled from the halide mixing.
Figure 26.
(a) Crystal structure of pure CHA2PbBr4 compound in the ferroelectric (FE) and paraelectric (PE) phases. Cation disorder is present in the PE phase. (b) Temperature dependence of the (b) SHG response and (c) real part ε′ of the complex dielectric permittivity (1 MHz) of the mixed CHA2Pb(Br1–xIx)4 system (x = 0, 0.11, and 0.18). (d) Ferroelectric hysteresis loops of the same compounds obtained at room temperature. Reprinted (adapted) with permission from ref (397). Copyright 2016 Wiley-VCH.
A full range of compositions (0 ≤ x ≤ 1) of the same system was studied by Yangui et al.37 It was shown that a composition-induced phase transition from the Cmc21 to Pbca phase occurs for 0.5 < x < 0.6 at room temperature. The change in composition also led to the appearance of white-light emission for x < 0.5.
8.2. (C9H19NH3)2PbI2Br2
Abid et al.398 studied phase transition in (C9H19NH3)2PbI2Br2 solid solution, which crystallizes in a 2D perovskite structure. Using DSC, PXRD, and IR spectroscopy, they showed that the phase transition occurs around 230 K. The crystal structure below 230 K was not solved, but spectroscopic studies revealed that the phase transition is related to a decreased conformational disorder of the methylene units in the alkyl chains.
8.3. Br-PEA2Pb(BrxCl1–x)4
Pareja-Rivera et al.40 studied the effect of mixing at the X-site on the PL properties of 2D perovskite comprising 4-bromophenethylammonium (Br-PEA) cation. No solubility limit was observed, and the crystal structure of this system was found to be orthorhombic Fmm2 in the whole range of mixing. Analysis of the structural data revealed a much larger distortion of the inorganic layers for x = 0 compared to x = 1. Lattice parameters b and c increased with increasing Br content, whereas an opposite behavior was observed for the a parameter. Mixing at the X-site had a pronounced effect on the band gap and PL, as the band gap narrowed with increasing x, while PL for the x = 1 composition was dominated by a sharp violet emission with the peak maximum near 420 nm. This emission shifted to the UV range with decreasing x. When x decreased to about 0.6, a new broadband emission appeared near 500 nm, and its intensity increased on further lowering of x.
8.4. PEA2Pb(BrxCl1–x)4and PEA2Pb(BrxI1–x)4
Mixing effects at the X-sites were also reported for 2D perovskites comprising the phenethylammoium (PEA) cation.39,399,400 Similarly to the Br-PEA analogue discussed above, Cai et al. claimed no solubility limit in the PEA2Pb(BrxCl1–x)4 system, but the crystal symmetry was lower, triclinic P1̅.39
The XRD and DFT calculations revealed phase separation for the x = 0.25 composition of the mixed PEA2Pb(BrxI1–x)4 analogue, while no phase separation was observed for x ≥ 0.5.399 For x = 0.75, two PL bands were observed and only one for x = 0.25 and 0.5, which was attributed to different crystal configurations simultaneously present in the prepared film. Another study of PEA2Pb(BrxI1–x)4 system showed the presence of two (002) diffraction peaks for the x = 0.79 composition, suggesting an immiscible system with two discrete halide phases.400
8.5. HA2Pb(BrxI1–x)4and HA2FAPb2(BrxI1–x)7
A mixing effect at the X-sites was also reported for HA2Pb(BrxI1–x)4 and HA2FAPb2(BrxI1–x)7 compounds (HA = n-hexylammonium).400 The situation for these systems is very similar to that observed for PEA2Pb(BrxI1–x)4,400 as these halides do not show full solubility, and there is a clear crossover from near-pure-I to near-pure-Br phases for x = 0.84–0.88.400 The authors also showed that these systems may show photoinduced phase separation over hundreds of ms for focused laser beam and over minutes under the diffuse UV light.
8.6. Et3PrNPb(Br1–xIx)3
Shao et al.124 reported investigation of the mixing effects at the X-site of 1D hybrid lead halide comprising globular-shaped triethylpropylammonium (Et3PrN) cations. The authors found no solubility limit for this mixed halide system.
The DSC experiments showed that the phase transition is weakly affected by mixing as its temperature gradually decreased from 480 K for x = 0 to 445 K for x = 1. The entropy associated with the phase transition was the largest for pure Et3PrNPbBr3 (43.54 J mol–1 K–1) and decreased to 32.36 J mol–1 K–1 for Et3PrNPbI3 indicating a lower degree of ordering. The mixing with iodine also results in the shift of the band gap from 3.50 eV for x = 0 to 2.93 eV for x = 1.
8.7. t-BA2Pb(Br1–xIx)4
The mixed halide system containing tert-butylammonium (t-BA) branched spacer cation was recently studied by Ovčar et al.36 The authors showed that the end members crystallize into the polar P21 structures: 1D for the iodide (t-BAPbI3), and layered 2D nonperovskite for the bromide (t-BAPb2Br5). Interestingly, the perovskite architecture was found to be stabilized in the mixed halide t-BA2PbBr2I2 system (2D layered perovskite structure, monoclinic P21/c space group), however, at the expense of the polar symmetry.
8.8. (3AMP)x(4AMP)1–x(FA)y(MA)1–yPb2Br7
Mixing of organic cations in a DJ type structure of (3AMP)x(4AMP)1–x(FA)y(MA)1–yPb2Br7 (3AMP = 3-(aminomethyl)piperidinium, 4AMP = 4-(aminomethyl)piperidinium) was studied by Mao et al.138 This system showed a composition-induced phase transition from the monoclinic Cm symmetry for x = 1 and y = 1 to the monoclinic Pc phase for the (3AMP)0.5(4AMP)0.5FAPb2Br7,(3AMP)0.5(4AMP)0.5FA0.5MA0.5Pb2Br7, (4AMP)FA0.5MA0.5Pb2Br7, (4AMP)FAPb2Br7, and (3AMP)MAPb2Br7 compositions. The monoclinic Cc symmetry was reported for x = 0 and y = 0 as well as for x = 0.5 and y = 0. It was found that these structural changes have a significant impact on the Pb–Br–Pb bond angle and the band gap.
8.9. (BA0.5PEA0.5)2MAPb2Br7
Mixing of organic spacer cations was also studied in an RP type structure of (BA0.5PEA0.5)2MAPb2Br7.136 The structure of BA2MAPb2Br7 was refined in the orthorhombic Ama2 space group, and PEA2MAPb2Br7 was reported to be isostructural to PEA2MAPb2I7136 with the triclinic space group P1̅.401 The XRD results showed that (BA0.5PEA0.5)2MAPb2Br7 has a distinct structure from its parent compounds (triclinic P1).
8.10. [AAxIdPA1–x]2MAn-1PbnI3n+1 (n = 2–4)
Another reported RP system with mixing at the A′-sites is [AAxIdPA1–x]2MAn-1PbnI3n+1 (n = 2–4) family of compounds (AA = allylammonium, IdPA = iodopropylammonium).402 The most important conclusion from studies of this system is that, whereas the end members crystallize in the centrosymmetric structures, the mixed A′ phases are acentric. For instance, AA2MA2Pb3I10 and IdPA2MA2Pb3I10 crystallize in the P21/c and P2/c structure, respectively, while [AAxIdPA1–x]2MA2Pb3I10 phases are described by the noncentrosymmetric Pc space group.402
8.11. (BA)2(MA1–xEAx)2Pb3I10
Mixing of organic cations in RP phases can also be realized at the “perovskitizer” sites, as demonstrated by Fu et al. in a study of (BA)2(MA1–xEAx)2Pb3I10.140 The end members BA2EA2Pb3I10 and BA2MA2Pb3I10 crystallize in the Cmc21 and C2cb space groups, respectively, and, due to larger ionic radius of EA, BA2EA2Pb3I10 feature a higher degree of structural deformation. The authors showed that the symmetry of the mixed cation phases decreased to monoclinic Cc, and the average equatorial Pb–I–Pb bond angle of the inner layers strongly decreased with increasing EA fraction indicating higher structural distortion.
8.12. (PEA)2(MA1–xGAx)2Pb3I10
Another example of the RP phase with mixed cations at the “perovskitizer” sites is the (PEA)2(MA1–xGAx)2Pb3I10 system studied by Ramos-Terrón et al.141 In this work, thin films up to x = 0.5 were prepared, which is a much higher incorporation fraction compared to the 3D analogues (e.g., 25% in MA1–xGAxPbI369). The authors observed shifts of diffraction peaks with increasing x due to enlargement of the unit cell. When the GA content approached 30%, a clear change in the width of the diffraction peaks was observed. This behavior was attributed to a sudden change in the arrangement of the mixed cations. To obtain information on the degree of lattice distortion induced by incorporation of GA, the DFT calculations were performed for the x = 0.25 and x = 0.5 compositions. The obtained results revealed low distortion of the inorganic layers on GA incorporation, even for a high GA content.
8.13. (BA)2(MA1–xCsx)Pb2Br7
Mixing at the “perovskitizer” sites was also reported by Ma et al. for (BA)2(MA1–xCsx)Pb2Br7.403 The authors grew single crystals for x = 1, 0.93, 0.79, and 0.66 and showed that incorporation of MA led to a decrease of the phase transition temperature from 411 K for x = 1 to 399, 383, and 379 K for x = 0.93, 0.79, and 0.66, respectively. For all solid solutions, the ferroelectricity of the low-temperature phase was confirmed by observation of the polarization hysteresis loops. The incorporation of MA led to narrowing of the band gap from 2.7 eV for x = 1 to 2.43 eV for x = 0.66. Most interestingly, the introduction of MA also resulted in a very large, at least 10-fold, enhancement of the photoactivity of this layered perovskite. This effect was attributed to the reduction of the exciton binding energy and enhanced ferroelectric polarization. As a result, the photoelectric activity of the mixed-cation crystals (x = 0.66) exceeded that of previously reported self-driven detectors based on hybrid lead halide perovskites.
8.14. PEA2Pb1–xSnxBr4
Sui et al.127 reported investigation of the mixing effects at the B-site of 2D hybrid lead halide comprising aromatic PEA cations. This system shows a full solubility, while maintaining the same triclinic P1 crystal structure with decreased lattice parameters due to smaller ionic radius of Sn2+ (93 pm) compared to Pb2+ (120 pm). The smaller ionic radius of Sn2+ was also shown to be responsible for a shift of the band gap and PL peak to lower energies with increasing x.
In summary, literature data show that ion mixing in low-dimensional perovskites may significantly tune optoelectronic properties and lead to new symmetry of the mixed phases and loss of the inversion center.
9. Conclusions and Outlook
In this review, we provided a systematic and comprehensive picture of the mixing effects on the structural and dynamic properties of 3D and low-dimensional lead halide perovskites. The following important generalizations can be drawn from the reviewed studies:
(i) Frequently, the extent of mixing is limited by phase separation, which is especially pronounced, if the parent compounds have structures of different topologies. The phase separation directly affects the phase diagrams and material properties and thus must be carefully considered and determined. In general, for lead halide perovskites, kinetics often have a strong impact on the phases that form. This problem is especially important in the fabrication of thin films, since the spin-coating process often leads to pronounced heterogeneity.361,395,396 The presence of nanoscale local compositional variations may lead to structural and optoelectronic property heterogeneity on multiple length scales.361,396 It also complicates the phase transition behavior since slight variations in chemical composition can result in a different crystallographic phase being thermodynamically favorable and shift the transition temperature, leading to smearing of the transition anomaly.361,396 It has been shown that the A-site segregation can be often mitigated by annealing the samples either long enough or at a sufficiently high temperature.361,396 Another way is to perform rapid crystallization by antisolvent dripping.404,405 Trapping of the metastable phases can also be done by fast quenching.264
(ii) The mixing tends to partially or completely suppress the structural phase transitions resulting in the stabilization of the cubic phase and thus symmetrization of the crystal lattice, which is related to the improved stability, performance, and functionality of the mixed compounds.
(iii) Symmetrization of the lattice is frequently accompanied by signatures of the orientational (dipolar) glass phase, which occurs due to the multivalley potentials and frustrated interactions introduced by mixing and mediated by the local lattice deformations.
(iv) The symmetrization effect seems to be less pronounced or absent, if parent compounds exhibit highly stable low-symmetry phases at room temperature (e.g., Cs- and MHy-based perovskites). In this case, the evolution of the phase transitions with mixing is continuous, and signatures of the glassy phase are absent.
(v) The dynamics of molecular cations typically become slower upon mixing. This is especially pronounced at low temperature, where more bulky guest cations tend to freeze (e.g., DMA) and cause local deformation of the crystal lattice, which in turn increases the rotation barrier and hinders the dynamics of the smaller cations (e.g., MA).
(vi) Mixing in low-dimensional lead halide perovskites shows interesting behavior related to the occurrence of new crystal symmetries and loss of the inversion center. However, mixing in these compounds is significantly less studied compared to the 3D perovskites and therefore requires more attention. This is especially important in the context of ferroelectric low-dimensional perovskites, which upon mixing may form a ferroelectric relaxor phase instead of the dipolar glass.
Lastly, we identify other new promising systems and directions for further studies of the mixing effects on the structural and dynamic properties of lead halide and related perovskites. Such systems include hollow perovskites,72 deficient perovskites,406 and quasi 3D perovskites,407 all of which exhibit a substantial degree of disruption of the inorganic framework. It can be expected that in such compounds the molecular cation frustration would be even higher, as both organic and inorganic sublattices would significantly contribute to disorder.
Mixing in the newly discovered aziridinium-based 3D perovskites237,298,408,409 is also of high interest, especially from the stability point of view. As these perovskites have a 3D topology and high stability of the cubic phase, their solid solutions with high fractions of MA and FA should be feasible.
The B-site mixing in 3D perovskites is also significantly understudied compared to the A- and X-site mixing despite seemingly interesting effects on the structural phase transitions. In addition, mixing at the B-site should also significantly affect the off-centering dynamics of the inorganic framework,342,410 which should be directly visible in the low-temperature dielectric response of these materials. This is also related to lattice fluctuations resulting in low-symmetry correlated octahedral distortions observed in lead halide perovskites. It is very likely that mixing (especially at the B-site) would significantly affect such octahedral rotational instabilities. However, due to the requirement of advanced structural characterization techniques, such studies of the mixed systems are very rare.
As discussed, a major area of research in lead halide perovskites is related to the stabilization of the black photoactive phases of FAPbI3 and CsPbI3 compounds by utilizing several strategies including ion mixing—a topic which is out of scope for this review. However, the stabilization by mixing should inevitably affect the structural phase transitions and dynamics in the stabilized black phases of FAPbI3 and CsPbI3. One could expect a connection between the induced phase stability and suppression of the structural phase transitions in the black phases of these compounds. For example, Marshall et al. demonstrated a successful stabilization of the black phase of CsPbI3 by DMA cations reaching incorporation fractions up to 25%,411 while similar values of DMA proved to be sufficient to fully stabilize the cubic phase in the MA1–xDMAxPbBr3 system.164 Interestingly, a recent NMR study found that the mixed Cs1–xDMAxPbI3 perovskite phases only form when they are kinetically trapped by rapid antisolvent-induced crystallization.404 Of potential interest is also doping with Rb, which can be incorporated into the CsPbBr3, the black phase of CsPbI3, and CsPb(I/Br)3 solid solutions.412
As interesting example of the black phase stabilization was also recently reported by Duijnstee et al., where a small amount (x = 0.005) of tetrahydrotriazinium (THTZ-H) molecular cation was incorporated in the structure of FAPbI3.413 The samples were obtained from solutions containing methylenediammonium dichloride or hexamethylenetetramine, which then formed THTZ-H during the subsequent reactions with FA. Remarkably, the authors reported that such a mixing leads to an impressive enhancement of the black phase stability of FAPbI3 lasting for more than one year under ambient conditions. It can be expected that such a pronounced effect of THTZ-H on the black phase stabilization could also significantly alter the structural phase transitions and FA cation dynamics in α-FAPbI3.
The mixing effects and the dielectric response are also significantly understudied in the related lead-free families of hybrid materials such as tin halides,297,414,415 antimony halides,416,417 bismuth halides,417−421 tellurium halides,422 and double (e.g., Ag/Bi) perovskites.423−427 Many of these compounds undergo structural phase transitions related to the ordering-disordering phenomena and ferroelectric properties. Thus, one can expect a formation of similar frustrated phases in these materials upon mixing.
We note that, apart from dipolar glass, mixing-induced frustration may also lead to other exotic disordered phases. From the applications point of view, of particular interest is the relaxor phase,58 which is rather common in the mixed classical inorganic compounds. However, the formation of the relaxor phase requires mixing of ferroelectrics, while the ferroelectricity in 3D lead halide perovskites is questionable.183 On the other hand, numerous lower-dimensional perovskites are ferroelectric and antiferroelectric (see Table 2), and thus their mixing holds a significant potential for formation of relaxor and other complex frustrated phases.
There is presently a lot of interest in chiral layered lead halides, in which employment of a chiral organic cation leads to transfer of the chirality into the achiral inorganic framework.428−435 Due to the intrinsic noncentrosymmetry of chiral cations, chiral lead halides show unique and highly attractive features such as optical rotation, circular dichroism, circularly polarized PL, polarization-dependent SHG, ferroelectricity, bulk photovoltaic effect, chirality-induced spin selectivity, and topological quantum properties.428−434 Furthermore, the strong spin–orbit coupling of metal and halide ions imparts significant Rashba-Dresselhaus spin-splitting of otherwise 2-fold spin-degenerate electronic bands.430,431,436 We are not aware of any report on the mixing effect on phase transitions and lattice dynamics in chiral lead halides. However, there are some examples of mixed chiral hybrid perovskites reporting a significant tuning of optical properties upon mixing at the B- and X-sites.433,437,438 Thus, a substantial effect on the structural phase transitions and dynamics can be also expected.
Another intriguing feature of certain mixed inorganic ferroelectrics is the presence of a morphotropic phase boundary, where the crystal structure changes abruptly upon mixing, and the electromechanical (piezoelectric) properties become maximized.439 The observation of such a phase boundary in mixed lead halide perovskites would undoubtedly open new research and application frontiers of these materials.
Acknowledgments
This project has been funded by the Research Council of Lithuania (LMTLT) (Agreement No. S-MIP-22-73). The authors thank the anonymous reviewers for insightful suggestions.
Glossary
Abbreviations
- AA
allylammonium
- i-AM
isoamylammonium
- ACE
acetamidinium cation
- i-BA
isobutylammonium
- AM
amylammonium
- IM
imidazolium cation
- AZR
aziridinium
- i-PA
isopropylammonium
- BA
butylammonium
- MA
methylammonium cation
- BPA
3-bromopropylammonium
- MACH
cyclohexanemethylammonium
- Br-PEA
4-bromophenethylammonium
- MHy
methylhydrazinium cation
- BZA
benzylammonium
- PA
propylammonium
- CHA
cyclohexylammonium
- PEA
phenethylammoium
- CPA
3-chloropropylammonium
- t-ACH
4-aminomethyl-1-cyclohexanecarboxylate
- DMA
dimethylammonium cation
- t-BA
tert-butylammonium
- EA
ethylammonium cation
- TEA
thioethylammonium cation
- Et3PrN
triethylpropylammonium
- THTZ-H
tetrahydrotriazinium cation
- FA
formamidinium cation
- TZ
1,2,4-triazolium
- GA
guanidinium cation
- 3AMP
3-(aminomethyl)piperidinium
- HA
n-hexylammonium
- 4AMP
4-(aminomethyl)piperidinium
- HY
hydrazinium cation
- 4IBA
4-iodobutylammonium
- HEA
hydroxyethylammonium cation
- ACI
alternating cations in the interlayer space
- NMR
nuclear magnetic resonance
- DFT
density functional theory
- PL
photoluminescence
- DJ
Dion–Jacobson
- PXRD
powder X-ray diffraction
- DS
dielectric spectroscopy
- QENS
quasi elastic neutron scattering
- DSC
differential scanning calorimetry
- RP
Ruddlesden–Popper
- IR
infrared
- SCXRD
single crystal X-ray diffraction
- MD
molecular dynamics
- SHG
second-harmonic generation
- MC
Monte Carlo
- XRD
X-ray diffraction
- NLO
nonlinear optical
Biographies
Mantas Šimėnas received his Ph.D. degree in Physics from Vilnius University (Lithuania). Afterwards, he was a research fellow in magnetic resonance at University College London (UK). Currently, he is a professor in physics at the Faculty of Physics at Vilnius University focusing, among other topics, on the solid state physics, structural phase transitions, dielectric response, and dynamic phenomena of hybrid perovskites and related materials.
Anna Ga̧gor received her Ph.D. degree in Physics from the Institute of Low Temperature and Structure Research, Polish Academy of Sciences (Poland). Since then, she has been associated with the Institute, where she currently holds the position of Associate Professor and serves as the Head of the Structure Research Department. She is a member of the Polish Crystallographic Society and the Commission on Inorganic and Mineral Structures associated with the International Union of Crystallography. Her research focuses on the structural analysis of crystalline materials using X-ray diffraction, including the investigation of phase transitions, structure–property relationships in organic–inorganic hybrids, and ferroic and multiferroic materials.
Ju̅ras Banys obtained his Ph.D. degree at the Faculty of Physics at Vilnius University (Lithuania). He completed his postdoc training in solid state physics at Oxford University and Leipzig University. Currently, he is a professor in physics at Vilnius University. Ju̅ras is the world-leading expert in the dielectric characterization of the structural phase transitions, relaxor and dipolar glass phases, and lattice dynamics in mixed inorganic perovskites. He is best known for the development and application of the ultrabroadband dielectric spectroscopy to characterize different perovskite systems.
Mirosław Ma̧czka received his Chemistry diploma from the Wrocław University of Science and Technology (Poland) and his Ph.D. degree in Chemistry from the Institute of Low Temperature and Structure Research, Polish Academy of Sciences (Poland). He is currently a Professor in Chemistry at the Institute of Low Temperature and Structure Research. His research focuses broadly on hybrid perovskites, optoelectronic materials, synthesis science, and vibrational spectroscopy.
The authors declare no competing financial interest.
References
- Kojima A.; Teshima K.; Shirai Y.; Miyasaka T. Organometal Halide Perovskites as Visible-Light Sensitizers for Photovoltaic Cells. J. Am. Chem. Soc. 2009, 131, 6050–6051. 10.1021/ja809598r. [DOI] [PubMed] [Google Scholar]
- Liu M.; Johnston M. B.; Snaith H. J. Efficient Planar Heterojunction Perovskite Solar Cells by Vapour Deposition. Nature 2013, 501, 395–398. 10.1038/nature12509. [DOI] [PubMed] [Google Scholar]
- Park N.-G. Organometal Perovskite Light Absorbers Toward a 20% Efficiency Low-Cost Solid-State Mesoscopic Solar Cell. J. Phys. Chem. Lett. 2013, 4, 2423–2429. 10.1021/jz400892a. [DOI] [Google Scholar]
- Shang Y.; Liao Y.; Wei Q.; Wang Z.; Xiang B.; Ke Y.; Liu W.; Ning Z. Highly Stable Hybrid Perovskite Light-Emitting Diodes Based on Dion-Jacobson Structure. Sci. Adv. 2019, 5, eaaw8072 10.1126/sciadv.aaw8072. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang X.; Liu H.; Wang W.; Zhang J.; Xu B.; Karen K. L.; Zheng Y.; Liu S.; Chen S.; Wang K.; Sun X. W. Hybrid Perovskite Light-Emitting Diodes Based on Perovskite Nanocrystals with Organic-Inorganic Mixed Cations. Adv. Mater. 2017, 29, 1606405 10.1002/adma.201606405. [DOI] [PubMed] [Google Scholar]
- Kim Y.-H.; Cho H.; Lee T.-W. Metal Halide Perovskite Light Emitters. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 11694–11702. 10.1073/pnas.1607471113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Quan L. N.; García de Arquer F. P.; Sabatini R. P.; Sargent E. H. Perovskites for Light Emission. Adv. Mater. 2018, 30, 1801996 10.1002/adma.201801996. [DOI] [PubMed] [Google Scholar]
- Jeong J.; Kim M.; Seo J.; Lu H.; Ahlawat P.; Mishra A.; Yang Y.; Hope M. A.; Eickemeyer F. T.; Kim M.; et al. Pseudo-Halide Anion Engineering for α-FaPbI3 Perovskite Solar Cells. Nature 2021, 592, 381–385. 10.1038/s41586-021-03406-5. [DOI] [PubMed] [Google Scholar]
- Yoo J. J.; Seo G.; Chua M. R.; Park T. G.; Lu Y.; Rotermund F.; Kim Y.-K.; Moon C. S.; Jeon N. J.; Correa-Baena J.-P.; et al. Efficient Perovskite Solar Cells via Improved Carrier Management. Nature 2021, 590, 587–593. 10.1038/s41586-021-03285-w. [DOI] [PubMed] [Google Scholar]
- Zhou Y.; Zhou Z.; Chen M.; Zong Y.; Huang J.; Pang S.; Padture N. P. Doping and Alloying for Improved Perovskite Solar Cells. J. Mater. Chem. A 2016, 4, 17623–17635. 10.1039/C6TA08699C. [DOI] [Google Scholar]
- Ono L. K.; Juarez-Perez E. J.; Qi Y. Progress on Perovskite Materials and Solar Cells with Mixed Cations and Halide Anions. ACS Appl. Mater. Interfaces 2017, 9, 30197–30246. 10.1021/acsami.7b06001. [DOI] [PubMed] [Google Scholar]
- Lin H.; Zhou C.; Tian Y.; Siegrist T.; Ma B. Low-Dimensional Organometal Halide Perovskites. ACS Energy Lett. 2018, 3, 54–62. 10.1021/acsenergylett.7b00926. [DOI] [Google Scholar]
- Hong K.; Le Q. V.; Kim S. Y.; Jang H. W. Low-Dimensional Halide Perovskites: Review and Issues. J. Mater. Chem. C 2018, 6, 2189–2209. 10.1039/C7TC05658C. [DOI] [Google Scholar]
- Goldschmidt V. M. Die Gesetze der Krystallochemie. Naturwissenschaften 1926, 14, 477–485. 10.1007/BF01507527. [DOI] [Google Scholar]
- Kieslich G.; Sun S.; Cheetham A. K. Solid-State Principles Applied to Organic-Inorganic Perovskites: New Tricks for an Old Dog. Chem. Sci. 2014, 5, 4712–4715. 10.1039/C4SC02211D. [DOI] [Google Scholar]
- Travis W.; Glover E. N. K.; Bronstein H.; Scanlon D. O.; Palgrave R. G. On the Application of the Tolerance Factor to Inorganic and Hybrid Halide Perovskites: A Revised System. Chem. Sci. 2016, 7, 4548–4556. 10.1039/C5SC04845A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Breternitz J.; Lehmann F.; Barnett S. A.; Nowell H.; Schorr S. Role of the Iodide-Methylammonium Interaction in the Ferroelectricity of CH3NH3PbI3. Angew. Chem., Int. Ed. 2020, 59, 424–428. 10.1002/anie.201910599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Quan L. N.; Rand B. P.; Friend R. H.; Mhaisalkar S. G.; Lee T.-W.; Sargent E. H. Perovskites for Next-Generation Optical Sources. Chem. Rev. 2019, 119, 7444–7477. 10.1021/acs.chemrev.9b00107. [DOI] [PubMed] [Google Scholar]
- Han Y.; Yue S.; Cui B.-B. Low-Dimensional Metal Halide Perovskite Crystal Materials: Structure Strategies and Luminescence Applications. Adv. Sci. 2021, 8, 2004805 10.1002/advs.202004805. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maczka M.; Drozdowski D.; Stefanska D.; Gagor A.. Zero-Dimensional Mixed-Cation Hybrid Lead Halides with Broadband Emissions. Inorg. Chem. Front. 2023,.107222. 10.1039/D3QI01749D [DOI] [Google Scholar]
- Elleuch S.; Lusson A.; Pillet S.; Boukheddaden K.; Abid Y. White Light Emission from a Zero-Dimensional Lead Chloride Hybrid Material. ACS Photonics 2020, 7, 1178–1187. 10.1021/acsphotonics.9b01817. [DOI] [Google Scholar]
- Akkerman Q. A.; Manna L. What Defines a Halide Perovskite?. ACS Energy Lett. 2020, 5, 604–610. 10.1021/acsenergylett.0c00039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shao S.; Cui X.; Li Z. Recent Progress in Understanding the Structural, Optoelectronic, and Photophysical Properties of Lead Based Dion-Jacobson Perovskites as Well as Their Application in Solar Cells. ACS Mater. Lett. 2022, 4, 891–917. 10.1021/acsmaterialslett.1c00684. [DOI] [Google Scholar]
- Niu T.; Xue Q.; Yip H.-L. Advances in Dion-Jacobson Phase Two-Dimensional Metal Halide Perovskite Solar Cells. Nanophotonics 2021, 10, 2069–2102. 10.1515/nanoph-2021-0052. [DOI] [Google Scholar]
- Gao X.; Zhang X.; Yin W.; Wang H.; Hu Y.; Zhang Q.; Shi Z.; Colvin V. L.; Yu W. W.; Zhang Y. Ruddlesden-Popper Perovskites: Synthesis and Optical Properties for Optoelectronic Applications. Adv. Sci. 2019, 6, 1900941 10.1002/advs.201900941. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Siwach P.; Sikarwar P.; Halpati J. S.; Chandiran A. K. Design of Above-Room-Temperature Ferroelectric Two-Dimensional Layered Halide Perovskites. J. Mater. Chem. A 2022, 10, 8719–8738. 10.1039/D1TA09537D. [DOI] [Google Scholar]
- Cortecchia D.; Yin J.; Petrozza A.; Soci C. White Light Emission in Low-Dimensional Perovskites. J. Mater. Chem. C 2019, 7, 4956–4969. 10.1039/C9TC01036J. [DOI] [Google Scholar]
- Dohner E. R.; Jaffe A.; Bradshaw L. R.; Karunadasa H. I. Intrinsic White-Light Emission from Layered Hybrid Perovskites. J. Am. Chem. Soc. 2014, 136, 13154–13157. 10.1021/ja507086b. [DOI] [PubMed] [Google Scholar]
- Wang Y.; Song L.; Chen Y.; Huang W. Emerging New-Generation Photodetectors Based on Low-Dimensional Halide Perovskites. ACS Photonics 2020, 7, 10–28. 10.1021/acsphotonics.9b01233. [DOI] [Google Scholar]
- Chen X.-G.; Song X.-J.; Zhang Z.-X.; Li P.-F.; Ge J.-Z.; Tang Y.-Y.; Gao J.-X.; Zhang W.-Y.; Fu D.-W.; You Y.-M.; et al. Two-Dimensional Layered Perovskite Ferroelectric with Giant Piezoelectric Voltage Coefficient. J. Am. Chem. Soc. 2020, 142, 1077–1082. 10.1021/jacs.9b12368. [DOI] [PubMed] [Google Scholar]
- Xie A.; Maddalena F.; Witkowski M. E.; Makowski M.; Mahler B.; Drozdowski W.; Springham S. V.; Coquet P.; Dujardin C.; Birowosuto M. D.; et al. Library of Two-Dimensional Hybrid Lead Halide Perovskite Scintillator Crystals. Chem. Mater. 2020, 32, 8530–8539. 10.1021/acs.chemmater.0c02789. [DOI] [Google Scholar]
- Han K.; Wei Z.; Ye X.; Li B.; Wang P.; Cai H. A Lead Bromide Organic-Inorganic Hybrid Perovskite Material Showing Reversible Dual Phase Transition and Robust SHG Switching. Dalton Trans 2022, 51, 8273–8278. 10.1039/D2DT01040B. [DOI] [PubMed] [Google Scholar]
- Yang M.; Cheng H.; Xu Y.; Li M.; Ai Y. A Hybrid Organic-Inorganic Perovskite with Robust SHG Switching. Chin. Chem. Lett. 2022, 33, 2143–2146. 10.1016/j.cclet.2021.08.098. [DOI] [Google Scholar]
- Lun M.-M.; Zhou F.-L.; Fu D.-W.; Ye Q. Multi-Functional Hybrid Perovskites with Triple-Channel Switches and Optical Properties. J. Mater. Chem. C 2022, 10, 11371–11378. 10.1039/D2TC01063A. [DOI] [Google Scholar]
- Li H.-J.; Liu Y.-L.; Chen X.-G.; Gao J.-X.; Wang Z.-X.; Liao W.-Q. High-Temperature Dielectric Switching and Photoluminescence in a Corrugated Lead Bromide Layer Hybrid Perovskite Semiconductor. Inorg. Chem. 2019, 58, 10357–10363. 10.1021/acs.inorgchem.9b01538. [DOI] [PubMed] [Google Scholar]
- Ovčar J.; Leung T. L.; Grisanti L.; Skoko Z.; Vrankić M.; Low K.-H.; Wang S.; You P.-Y.; Ahn H.; Lončarić I.; et al. Mixed Halide Ordering as a Tool for the Stabilization of Ruddlesden-Popper Structures. Chem. Mater. 2022, 34, 4286–4297. 10.1021/acs.chemmater.1c03815. [DOI] [Google Scholar]
- Yangui A.; Pillet S.; Lusson A.; Bendeif E.-E.; Triki S.; Abid Y.; Boukheddaden K. Control of the White-Light Emission in the Mixed Two-Dimensional Hybrid Perovskites (C6H11NH3)2[PbBr4–xIx]. J. Alloys Compd. 2017, 699, 1122–1133. 10.1016/j.jallcom.2017.01.032. [DOI] [Google Scholar]
- Ma D.; Fu Y.; Dang L.; Zhai J.; Guzei I. A.; Jin S. Single-Crystal Microplates of Two-Dimensional Organic-Inorganic Lead Halide Layered Perovskites for Optoelectronics. Nano Res. 2017, 10, 2117–2129. 10.1007/s12274-016-1401-6. [DOI] [Google Scholar]
- Cai P.; Wang X.; Seo H. J.; Yan X. Bluish-White-Light-Emitting Diodes Based on Two-Dimensional Lead Halide Perovskite (C6H5C2H4NH3)2PbCl2Br2. Appl. Phys. Lett. 2018, 112, 153901 10.1063/1.5023797. [DOI] [Google Scholar]
- Pareja-Rivera C.; Morán-Muñoz J. A.; Gómora-Figueroa A. P.; Jancik V.; Vargas B.; Rodríguez-Hernández J.; Solis-Ibarra D. Optimizing Broadband Emission in 2D Halide Perovskites. Chem. Mater. 2022, 34, 9344–9349. 10.1021/acs.chemmater.2c00937. [DOI] [Google Scholar]
- Lin Y. L.; Johnson J. C. Interlayer Triplet Energy Transfer in Dion-Jacobson Two-Dimensional Lead Halide Perovskites Containing Naphthalene Diammonium Cations. J. Phys. Chem. Lett. 2021, 12, 4793–4798. 10.1021/acs.jpclett.1c01232. [DOI] [PubMed] [Google Scholar]
- Yan L.; Jana M. K.; Sercel P. C.; Mitzi D. B.; You W. Alkyl-Aryl Cation Mixing in Chiral 2D Perovskites. J. Am. Chem. Soc. 2021, 143, 18114–18120. 10.1021/jacs.1c06841. [DOI] [PubMed] [Google Scholar]
- Guo Y.-Y.; Yang L.-J.; Biberger S.; McNulty J. A.; Li T.; Schötz K.; Panzer F.; Lightfoot P. Structural Diversity in Layered Hybrid Perovskites, A2PbBr4 or AA$′$PbBr4, Templated by Small Disc-Shaped Amines. Inorg. Chem. 2020, 59, 12858–12866. 10.1021/acs.inorgchem.0c01807. [DOI] [PubMed] [Google Scholar]
- Guo Y.-Y.; McNulty J. A.; Mica N. A.; Samuel I. D. W.; Slawin A. M. Z.; Bühl M.; Lightfoot P. Structure-Directing Effects in (110)-Layered Hybrid Perovskites Containing Two Distinct Organic Moieties. Chem. Commun. 2019, 55, 9935–9938. 10.1039/C9CC04964A. [DOI] [PubMed] [Google Scholar]
- Stoumpos C. C.; Cao D. H.; Clark D. J.; Young J.; Rondinelli J. M.; Jang J. I.; Hupp J. T.; Kanatzidis M. G. Ruddlesden-Popper Hybrid Lead Iodide Perovskite 2D Homologous Semiconductors. Chem. Mater. 2016, 28, 2852–2867. 10.1021/acs.chemmater.6b00847. [DOI] [Google Scholar]
- Li X.; Hoffman J. M.; Kanatzidis M. G. The 2D Halide Perovskite Rulebook: How the Spacer Influences Everything from the Structure to Optoelectronic Device Efficiency. Chem. Rev. 2021, 121, 2230–2291. 10.1021/acs.chemrev.0c01006. [DOI] [PubMed] [Google Scholar]
- Li D.; Wu W.; Han S.; Liu X.; Peng Y.; Li X.; Li L.; Hong M.; Luo J. A Reduced-Dimensional Polar Hybrid Perovskite for Self-Powered Broad-Spectrum Photodetection. Chem. Sci. 2021, 12, 3050–3054. 10.1039/D0SC06112C. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li X.; Dong H.; Volonakis G.; Stoumpos C. C.; Even J.; Katan C.; Guo P.; Kanatzidis M. G. Ordered Mixed-Spacer 2D Bromide Perovskites and the Dual Role of 1,2,4-Triazolium Cation. Chem. Mater. 2022, 34, 6541–6552. 10.1021/acs.chemmater.2c01432. [DOI] [Google Scholar]
- Stoumpos C. C.; Mao L.; Malliakas C. D.; Kanatzidis M. G. Structure-Band Gap Relationships in Hexagonal Polytypes and Low-Dimensional Structures of Hybrid Tin Iodide Perovskites. Inorg. Chem. 2017, 56, 56–73. 10.1021/acs.inorgchem.6b02764. [DOI] [PubMed] [Google Scholar]
- Zhang Y.; Liu Y.; Liu S. F. Composition Engineering of Perovskite Single Crystals for High-Performance Optoelectronics. Adv. Funct. Mater. 2023, 33, 2210335 10.1002/adfm.202210335. [DOI] [Google Scholar]
- Kim J. Y.; Lee J.-W.; Jung H. S.; Shin H.; Park N.-G. High-Efficiency Perovskite Solar Cells. Chem. Rev. 2020, 120, 7867–7918. 10.1021/acs.chemrev.0c00107. [DOI] [PubMed] [Google Scholar]
- Jin R.-J.; Lou Y.-H.; Wang Z.-K. Doping Strategies for Promising Organic-Inorganic Halide Perovskites. Small 2023, 19, 2206581 10.1002/smll.202206581. [DOI] [PubMed] [Google Scholar]
- Frost J. M.; Butler K. T.; Brivio F.; Hendon C. H.; van Schilfgaarde M.; Walsh A. Atomistic Origins of High-Performance in Hybrid Halide Perovskite Solar Cells. Nano Lett. 2014, 14, 2584–2590. 10.1021/nl500390f. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Egger D. A.; Rappe A. M.; Kronik L. Hybrid Organic-Inorganic Perovskites on the Move. Acc. Chem. Res. 2016, 49, 573–581. 10.1021/acs.accounts.5b00540. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Herz L. M. How Lattice Dynamics Moderate the Electronic Properties of Metal-Halide Perovskites. J. Phys. Chem. Lett. 2018, 9, 6853–6863. 10.1021/acs.jpclett.8b02811. [DOI] [PubMed] [Google Scholar]
- Frost J. M.; Walsh A. What Is Moving in Hybrid Halide Perovskite Solar Cells?. Acc. Chem. Res. 2016, 49, 528–535. 10.1021/acs.accounts.5b00431. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu S.; Guo R.; Xie F. The Effects of Organic Cation Rotation in Hybrid Organic-Inorganic Perovskites: A Critical Review. Mater. Des. 2022, 221, 110951 10.1016/j.matdes.2022.110951. [DOI] [Google Scholar]
- Cowley R. A.; Gvasaliya S. N.; Lushnikov S. G.; Roessli B.; Rotaru G. M. Relaxing with Relaxors: A Review of Relaxor Ferroelectrics. Adv. Phys. 2011, 60, 229–327. 10.1080/00018732.2011.555385. [DOI] [Google Scholar]
- Vugmeister B. E.; Glinchuk M. D. Dipole Glass and Ferroelectricity in Random-Site Electric Dipole Systems. Rev. Mod. Phys. 1990, 62, 993–1026. 10.1103/RevModPhys.62.993. [DOI] [Google Scholar]
- Loidl A. Orientational Glasses. Annu. Rev. Phys. Chem. 1989, 40, 29–60. 10.1146/annurev.pc.40.100189.000333. [DOI] [Google Scholar]
- Samara G. A. The Relaxational Properties of Compositionally Disordered ABO3 Perovskites. J. Phys.: Condens. Matter 2003, 15, R367. 10.1088/0953-8984/15/9/202. [DOI] [Google Scholar]
- Bokov A. A.; Ye Z.-G. Recent Progress in Relaxor Ferroelectrics with Perovskite Structure. J. Mater. Sci. 2006, 41, 31–52. 10.1007/s10853-005-5915-7. [DOI] [Google Scholar]
- Ellis C. L. C.; Javaid H.; Smith E. C.; Venkataraman D. Hybrid Perovskites with Larger Organic Cations Reveal Autocatalytic Degradation Kinetics and Increased Stability under Light. Inorg. Chem. 2020, 59, 12176–12186. 10.1021/acs.inorgchem.0c01133. [DOI] [PubMed] [Google Scholar]
- Hsiao K.-C.; Jao M.-H.; Tian K.-Y.; Lin T.-H.; Tran D.-P.; Liao H.-C.; Hou C.-H.; Shyue J.-J.; Wu M.-C.; Su W.-F. Acetamidinium Cation to Confer Ion Immobilization and Structure Stabilization of Organometal Halide Perovskite Toward Long Life and High-Efficiency p-i-n Planar Solar Cell via Air-Processable Method. Solar RRL 2020, 4, 2000197 10.1002/solr.202070092. [DOI] [Google Scholar]
- Singh P.; Mukherjee R.; Avasthi S. Acetamidinium-Substituted Methylammonium Lead Iodide Perovskite Solar Cells with Higher Open-Circuit Voltage and Improved Intrinsic Stability. ACS Appl. Mater. Interfaces 2020, 12, 13982–13987. 10.1021/acsami.0c00663. [DOI] [PubMed] [Google Scholar]
- Kim J.; Hwang T.; Lee B.; Lee S.; Park K.; Park H. H.; Park B. An Aromatic Diamine Molecule as the A-Site Solute for Highly Durable and Efficient Perovskite Solar Cells. Small Methods 2019, 3, 1800361 10.1002/smtd.201800361. [DOI] [Google Scholar]
- Wang Q.; Lin F.; Chueh C.-C.; Zhao T.; Eslamian M.; Jen A. K.-Y. Enhancing Efficiency of Perovskite Solar Cells by Reducing Defects Through Imidazolium Cation Incorporation. Mater. Today Energy 2018, 7, 161–168. 10.1016/j.mtener.2017.09.007. [DOI] [Google Scholar]
- Shi Z.; Zhang Y.; Cui C.; Li B.; Zhou W.; Ning Z.; Mi Q. Symmetrization of the Crystal Lattice of MAPbI3 Boosts the Performance and Stability of Metal-Perovskite Photodiodes. Adv. Mater. 2017, 29, 1701656 10.1002/adma.201701656. [DOI] [PubMed] [Google Scholar]
- Jodlowski A. D.; Roldán-Carmona C.; Grancini G.; Salado M.; Ralaiarisoa M.; Ahmad S.; Koch N.; Camacho L.; de Miguel G.; Nazeeruddin M. K. Large Guanidinium Cation Mixed with Methylammonium in Lead Iodide Perovskites for 19% Efficient Solar Cells. Nat. Energy 2017, 2, 972–979. 10.1038/s41560-017-0054-3. [DOI] [Google Scholar]
- Akbulatov A. F.; Frolova L. A.; Anokhin D. V.; Gerasimov K. L.; Dremova N. N.; Troshin P. A. Hydrazinium-Loaded Perovskite Solar Cells with Enhanced Performance and Stability. J. Mater. Chem. A 2016, 4, 18378–18382. 10.1039/C6TA08215G. [DOI] [Google Scholar]
- Peng W.; Miao X.; Adinolfi V.; Alarousu E.; El Tall O.; Emwas A.-H.; Zhao C.; Walters G.; Liu J.; Ouellette O.; et al. Engineering of CH3NH3PbI3 Perovskite Crystals by Alloying Large Organic Cations for Enhanced Thermal Stability and Transport Properties. Angew. Chem., Int. Ed. 2016, 55, 10686–10690. 10.1002/anie.201604880. [DOI] [PubMed] [Google Scholar]
- Leblanc A.; Mercier N.; Allain M.; Dittmer J.; Pauporté T.; Fernandez V.; Boucher F.; Kepenekian M.; Katan C. Enhanced Stability and Band Gap Tuning of α-[HC(NH2)2]PbI3 Hybrid Perovskite by Large Cation Integration. ACS Appl. Mater. Interfaces 2019, 11, 20743–20751. 10.1021/acsami.9b00210. [DOI] [PubMed] [Google Scholar]
- Kumar G. S.; Sarkar P. K.; Pradhan B.; Hossain M.; Rao K. D. M.; Acharya S. Large-Area Transparent Flexible Guanidinium Incorporated MAPbI3 Microstructures for High-Performance Photodetectors with Enhanced Stability. Nanoscale Horiz 2020, 5, 696–704. 10.1039/C9NH00774A. [DOI] [PubMed] [Google Scholar]
- Ray A.; Martin-Garcia B.; Moliterni A.; Casati N.; Boopathi K. M.; Spirito D.; Goldoni L.; Prato M.; Giacobbe C.; Giannini C.; Di Stasio F.; Krahne R.; Manna L.; Abdelhady A. L. Mixed Dimethylammonium/Methylammonium Lead Halide Perovskite Crystals for Improved Structural Stability and Enhanced Photodetection. Adv. Mater. 2022, 34, 2106160 10.1002/adma.202106160. [DOI] [PubMed] [Google Scholar]
- Kubicki D. J.; Prochowicz D.; Hofstetter A.; Saski M.; Yadav P.; Bi D.; Pellet N.; Lewiński J.; Zakeeruddin S. M.; Grätzel M.; et al. Formation of Stable Mixed Guanidinium-Methylammonium Phases with Exceptionally Long Carrier Lifetimes for High-Efficiency Lead Iodide-Based Perovskite Photovoltaics. J. Am. Chem. Soc. 2018, 140, 3345–3351. 10.1021/jacs.7b12860. [DOI] [PubMed] [Google Scholar]
- Pellet N.; Gao P.; Gregori G.; Yang T.-Y.; Nazeeruddin M. K.; Maier J.; Grätzel M. Mixed-Organic-Cation Perovskite Photovoltaics for Enhanced Solar-Light Harvesting. Angew. Chem., Int. Ed. 2014, 53, 3151–3157. 10.1002/anie.201309361. [DOI] [PubMed] [Google Scholar]
- Wu S.; Li Z.; Zhang J.; Liu T.; Zhu Z.; Jen A. K.-Y. Efficient Large Guanidinium Mixed Perovskite Solar Cells with Enhanced Photovoltage and Low Energy Losses. Chem. Commun. 2019, 55, 4315–4318. 10.1039/C9CC00016J. [DOI] [PubMed] [Google Scholar]
- Shao F.; Qin P.; Wang D.; Zhang G.; Wu B.; He J.; Peng W.; Sum T. C.; Wang D.; Huang F. Enhanced Photovoltaic Performance and Thermal Stability of CH3NH3PbI3 Perovskite through Lattice Symmetrization. ACS Appl. Mater. Interfaces 2019, 11, 740–746. 10.1021/acsami.8b17068. [DOI] [PubMed] [Google Scholar]
- Kim G.; Min H.; Lee K. S.; Lee D. Y.; Yoon S. M.; Seok S. I. Impact of Strain Relaxation on Performance of α-Formamidinium Lead Iodide Perovskite Solar Cells. Science 2020, 370, 108–112. 10.1126/science.abc4417. [DOI] [PubMed] [Google Scholar]
- Gao L.; Li X.; Liu Y.; Fang J.; Huang S.; Spanopoulos I.; Li X.; Wang Y.; Chen L.; Yang G.; et al. Incorporated Guanidinium Expands the CH3NH3PbI3 Lattice and Enhances Photovoltaic Performance. ACS Appl. Mater. Interfaces 2020, 12, 43885–43891. 10.1021/acsami.0c14925. [DOI] [PubMed] [Google Scholar]
- Huang Y.; Qiao L.; Jiang Y.; He T.; Long R.; Yang F.; Wang L.; Lei X.; Yuan M.; Chen J. A-site Cation Engineering for Highly Efficient MAPbI3 Single-Crystal X-ray Detector. Angew. Chem., Int. Ed. 2019, 58, 17834–17842. 10.1002/anie.201911281. [DOI] [PubMed] [Google Scholar]
- Zhang X.; Xiao S.; Li R.; He T.; Xing G.; Chen R. Influence of Mixed Organic Cations on the Nonlinear Optical Properties of Lead Tri-Iodide Perovskites. Photon. Res. 2020, 8, A25–A30. 10.1364/PRJ.398975. [DOI] [Google Scholar]
- Choi H.; Jeong J.; Kim H.-B.; Kim S.; Walker B.; Kim G.-H.; Kim J. Y. Cesium-Doped Methylammonium Lead Iodide Perovskite Light Absorber for Hybrid Solar Cells. Nano Energy 2014, 7, 80–85. 10.1016/j.nanoen.2014.04.017. [DOI] [Google Scholar]
- Liu D.; Li Q.; Wu K. Ethylammonium as an Alternative Cation for Efficient Perovskite Solar Cells from First-Principles Calculations. RSC Adv. 2019, 9, 7356–7361. 10.1039/C9RA00853E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gholipour S.; Ali A. M.; Correa-Baena J.-P.; Turren-Cruz S.-H.; Tajabadi F.; Tress W.; Taghavinia N.; Gratzel M.; Abate A.; De Angelis F.; Gaggioli C. A.; Mosconi E.; Hagfeldt A.; Saliba M. Globularity-Selected Large Molecules for a New Generation of Multication Perovskites. Adv. Mater. 2017, 29, 1702005 10.1002/adma.201702005. [DOI] [PubMed] [Google Scholar]
- Anelli C.; Chierotti M. R.; Bordignon S.; Quadrelli P.; Marongiu D.; Bongiovanni G.; Malavasi L. Investigation of Dimethylammonium Solubility in MAPbBr3 Hybrid Perovskite: Synthesis, Crystal Structure, and Optical Properties. Inorg. Chem. 2019, 58, 944–949. 10.1021/acs.inorgchem.8b03072. [DOI] [PubMed] [Google Scholar]
- Huang J.; Xu P.; Liu J.; You X.-Z. Sequential Introduction of Cations Deriving Large-Grain CsxFA1–xPbI3 Thin Film for Planar Hybrid Solar Cells: Insight into Phase-Segregation and Thermal-Healing Behavior. Small 2017, 13, 1603225 10.1002/smll.201603225. [DOI] [PubMed] [Google Scholar]
- Francisco-López A.; Charles B.; Alonso M. I.; Garriga M.; Campoy-Quiles M.; Weller M. T.; Goñi A. R. Phase Diagram of Methylammonium/Formamidinium Lead Iodide Perovskite Solid Solutions from Temperature-Dependent Photoluminescence and Raman Spectroscopies. J. Phys. Chem. C 2020, 124, 3448–3458. 10.1021/acs.jpcc.9b10185. [DOI] [Google Scholar]
- Parrott E. S.; Green T.; Milot R. L.; Johnston M. B.; Snaith H. J.; Herz L. M. Interplay of Structural and Optoelectronic Properties in Formamidinium Mixed Tin-Lead Triiodide Perovskites. Adv. Funct. Mater. 2018, 28, 1802803 10.1002/adfm.201802803. [DOI] [Google Scholar]
- Wang C.; Song Z.; Li C.; Zhao D.; Yan Y. Low-Bandgap Mixed Tin-Lead Perovskites and Their Applications in All-Perovskite Tandem Solar Cells. Adv. Funct. Mater. 2019, 29, 1808801 10.1002/adfm.201808801. [DOI] [Google Scholar]
- Tong J.; Song Z.; Kim D. H.; Chen X.; Chen C.; Palmstrom A. F.; Ndione P. F.; Reese M. O.; Dunfield S. P.; Reid O. G.; et al. Carrier Lifetimes of > 1μs in Sn-Pb Perovskites Enable Efficient All-Perovskite Tandem Solar Cells. Science 2019, 364, 475–479. 10.1126/science.aav7911. [DOI] [PubMed] [Google Scholar]
- Savill K. J.; Ulatowski A. M.; Herz L. M. Optoelectronic Properties of Tin-Lead Halide Perovskites. ACS Energy Lett. 2021, 6, 2413–2426. 10.1021/acsenergylett.1c00776. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stoumpos C. C.; Malliakas C. D.; Kanatzidis M. G. Semiconducting Tin and Lead Iodide Perovskites with Organic Cations: Phase Transitions, High Mobilities, and Near-Infrared Photoluminescent Properties. Inorg. Chem. 2013, 52, 9019–9038. 10.1021/ic401215x. [DOI] [PubMed] [Google Scholar]
- Hao F.; Stoumpos C. C.; Chang R. P. H.; Kanatzidis M. G. Anomalous Band Gap Behavior in Mixed Sn and Pb Perovskites Enables Broadening of Absorption Spectrum in Solar Cells. J. Am. Chem. Soc. 2014, 136, 8094–8099. 10.1021/ja5033259. [DOI] [PubMed] [Google Scholar]
- Prasanna R.; Gold-Parker A.; Leijtens T.; Conings B.; Babayigit A.; Boyen H.-G.; Toney M. F.; McGehee M. D. Band Gap Tuning via Lattice Contraction and Octahedral Tilting in Perovskite Materials for Photovoltaics. J. Am. Chem. Soc. 2017, 139, 11117–11124. 10.1021/jacs.7b04981. [DOI] [PubMed] [Google Scholar]
- Senanayak S. P.; Dey K.; Shivanna R.; Li W.; Ghosh D.; Zhang Y.; Roose B.; Zelewski S. J.; Andaji-Garmaroudi Z.; Wood W.; et al. Charge Transport in Mixed Metal Halide Perovskite Semiconductors. Nat. Mater. 2023, 22, 216–224. 10.1038/s41563-022-01448-2. [DOI] [PubMed] [Google Scholar]
- Zhao D.; Yu Y.; Wang C.; Liao W.; Shrestha N.; Grice C. R.; Cimaroli A. J.; Guan L.; Ellingson R. J.; Zhu K.; Zhao X.; Xiong R.-G.; Yan Y. Low-Bandgap Mixed Tin-Lead Iodide Perovskite Absorbers with Long Carrier Lifetimes for All-Perovskite Tandem Solar Cells. Nat. Energy 2017, 2, 17018 10.1038/nenergy.2017.18. [DOI] [Google Scholar]
- Elsayed M. R. A.; Elseman A. M.; Abdelmageed A. A.; Hashem H. M.; Hassen A. Green and Cost-Effective Morter Grinding Synthesis of Bismuth-Doped Halide Perovskites as Efficient Absorber Materials. J. Mater. Sci.: Mater. Electron 2023, 34, 194. 10.1007/s10854-022-09574-y. [DOI] [Google Scholar]
- Wieghold S.; Luo Y.; Bieber A. S.; Lackner J.; Shirato N.; VanOrman Z. A.; Rosenmann D.; Nienhaus K.; Lai B.; Nienhaus G. U.; et al. Impact of Transition Metal Doping on the Structural and Optical Properties of Halide Perovskites. Chem. Mater. 2021, 33, 6099–6107. 10.1021/acs.chemmater.1c01645. [DOI] [Google Scholar]
- Zhou Y.; Chen J.; Bakr O. M.; Sun H.-T. Metal-Doped Lead Halide Perovskites: Synthesis, Properties, and Optoelectronic Applications. Chem. Mater. 2018, 30, 6589–6613. 10.1021/acs.chemmater.8b02989. [DOI] [Google Scholar]
- Swarnkar A.; Mir W. J.; Nag A. Can B-Site Doping or Alloying Improve Thermal- and Phase-Stability of All-Inorganic CsPbX3 (X = Cl, Br, I) Perovskites?. ACS Energy Lett. 2018, 3, 286–289. 10.1021/acsenergylett.7b01197. [DOI] [Google Scholar]
- Kubicki D. J.; Prochowicz D.; Hofstetter A.; Péchy P.; Zakeeruddin S. M.; Grätzel M.; Emsley L. Cation Dynamics in Mixed-Cation (MA)x(FA)1–xPbI3 Hybrid Perovskites from Solid-State NMR. J. Am. Chem. Soc. 2017, 139, 10055–10061. 10.1021/jacs.7b04930. [DOI] [PubMed] [Google Scholar]
- Liu F.; Zhang Y.; Ding C.; Kawabata K.; Yoshihara Y.; Toyoda T.; Hayase S.; Minemoto T.; Wang R.; Shen Q. Trioctylphosphine Oxide Acts as Alkahest for SnX2/PbX2: A General Synthetic Route to Perovskite ASnxPb1–xX3 (A = Cs, FA, MA; X = Cl, Br, I) Quantum Dots. Chem. Mater. 2020, 32, 1089–1100. 10.1021/acs.chemmater.9b03918. [DOI] [Google Scholar]
- Eperon G. E.; Leijtens T.; Bush K. A.; Prasanna R.; Green T.; Wang J. T.-W.; McMeekin D. P.; Volonakis G.; Milot R. L.; May R.; et al. Perovskite-Perovskite Tandem Photovoltaics with Optimized Band Gaps. Science 2016, 354, 861–865. 10.1126/science.aaf9717. [DOI] [PubMed] [Google Scholar]
- Bush K. A.; Palmstrom A. F.; Yu Z. J.; Boccard M.; Cheacharoen R.; Mailoa J. P.; McMeekin D. P.; Hoye R. L. Z.; Bailie C. D.; Leijtens T.; Peters I. M.; Minichetti M. C.; Rolston N.; Prasanna R.; Sofia S.; Harwood D.; Ma W.; Moghadam F.; Snaith H. J.; Buonassisi T.; Holman Z. C.; Bent S. F.; McGehee M. D. 23.6%-Efficient Monolithic Perovskite/Silicon Tandem Solar Cells with Improved Stability. Nat. Energy 2017, 2, 17009 10.1038/nenergy.2017.9. [DOI] [Google Scholar]
- Koh T. M.; Fu K.; Fang Y.; Chen S.; Sum T. C.; Mathews N.; Mhaisalkar S. G.; Boix P. P.; Baikie T. Formamidinium-Containing Metal-Halide: An Alternative Material for Near-IR Absorption Perovskite Solar Cells. J. Phys. Chem. C 2014, 118, 16458–16462. 10.1021/jp411112k. [DOI] [Google Scholar]
- Eperon G. E.; Paternò G. M.; Sutton R. J.; Zampetti A.; Haghighirad A. A.; Cacialli F.; Snaith H. J. Inorganic Caesium Lead Iodide Perovskite Solar Cells. J. Mater. Chem. A 2015, 3, 19688–19695. 10.1039/C5TA06398A. [DOI] [Google Scholar]
- Binek A.; Hanusch F. C.; Docampo P.; Bein T. Stabilization of the Trigonal High-Temperature Phase of Formamidinium Lead Iodide. J. Phys. Chem. Lett. 2015, 6, 1249–1253. 10.1021/acs.jpclett.5b00380. [DOI] [PubMed] [Google Scholar]
- Merten L.; Hinderhofer A.; Baumeler T.; Arora N.; Hagenlocher J.; Zakeeruddin S. M.; Dar M. I.; Grätzel M.; Schreiber F. Quantifying Stabilized Phase Purity in Formamidinium-Based Multiple-Cation Hybrid Perovskites. Chem. Mater. 2021, 33, 2769–2776. 10.1021/acs.chemmater.0c04185. [DOI] [Google Scholar]
- Li Z.; Yang M.; Park J.-S.; Wei S.-H.; Berry J. J.; Zhu K. Stabilizing Perovskite Structures by Tuning Tolerance Factor: Formation of Formamidinium and Cesium Lead Iodide Solid-State Alloys. Chem. Mater. 2016, 28, 284–292. 10.1021/acs.chemmater.5b04107. [DOI] [Google Scholar]
- Jeon N. J.; Noh J. H.; Yang W. S.; Kim Y. C.; Ryu S.; Seo J.; Seok S. I. Compositional Engineering of Perovskite Materials for High-Performance Solar Cells. Nature 2015, 517, 476–480. 10.1038/nature14133. [DOI] [PubMed] [Google Scholar]
- Baumeler T.; Arora N.; Hinderhofer A.; Akin S.; Greco A.; Abdi-Jalebi M.; Shivanna R.; Uchida R.; Liu Y.; Schreiber F.; et al. Minimizing the Trade-Off between Photocurrent and Photovoltage in Triple-Cation Mixed-Halide Perovskite Solar Cells. J. Phys. Chem. Lett. 2020, 11, 10188–10195. 10.1021/acs.jpclett.0c02791. [DOI] [PubMed] [Google Scholar]
- McMeekin D. P.; Sadoughi G.; Rehman W.; Eperon G. E.; Saliba M.; Hörantner M. T.; Haghighirad A.; Sakai N.; Korte L.; Rech B.; et al. A Mixed-Cation Lead Mixed-Halide Perovskite Absorber for Tandem Solar Cells. Science 2016, 351, 151–155. 10.1126/science.aad5845. [DOI] [PubMed] [Google Scholar]
- Min H.; Kim M.; Lee S.-U.; Kim H.; Kim G.; Choi K.; Lee J. H.; Seok S. I. Efficient, Stable Solar Cells by using Inherent Bandgap of α-Phase Formamidinium Lead Iodide. Science 2019, 366, 749–753. 10.1126/science.aay7044. [DOI] [PubMed] [Google Scholar]
- Schmidt-Mende L.; Dyakonov V.; Olthof S.; Unlu F.; Le K. M. T.; Mathur S.; Karabanov A. D.; Lupascu D. C.; Herz L. M.; Hinderhofer A.; Schreiber F.; Chernikov A.; Egger D. A.; Shargaieva O.; Cocchi C.; Unger E.; Saliba M.; Byranvand M. M.; Kroll M.; Nehm F.; Leo K.; Redinger A.; Hocker J.; Kirchartz T.; Warby J.; Gutierrez-Partida E.; Neher D.; Stolterfoht M.; Wurfel U.; Unmussig M.; Herterich J.; Baretzky C.; Mohanraj J.; Thelakkat M.; Maheu C.; Jaegermann W.; Mayer T.; Rieger J.; Fauster T.; Niesner D.; Yang F.; Albrecht S.; Riedl T.; Fakharuddin A.; Vasilopoulou M.; Vaynzof Y.; Moia D.; Maier J.; Franckevicius M.; Gulbinas V.; Kerner R. A.; Zhao L.; Rand B. P.; Gluck N.; Bein T.; Matteocci F.; Castriotta L. A.; Di Carlo A.; Scheffler M.; Draxl C. Roadmap on Organic-Inorganic Hybrid Perovskite Semiconductors and Devices. APL Mater. 2021, 9, 109202 10.1063/5.0047616. [DOI] [Google Scholar]
- Eperon G. E.; Stranks S. D.; Menelaou C.; Johnston M. B.; Herz L. M.; Snaith H. J. Formamidinium Lead Trihalide: a Broadly Tunable Perovskite for Efficient Planar Heterojunction Solar Cells. Energy Environ. Sci. 2014, 7, 982–988. 10.1039/c3ee43822h. [DOI] [Google Scholar]
- Rybin N.; Ghosh D.; Tisdale J.; Shrestha S.; Yoho M.; Vo D.; Even J.; Katan C.; Nie W.; Neukirch A. J.; et al. Effects of Chlorine Mixing on Optoelectronics, Ion Migration, and Gamma-Ray Detection in Bromide Perovskites. Chem. Mater. 2020, 32, 1854–1863. 10.1021/acs.chemmater.9b04244. [DOI] [Google Scholar]
- Drozdowski D.; Gagor A.; Stefanska D.; Zareba J. K.; Fedoruk K.; Maczka M.; Sieradzki A. Three-Dimensional Methylhydrazinium Lead Halide Perovskites: Structural Changes and Effects on Dielectric, Linear, and Nonlinear Optical Properties Entailed by the Halide Tuning. J. Phys. Chem. C 2022, 126, 1600–1610. 10.1021/acs.jpcc.1c07911. [DOI] [Google Scholar]
- Chen Y.; Motti S. G.; Oliver R. D. J.; Wright A. D.; Snaith H. J.; Johnston M. B.; Herz L. M.; Filip M. R. Optoelectronic Properties of Mixed Iodide-Bromide Perovskites from First-Principles Computational Modeling and Experiment. J. Phys. Chem. Lett. 2022, 13, 4184–4192. 10.1021/acs.jpclett.2c00938. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang H.; Ling F.; Luo C.; Li D.; Xiao Y.; Chang Z.; Xu Z.; Zeng Y.; Wang W.; Yao J. Active Terahertz Modulator Based on Optically Controlled Organometal Halide Perovskite MAPbI2Br. Appl. Opt. 2022, 61, 1171–1176. 10.1364/AO.444667. [DOI] [PubMed] [Google Scholar]
- Suárez I.; Vallés-Pelarda M.; Gualdrón-Reyes A. F.; Mora-Seró I.; Ferrando A.; Michinel H.; Salgueiro J. R.; Pastor J. P. M. Outstanding Nonlinear Optical Properties of Methylammonium- and Cs– PbX3 (X = Br, I, and Br-I) Perovskites: Polycrystalline Thin Films and Nanoparticles. APL Mater. 2019, 7, 041106 10.1063/1.5090926. [DOI] [Google Scholar]
- Wei Q.; Liang H.; Haruta Y.; Saidaminov M.; Mi Q.; Saliba M.; Cui G.; Ning Z. From Tetragonal to Cubic: Perovskite Phase Structure Evolution for High-Performance Solar Cells. Sci. Bull. 2023, 68, 141–145. 10.1016/j.scib.2023.01.008. [DOI] [PubMed] [Google Scholar]
- He J.; Li T.; Liu X.; Su H.; Ku Z.; Zhong J.; Huang F.; Peng Y.; Cheng Y.-B. Influence of Phase Transition on Stability of Perovskite Solar Cells under Thermal Cycling Conditions. Sol. Energy 2019, 188, 312–317. 10.1016/j.solener.2019.06.025. [DOI] [Google Scholar]
- Shao D.-S.; Sang L.; Kong Y.-R.; Deng Z.-R.; Luo H.-B.; Tian Z.-F.; Ren X.-M. Tunable Thermotropic Phase Transition Triggering Large Dielectric Response and Superionic Conduction in Lead Halide Perovskites. Inorg. Chem. Front. 2022, 9, 5653–5662. 10.1039/D2QI01650H. [DOI] [Google Scholar]
- Yu J.; Kong J.; Hao W.; Guo X.; He H.; Leow W. R.; Liu Z.; Cai P.; Qian G.; Li S.; Chen X.; Chen X. Broadband Extrinsic Self-Trapped Exciton Emission in Sn-Doped 2D Lead-Halide Perovskites. Adv. Mater. 2019, 31, 1806385 10.1002/adma.201806385. [DOI] [PubMed] [Google Scholar]
- Li T.; Chen X.; Wang X.; Lu H.; Yan Y.; Beard M. C.; Mitzi D. B. Origin of Broad-Band Emission and Impact of Structural Dimensionality in Tin-Alloyed Ruddlesden-Popper Hybrid Lead Iodide Perovskites. ACS Energy Lett. 2020, 5, 347–352. 10.1021/acsenergylett.9b02490. [DOI] [Google Scholar]
- Sui S.; Zhou J.; Wang A.; Hu G.; Meng W.; Wang C.; Liu Y.; Wu J.; Deng Z. Synthesis of Two-Dimensional Phenylethylamine Tin-Lead Halide Perovskites with Bandgap Bending Behavior. Nanoscale Adv. 2021, 3, 3875–3880. 10.1039/D0NA00939C. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Biswas A.; Bakthavatsalam R.; Kundu J. Efficient Exciton to Dopant Energy Transfer in Mn2+-Doped (C4H9NH3)2PbBr4 Two-Dimensional (2D) Layered Perovskites. Chem. Mater. 2017, 29, 7816–7825. 10.1021/acs.chemmater.7b02429. [DOI] [Google Scholar]
- Liao J.-F.; Zhang Z.; Wang B.; Tang Z.; Xing G. Full-Color-Tunable Phosphorescence of Antimony-Doped Lead Halide Single Crystal. npj Flexible Electron 2022, 6, 57. 10.1038/s41528-022-00194-4. [DOI] [Google Scholar]
- Wu Z.; Zhang W.; Ye H.; Yao Y.; Liu X.; Li L.; Ji C.; Luo J. Bromine-Substitution-Induced High-Tc Two-Dimensional Bilayered Perovskite Photoferroelectric. J. Am. Chem. Soc. 2021, 143, 7593–7598. 10.1021/jacs.1c00459. [DOI] [PubMed] [Google Scholar]
- Park I.-H.; Kwon K. C.; Zhu Z.; Wu X.; Li R.; Xu Q.-H.; Loh K. P. Self-Powered Photodetector Using Two-Dimensional Ferroelectric Dion-Jacobson Hybrid Perovskites. J. Am. Chem. Soc. 2020, 142, 18592–18598. 10.1021/jacs.0c08189. [DOI] [PubMed] [Google Scholar]
- Zhang Q.; Solanki A.; Parida K.; Giovanni D.; Li M.; Jansen T. L. C.; Pshenichnikov M. S.; Sum T. C. Tunable Ferroelectricity in Ruddlesden-Popper Halide Perovskites. ACS Appl. Mater. Interfaces 2019, 11, 13523–13532. 10.1021/acsami.8b21579. [DOI] [PubMed] [Google Scholar]
- Han S.; Liu X.; Liu Y.; Xu Z.; Li Y.; Hong M.; Luo J.; Sun Z. High-Temperature Antiferroelectric of Lead Iodide Hybrid Perovskites. J. Am. Chem. Soc. 2019, 141, 12470–12474. 10.1021/jacs.9b05124. [DOI] [PubMed] [Google Scholar]
- Tang L.; Han S.; Ma Y.; Liu Y.; Hua L.; Xu H.; Guo W.; Wang B.; Sun Z.; Luo J. Giant Near-Room-Temperature Pyroelectric Figures-of-Merit Originating from Unusual Dielectric Bistability of Two-Dimensional Perovskite Ferroelectric Crystals. Chem. Mater. 2022, 34, 8898–8904. 10.1021/acs.chemmater.2c02211. [DOI] [Google Scholar]
- Ko B. A.; Berry K.; Qin Z.; Sokolov A. V.; Hu J.; Scully M. O.; Bao J.; Zhang Z. Resonant Degenerate Four-Wave Mixing at the Defect Energy Levels of 2D Organic-Inorganic Hybrid Perovskite Crystals. ACS Appl. Mater. Interfaces 2021, 13, 57075–57083. 10.1021/acsami.1c14092. [DOI] [PubMed] [Google Scholar]
- Leung T. L.; Tam H. W.; Liu F.; Lin J.; Ng A. M. C.; Chan W. K.; Chen W.; He Z.; Loncaric I.; Grisanti L.; Ma C.; Wong K. S.; Lau Y. S.; Zhu F.; Skoko Z.; Popovic J.; Djurisic A. B. Mixed Spacer Cation Stabilization of Blue-Emitting n = 2 Ruddlesden-Popper Organic-Inorganic Halide Perovskite Films. Adv. Opt. Mater. 2020, 8, 1901679 10.1002/adom.201901679. [DOI] [Google Scholar]
- Shi J.; Jin X.; Wu Y.; Shao M. Mixed Bulky Cations for Efficient and Stable Ruddlesden-Popper Perovskite Solar Cells. APL Mater. 2020, 8, 101102 10.1063/5.0024135. [DOI] [Google Scholar]
- Mao L.; Guo P.; Kepenekian M.; Spanopoulos I.; He Y.; Katan C.; Even J.; Schaller R. D.; Seshadri R.; Stoumpos C. C.; et al. Organic Cation Alloying on Intralayer A and Interlayer A’ sites in 2D Hybrid Dion-Jacobson Lead Bromide Perovskites (A’)(A)Pb2Br7. J. Am. Chem. Soc. 2020, 142, 8342–8351. 10.1021/jacs.0c01625. [DOI] [PubMed] [Google Scholar]
- Qin X.; Liu F.; Leung T. L.; Sun W.; Chan C. C. S.; Wong K. S.; Kanižaj L.; Popović J.; Djurišić A. B. Compositional Optimization of Mixed Cation Dion-Jacobson Perovskites for Efficient Green Light Emission. J. Mater. Chem. C 2021, 10, 108–114. 10.1039/D1TC04743D. [DOI] [Google Scholar]
- Fu Y.; Jiang X.; Li X.; Traore B.; Spanopoulos I.; Katan C.; Even J.; Kanatzidis M. G.; Harel E. Cation Engineering in Two-Dimensional Ruddlesden-Popper Lead Iodide Perovskites with Mixed Large A-Site Cations in the Cages. J. Am. Chem. Soc. 2020, 142, 4008–4021. 10.1021/jacs.9b13587. [DOI] [PubMed] [Google Scholar]
- Ramos-Terrón S.; Jodlowski A. D.; Verdugo-Escamilla C.; Camacho L.; de Miguel G. Relaxing the Goldschmidt Tolerance Factor: Sizable Incorporation of the Guanidinium Cation into a Two-Dimensional Ruddlesden-Popper Perovskite. Chem. Mater. 2020, 32, 4024–4037. 10.1021/acs.chemmater.0c00613. [DOI] [Google Scholar]
- Ke W.; Chen C.; Spanopoulos I.; Mao L.; Hadar I.; Li X.; Hoffman J. M.; Song Z.; Yan Y.; Kanatzidis M. G. Narrow-Bandgap Mixed Lead/Tin-Based 2D Dion-Jacobson Perovskites Boost the Performance of Solar Cells. J. Am. Chem. Soc. 2020, 142, 15049–15057. 10.1021/jacs.0c06288. [DOI] [PubMed] [Google Scholar]
- Ramirez D.; Schutt K.; Wang Z.; Pearson A. J.; Ruggeri E.; Snaith H. J.; Stranks S. D.; Jaramillo F. Layered Mixed Tin-Lead Hybrid Perovskite Solar Cells with High Stability. ACS Energy Lett. 2018, 3, 2246–2251. 10.1021/acsenergylett.8b01411. [DOI] [Google Scholar]
- Vashishtha P.; Ng M.; Shivarudraiah S. B.; Halpert J. E. High Efficiency Blue and Green Light-Emitting Diodes Using Ruddlesden-Popper Inorganic Mixed Halide Perovskites with Butylammonium Interlayers. Chem. Mater. 2019, 31, 83–89. 10.1021/acs.chemmater.8b02999. [DOI] [Google Scholar]
- Toso S.; Gushchina I.; Oliver A. G.; Manna L.; Kuno M. Are Mixed-Halide Ruddlesden-Popper Perovskites Really Mixed?. ACS Energy Lett. 2022, 7, 4242–4247. 10.1021/acsenergylett.2c01967. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang W.; Xiong R.-G. Ferroelectric Metal-Organic Frameworks. Chem. Rev. 2012, 112, 1163–1195. 10.1021/cr200174w. [DOI] [PubMed] [Google Scholar]
- Kleemann W. Random Fields in Dipolar Glasses and Relaxors. J. Non-Cryst. Solids 2002, 307–310, 66–72. 10.1016/S0022-3093(02)01441-2. [DOI] [Google Scholar]
- Höchli U.; Knorr K.; Loidl A. Orientational Glasses. Adv. Phys. 1990, 39, 405–615. 10.1080/00018739000101521. [DOI] [Google Scholar]
- Sarkar S.; Ren X.; Otsuka K. Evidence for Strain Glass in the Ferroelastic-Martensitic System Ti50–xNi50+x. Phys. Rev. Lett. 2005, 95, 205702 10.1103/PhysRevLett.95.205702. [DOI] [PubMed] [Google Scholar]
- Kleemann W. Random Fields in Relaxor Ferroelectrics - a Jubilee Review. J. Adv. Dielectr. 2012, 02, 1241001 10.1142/S2010135X12410019. [DOI] [Google Scholar]
- Grigalaitis R.; Banys J.; Kania A.; Slodczyk A. Distribution of Relaxation Times in PMN Single Crystal. J. Phys. IV France 2005, 128, 127–131. 10.1051/jp4:2005128020. [DOI] [Google Scholar]
- Jeong I.-K.; Darling T. W.; Lee J. K.; Proffen T.; Heffner R. H.; Park J. S.; Hong K. S.; Dmowski W.; Egami T. Direct Observation of the Formation of Polar Nanoregions in Pb(Mg1/3Nb2/3)O3 Using Neutron Pair Distribution Function Analysis. Phys. Rev. Lett. 2005, 94, 147602 10.1103/PhysRevLett.94.147602. [DOI] [PubMed] [Google Scholar]
- Banys J.; Klimm C.; Völkel G.; Böttcher R.; Bauch H.; Klöpperpieper A. Proton Glass Behaviour in a Solid Solution of Irradiated Betaine Phosphate0.15 Betaine Phosphite0.85. J. Phys.: Condens. Matter 1996, 8, L245. 10.1088/0953-8984/8/16/001. [DOI] [PubMed] [Google Scholar]
- Samara G. A. Pressure and Temperature Dependence of the Dielectric Properties and Phase Transitions of the Ferroelectric Perovskites: PbTiO3 and BaTiO3. Ferroelectrics 1971, 2, 277–289. 10.1080/00150197108234102. [DOI] [Google Scholar]
- Levstik A.; Kutnjak Z.; Filipič C.; Pirc R. Glassy Freezing in Relaxor Ferroelectric Lead Magnesium Niobate. Phys. Rev. B 1998, 57, 11204–11211. 10.1103/PhysRevB.57.11204. [DOI] [Google Scholar]
- Courtens E. Vogel-Fulcher Scaling of the Susceptibility in a Mixed-Crystal Proton Glass. Phys. Rev. Lett. 1984, 52, 69–72. 10.1103/PhysRevLett.52.69. [DOI] [Google Scholar]
- Tachibana M.; Takayama-Muromachi E. Thermal Conductivity and Heat Capacity of the Relaxor Ferroelectric [PbMg1/3Nb2/3O3]1-x[PbTiO3]x. Phys. Rev. B 2009, 79, 100104 10.1103/PhysRevB.79.100104. [DOI] [Google Scholar]
- Yu S.; Li L.; Zhang W.; Sun Z.; Dong H. Multilayer Thin Films with Compositional PbZr0.52Ti0.48O3/Bi1.5Zn1.0Nb1.5O7 Layers for Tunable Applications. Sci. Rep. 2015, 5, 10173 10.1038/srep10173. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Park S.-E.; Shrout T. R. Ultrahigh Strain and Piezoelectric Behavior in Relaxor Based Ferroelectric Single Crystals. J. Appl. Phys. 1997, 82, 1804–1811. 10.1063/1.365983. [DOI] [Google Scholar]
- Zhang L.; Zhao C.; Zheng T.; Wu J. Large Electrocaloric Effect in (Bi0.5Na0.5)TiO3-Based Relaxor Ferroelectrics. ACS Appl. Mater. Interfaces 2020, 12, 33934–33940. 10.1021/acsami.0c09343. [DOI] [PubMed] [Google Scholar]
- Ptak M.; Sieradzki A.; Šimėnas M.; Maczka M. Molecular Spectroscopy of Hybrid Organic-Inorganic Perovskites and Related Compounds. Coord. Chem. Rev. 2021, 448, 214180 10.1016/j.ccr.2021.214180. [DOI] [Google Scholar]
- Weber O. J.; Charles B.; Weller M. T. Phase Behaviour and Composition in the Formamidinium-Methylammonium Hybrid Lead Iodide Perovskite Solid Solution. J. Mater. Chem. A 2016, 4, 15375–15382. 10.1039/C6TA06607K. [DOI] [Google Scholar]
- Mozur E. M.; Maughan A. E.; Cheng Y.; Huq A.; Jalarvo N.; Daemen L. L.; Neilson J. R. Orientational Glass Formation in Substituted Hybrid Perovskites. Chem. Mater. 2017, 29, 10168–10177. 10.1021/acs.chemmater.7b04017. [DOI] [Google Scholar]
- Simenas M.; Balciunas S.; Wilson J. N.; Svirskas S.; Kinka M.; Garbaras A.; Kalendra V.; Gagor A.; Szewczyk D.; Sieradzki A.; Maczka M.; Samulionis V.; Walsh A.; Grigalaitis R.; Banys J. Suppression of Phase Transitions and Glass Phase Signatures in Mixed Cation Halide Perovskites. Nat. Commun. 2020, 11, 5103. 10.1038/s41467-020-18938-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Simenas M.; Balciunas S.; Svirskas S.; Kinka M.; Ptak M.; Kalendra V.; Gagor A.; Szewczyk D.; Sieradzki A.; Grigalaitis R.; Walsh A.; Maczka M.; Banys J. Phase Diagram and Cation Dynamics of Mixed MA1–xFAxPbBr3 Hybrid Perovskites. Chem. Mater. 2021, 33, 5926–5934. 10.1021/acs.chemmater.1c00885. [DOI] [Google Scholar]
- Simenas M.; Balciunas S.; Gagor A.; Pieniazek A.; Tolborg K.; Kinka M.; Klimavicius V.; Svirskas S.; Kalendra V.; Ptak M.; Szewczyk D.; Herman A. P.; Kudrawiec R.; Sieradzki A.; Grigalaitis R.; Walsh A.; Maczka M.; Banys J. Mixology of MA1–xEAxPbI3 Hybrid Perovskites: Phase Transitions, Cation Dynamics, and Photoluminescence. Chem. Mater. 2022, 34, 10104–10112. 10.1021/acs.chemmater.2c02807. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Weller M. T.; Weber O. J.; Henry P. F.; Di Pumpo A. M.; Hansen T. C. Complete Structure and Cation Orientation in the Perovskite Photovoltaic Methylammonium Lead Iodide Between 100 and 352 K. Chem. Commun. 2015, 51, 4180–4183. 10.1039/C4CC09944C. [DOI] [PubMed] [Google Scholar]
- López C. A.; Martínez-Huerta M. V.; Alvarez-Galván M. C.; Kayser P.; Gant P.; Castellanos-Gomez A.; Fernández-Díaz M. T.; Fauth F.; Alonso J. A. Elucidating the Methylammonium (MA) Conformation in MAPbBr3 Perovskite with Application in Solar Cells. Inorg. Chem. 2017, 56, 14214–14219. 10.1021/acs.inorgchem.7b02344. [DOI] [PubMed] [Google Scholar]
- Yang B.; Ming W.; Du M.-H.; Keum J. K.; Puretzky A. A.; Rouleau C. M.; Huang J.; Geohegan D. B.; Wang X.; Xiao K. Real-Time Observation of Order-Disorder Transformation of Organic Cations Induced Phase Transition and Anomalous Photoluminescence in Hybrid Perovskites. Adv. Mater. 2018, 30, 1705801 10.1002/adma.201705801. [DOI] [PubMed] [Google Scholar]
- Schueller E. C.; Laurita G.; Fabini D. H.; Stoumpos C. C.; Kanatzidis M. G.; Seshadri R. Crystal Structure Evolution and Notable Thermal Expansion in Hybrid Perovskites Formamidinium Tin Iodide and Formamidinium Lead Bromide. Inorg. Chem. 2018, 57, 695–701. 10.1021/acs.inorgchem.7b02576. [DOI] [PubMed] [Google Scholar]
- Charles B.; Weller M. T.; Rieger S.; Hatcher L. E.; Henry P. F.; Feldmann J.; Wolverson D.; Wilson C. C. Phase Behavior and Substitution Limit of Mixed Cesium-Formamidinium Lead Triiodide Perovskites. Chem. Mater. 2020, 32, 2282–2291. 10.1021/acs.chemmater.9b04032. [DOI] [Google Scholar]
- Mozur E. M.; Hope M. A.; Trowbridge J. C.; Halat D. M.; Daemen L. L.; Maughan A. E.; Prisk T. R.; Grey C. P.; Neilson J. R. Cesium Substitution Disrupts Concerted Cation Dynamics in Formamidinium Hybrid Perovskites. Chem. Mater. 2020, 32, 6266–6277. 10.1021/acs.chemmater.0c01862. [DOI] [Google Scholar]
- Krogstad M. J.; Gehring P. M.; Rosenkranz S.; Osborn R.; Ye F.; Liu Y.; Ruff J. P. C.; Chen W.; Wozniak J. M.; Luo H.; et al. The Relation of Local Order to Material Properties in Relaxor Ferroelectrics. Nat. Mater. 2018, 17, 718–724. 10.1038/s41563-018-0112-7. [DOI] [PubMed] [Google Scholar]
- Paściak M.; Wołcyrz M.; Pietraszko A. Interpretation of the Diffuse Scattering in Pb-Based Relaxor Ferroelectrics in Terms of Three-Dimensional Nanodomains of the < 110>-Directed Relative Interdomain Atomic Shifts. Phys. Rev. B 2007, 76, 014117 10.1103/PhysRevB.76.014117. [DOI] [Google Scholar]
- Welberry T.; Weber T. One Hundred Years of Diffuse Scattering. Crystallogr. Rev. 2016, 22, 2–78. 10.1080/0889311X.2015.1046853. [DOI] [Google Scholar]
- Proffen T.; Kim H. Advances in Total Scattering Analysis. J. Mater. Chem. 2009, 19, 5078–5088. 10.1039/b821178g. [DOI] [Google Scholar]
- Proffen T.; Billinge S. J. L.; Egami T.; Louca D. Structural Analysis of Complex Materials Using the Atomic Pair Distribution Function - a Practical Guide. Z. Kristallogr. Cryst. Mater. 2003, 218, 132–143. 10.1524/zkri.218.2.132.20664. [DOI] [Google Scholar]
- Weadock N. J.; Sterling T. C.; Vigil J. A.; Gold-Parker A.; Smith I. C.; Ahammed B.; Krogstad M. J.; Ye F.; Voneshen D.; Gehring P. M.; et al. The Nature of Dynamic Local Order in CH3NH3PbI3 and CH3NH3PbBr3. Joule 2023, 7, 1051–1066. 10.1016/j.joule.2023.03.017. [DOI] [Google Scholar]
- Beecher A. N.; Semonin O. E.; Skelton J. M.; Frost J. M.; Terban M. W.; Zhai H.; Alatas A.; Owen J. S.; Walsh A.; Billinge S. J. L. Direct Observation of Dynamic Symmetry Breaking above Room Temperature in Methylammonium Lead Iodide Perovskite. ACS Energy Lett. 2016, 1, 880–887. 10.1021/acsenergylett.6b00381. [DOI] [Google Scholar]
- Bertolotti F.; Protesescu L.; Kovalenko M. V.; Yakunin S.; Cervellino A.; Billinge S. J. L.; Terban M. W.; Pedersen J. S.; Masciocchi N.; Guagliardi A. Coherent Nanotwins and Dynamic Disorder in Cesium Lead Halide Perovskite Nanocrystals. ACS Nano 2017, 11, 3819–3831. 10.1021/acsnano.7b00017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lanigan-Atkins T.; He X.; Krogstad M. J.; Pajerowski D. M.; Abernathy D. L.; Xu G. N. M. N.; Xu Z.; Chung D.-Y.; Kanatzidis M. G.; Rosenkranz S.; et al. Two-Dimensional Overdamped Fluctuations of the Soft Perovskite Lattice in CsPbBr3. Nat. Mater. 2021, 20, 977–983. 10.1038/s41563-021-00947-y. [DOI] [PubMed] [Google Scholar]
- Onoda-Yamamuro N.; Matsuo T.; Suga H. Calorimetric and IR Spectroscopic Studies of Phase Transitions in Methylammonium Trihalogenoplumbates (II). J. Phys. Chem. Solids 1990, 51, 1383–1395. 10.1016/0022-3697(90)90021-7. [DOI] [Google Scholar]
- Anusca I.; Balciunas S.; Gemeiner P.; Svirskas S.; Sanlialp M.; Lackner G.; Fettkenhauer C.; Belovickis J.; Samulionis V.; Ivanov M.; Dkhil B.; Banys J.; Shvartsman V. V.; Lupascu D. C. Dielectric Response: Answer to Many Questions in the Methylammonium Lead Halide Solar Cell Absorbers. Adv. Energy Mater. 2017, 7, 1700600 10.1002/aenm.201700600. [DOI] [Google Scholar]
- Maczka M.; Gagor A.; Zareba J. K.; Stefanska D.; Drozd M.; Balciunas S.; Simenas M.; Banys J.; Sieradzki A. Three-Dimensional Perovskite Methylhydrazinium Lead Chloride with Two Polar Phases and Unusual Second-Harmonic Generation Bistability above Room Temperature. Chem. Mater. 2020, 32, 4072–4082. 10.1021/acs.chemmater.0c00973. [DOI] [Google Scholar]
- Maczka M.; Ptak M.; Gagor A.; Stefanska D.; Zareba J. K.; Sieradzki A. Methylhydrazinium Lead Bromide: Noncentrosymmetric Three-Dimensional Perovskite with Exceptionally Large Framework Distortion and Green Photoluminescence. Chem. Mater. 2020, 32, 1667–1673. 10.1021/acs.chemmater.9b05273. [DOI] [Google Scholar]
- Kawachi S.; Atsumi M.; Saito N.; Ohashi N.; Murakami Y.; Yamaura J.-i. Structural and Thermal Properties in Formamidinium and Cs-Mixed Lead Halides. J. Phys. Chem. Lett. 2019, 10, 6967–6972. 10.1021/acs.jpclett.9b02750. [DOI] [PubMed] [Google Scholar]
- Schönhals A.; Kremer F.. Broadband Dielectric Spectroscopy, 1st ed.; Springer-Verlag Berlin Heidelberg, 2003. [Google Scholar]
- Grigas J.; Brilingas A.; Kalesinskas V. Microwave and Millimetre Wave Dielectric Spectroscopy of Ferroelectrics. Ferroelectrics 1990, 107, 61–66. 10.1080/00150199008221514. [DOI] [Google Scholar]
- Banys J.; Lapinskas S.; Rudys S.; Greicius S.; Grigalaitis R. High Frequency Measurements of Ferroelecrics and Related Materials in Coaxial Line. Ferroelectrics 2011, 414, 64–69. 10.1080/00150193.2011.577308. [DOI] [Google Scholar]
- Onoda-Yamamuro N.; Matsuo T.; Suga H. Dielectric Study of CH3NH3PbX3 (X = Cl, Br, I). J. Phys. Chem. Solids 1992, 53, 935–939. 10.1016/0022-3697(92)90121-S. [DOI] [Google Scholar]
- Svirskas S.; Balciunas S.; Simenas M.; Usevicius G.; Kinka M.; Velicka M.; Kubicki D.; Castillo M. E.; Karabanov A.; Shvartsman V. V.; de Rosario Soares M.; Sablinskas V.; Salak A. N.; Lupascu D. C.; Banys J. Phase Transitions, Screening and Dielectric Response of CsPbBr3. J. Mater. Chem. A 2020, 8, 14015–14022. 10.1039/D0TA04155F. [DOI] [Google Scholar]
- Wilson J. N.; Frost J. M.; Wallace S. K.; Walsh A. Dielectric and Ferroic Properties of Metal Halide Perovskites. APL Mater. 2019, 7, 010901 10.1063/1.5079633. [DOI] [Google Scholar]
- Ibaceta-Jaña J.; Muydinov R.; Rosado P.; Mirhosseini H.; Chugh M.; Nazarenko O.; Dirin D. N.; Heinrich D.; Wagner M. R.; Kühne T. D.; et al. Vibrational Dynamics in Lead Halide Hybrid Perovskites Investigated by Raman Spectroscopy. Phys. Chem. Chem. Phys. 2020, 22, 5604–5614. 10.1039/C9CP06568G. [DOI] [PubMed] [Google Scholar]
- Kontos A. G.; Manolis G. K.; Kaltzoglou A.; Palles D.; Kamitsos E. I.; Kanatzidis M. G.; Falaras P. Halogen-NH2+ Interaction, Temperature-Induced Phase Transition, and Ordering in (NH2CHNH2)PbX3 (X = Cl, Br, I) Hybrid Perovskites. J. Phys. Chem. C 2020, 124, 8479–8487. 10.1021/acs.jpcc.9b11334. [DOI] [Google Scholar]
- Maczka M.; Zienkiewicz J. A.; Ptak M. Comparative Studies of Phonon Properties of Three-Dimensional Hybrid Organic-Inorganic Perovskites Comprising Methylhydrazinium, Methylammonium, and Formamidinium Cations. J. Phys. Chem. C 2022, 126, 4048–4056. 10.1021/acs.jpcc.1c09671. [DOI] [Google Scholar]
- Maczka M.; Ptak M. Lattice Dynamics and Structural Phase Transitions in Two-Dimensional Ferroelectric Methylhydrazinium Lead Bromide Investigated Using Raman and IR Spectroscopy. J. Phys. Chem. C 2022, 126, 7991–7998. 10.1021/acs.jpcc.2c01731. [DOI] [Google Scholar]
- Spirito D.; Asensio Y.; Hueso L. E.; Martín-García B. Raman Spectroscopy in Layered Hybrid Organic-Inorganic Metal Halide Perovskites. JPhys. Mater. 2022, 5, 034004 10.1088/2515-7639/ac7977. [DOI] [Google Scholar]
- Leguy A. M. A.; Goñi A. R.; Frost J. M.; Skelton J.; Brivio F.; Rodríguez-Martínez X.; Weber O. J.; Pallipurath A.; Alonso M. I.; Campoy-Quiles M.; et al. Dynamic Disorder, Phonon Lifetimes, and the Assignment of Modes to the Vibrational Spectra of Methylammonium Lead Halide Perovskites. Phys. Chem. Chem. Phys. 2016, 18, 27051–27066. 10.1039/C6CP03474H. [DOI] [PubMed] [Google Scholar]
- Schuck G.; Többens D. M.; Koch-Müller M.; Efthimiopoulos I.; Schorr S. Infrared Spectroscopic Study of Vibrational Modes across the Orthorhombic-Tetragonal Phase Transition in Methylammonium Lead Halide Single Crystals. J. Phys. Chem. C 2018, 122, 5227–5237. 10.1021/acs.jpcc.7b11499. [DOI] [Google Scholar]
- Ruan S.; McMeekin D. P.; Fan R.; Webster N. A. S.; Ebendorff-Heidepriem H.; Cheng Y.-B.; Lu J.; Ruan Y.; McNeill C. R. Raman Spectroscopy of Formamidinium-Based Lead Halide Perovskite Single Crystals. J. Phys. Chem. C 2020, 124, 2265–2272. 10.1021/acs.jpcc.9b08917. [DOI] [Google Scholar]
- Sendner M.; Nayak P. K.; Egger D. A.; Beck S.; Müller C.; Epding B.; Kowalsky W.; Kronik L.; Snaith H. J.; Pucci A.; et al. Optical Phonons in Methylammonium Lead Halide Perovskites and Implications for Charge Transport. Mater. Horiz. 2016, 3, 613–620. 10.1039/C6MH00275G. [DOI] [Google Scholar]
- Nakada K.; Matsumoto Y.; Shimoi Y.; Yamada K.; Furukawa Y. Temperature-Dependent Evolution of Raman Spectra of Methylammonium Lead Halide Perovskites, CH3NH3PbX3 (X = I, Br). Molecules 2019, 24, 626. 10.3390/molecules24030626. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Piveteau L.; Morad V.; Kovalenko M. V. Solid-State NMR and NQR Spectroscopy of Lead-Halide Perovskite Materials. J. Am. Chem. Soc. 2020, 142, 19413–19437. 10.1021/jacs.0c07338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mishra A.; Hope M. A.; Grätzel M.; Emsley L. A Complete Picture of Cation Dynamics in Hybrid Perovskite Materials from Solid-State NMR Spectroscopy. J. Am. Chem. Soc. 2023, 145, 978–990. 10.1021/jacs.2c10149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fykouras K.; Lahnsteiner J.; Leupold N.; Tinnemans P.; Moos R.; Panzer F.; de Wijs G. A.; Bokdam M.; Grüninger H.; Kentgens A. P. M. Disorder to Order: How Halide Mixing in MAPbI3–xBrx Perovskites Restricts MA Dynamics. J. Mater. Chem. A 2023, 11, 4587–4597. 10.1039/D2TA09069D. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Grüninger H.; Bokdam M.; Leupold N.; Tinnemans P.; Moos R.; De Wijs G. A.; Panzer F.; Kentgens A. P. M. Microscopic (Dis)order and Dynamics of Cations in Mixed FA/MA Lead Halide Perovskites. J. Phys. Chem. C 2021, 125, 1742–1753. 10.1021/acs.jpcc.0c10042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kubicki D. J.; Prochowicz D.; Hofstetter A.; Zakeeruddin S. M.; Grätzel M.; Emsley L. Phase Segregation in Cs-, Rb- and K-Doped Mixed-Cation (MA)x(FA)1–xPbI3 Hybrid Perovskites from Solid-State NMR. J. Am. Chem. Soc. 2017, 139, 14173–14180. 10.1021/jacs.7b07223. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Franssen W. M. J.; van Heumen C. M. M.; Kentgens A. P. M. Structural Investigations of MA1–xDMAxPbI3 Mixed-Cation Perovskites. Inorg. Chem. 2020, 59, 3730–3739. 10.1021/acs.inorgchem.9b03380. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fabini D. H.; Siaw T. A.; Stoumpos C. C.; Laurita G.; Olds D.; Page K.; Hu J. G.; Kanatzidis M. G.; Han S.; Seshadri R. Universal Dynamics of Molecular Reorientation in Hybrid Lead Iodide Perovskites. J. Am. Chem. Soc. 2017, 139, 16875–16884. 10.1021/jacs.7b09536. [DOI] [PubMed] [Google Scholar]
- Dahlman C. J.; Kubicki D. J.; Reddy G. N. M. Interfaces in Metal Halide Perovskites Probed by Solid-State NMR Spectroscopy. J. Mater. Chem. A 2021, 9, 19206–19244. 10.1039/D1TA03572J. [DOI] [Google Scholar]
- Kruteva M. Dynamics Studied by Quasielastic Neutron Scattering (QENS). Adsorption 2021, 27, 875–889. 10.1007/s10450-020-00295-4. [DOI] [Google Scholar]
- Telling M. T.A Practical Guide to Quasi-Elastic Neutron Scattering; Royal Society of Chemistry, 2020. [Google Scholar]
- Leguy A. M. A.; Frost J. M.; McMahon A. P.; Sakai V. G.; Kockelmann W.; Law C.; Li X.; Foglia F.; Walsh A.; O’Regan B. C.; et al. The Dynamics of Methylammonium Ions in Hybrid Organic-Inorganic Perovskite Solar Cells. Nat. Commun. 2015, 6, 7124. 10.1038/ncomms8124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen T.; Foley B. J.; Ipek B.; Tyagi M.; Copley J. R. D.; Brown C. M.; Choi J. J.; Lee S.-H. Rotational Dynamics of Organic Cations in the CH3NH3PbI3 Perovskite. Phys. Chem. Chem. Phys. 2015, 17, 31278–31286. 10.1039/C5CP05348J. [DOI] [PubMed] [Google Scholar]
- Li J.; Bouchard M.; Reiss P.; Aldakov D.; Pouget S.; Demadrille R.; Aumaitre C.; Frick B.; Djurado D.; Rossi M.; et al. Activation Energy of Organic Cation Rotation in CH3NH3PbI3 and CD3NH3PbI3: Quasi-Elastic Neutron Scattering Measurements and First-Principles Analysis Including Nuclear Quantum Effects. J. Phys. Chem. Lett. 2018, 9, 3969–3977. 10.1021/acs.jpclett.8b01321. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li B.; Kawakita Y.; Liu Y.; Wang M.; Matsuura M.; Shibata K.; Ohira-Kawamura S.; Yamada T.; Lin S.; Nakajima K.; Liu S. Polar Rotor Scattering as Atomic-Level Origin of Low Mobility and Thermal Conductivity of Perovskite CH3NH3PbI3. Nat. Commun. 2017, 8, 16086 10.1038/ncomms16086. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sharma V. K.; Mukhopadhyay R.; Mohanty A.; Sakai V. G.; Tyagi M.; Sarma D. D. Contrasting Effects of FA Substitution on MA/FA Rotational Dynamics in FAxMA1–xPbI3. J. Phys. Chem. C 2021, 125, 13666–13676. 10.1021/acs.jpcc.1c03280. [DOI] [Google Scholar]
- Zhou X.; Jankowska J.; Dong H.; Prezhdo O. V. Recent Theoretical Progress in The Development of Perovskite Photovoltaic Materials. J. Energy Chem. 2018, 27, 637–649. 10.1016/j.jechem.2017.10.010. [DOI] [Google Scholar]
- Motta C.; El-Mellouhi F.; Sanvito S. Exploring the Cation Dynamics in Lead-Bromide Hybrid Perovskites. Phys. Rev. B 2016, 93, 235412 10.1103/PhysRevB.93.235412. [DOI] [Google Scholar]
- Sukmas W.; Pinsook U.; Tsuppayakorn-aek P.; Pakornchote T.; Sukserm A.; Bovornratanaraks T. Organic Molecule Orientations and Rashba-Dresselhaus Effect in α-Formamidinium Lead Iodide. J. Phys. Chem. C 2019, 123, 16508–16515. 10.1021/acs.jpcc.9b02140. [DOI] [Google Scholar]
- Weller M. T.; Weber O. J.; Frost J. M.; Walsh A. Cubic Perovskite Structure of Black Formamidinium Lead Iodide, α-[HC(NH2)2]PbI3, at 298 K. J. Phys. Chem. Lett. 2015, 6, 3209–3212. 10.1021/acs.jpclett.5b01432. [DOI] [Google Scholar]
- Weber O. J.; Ghosh D.; Gaines S.; Henry P. F.; Walker A. B.; Islam M. S.; Weller M. T. Phase Behavior and Polymorphism of Formamidinium Lead Iodide. Chem. Mater. 2018, 30, 3768–3778. 10.1021/acs.chemmater.8b00862. [DOI] [Google Scholar]
- Carignano M. A.; Saeed Y.; Aravindh S. A.; Roqan I. S.; Even J.; Katan C. A Close Examination of the Structure and Dynamics of HC(NH2)2PbI3 by MD Simulations and Group Theory. Phys. Chem. Chem. Phys. 2016, 18, 27109–27118. 10.1039/C6CP02917E. [DOI] [PubMed] [Google Scholar]
- Bakulin A. A.; Selig O.; Bakker H. J.; Rezus Y. L.; Müller C.; Glaser T.; Lovrincic R.; Sun Z.; Chen Z.; Walsh A.; et al. Real-Time Observation of Organic Cation Reorientation in Methylammonium Lead Iodide Perovskites. J. Phys. Chem. Lett. 2015, 6, 3663–3669. 10.1021/acs.jpclett.5b01555. [DOI] [PubMed] [Google Scholar]
- Drużbicki K.; Lavén R.; Armstrong J.; Malavasi L.; Fernandez-Alonso F.; Karlsson M. Cation Dynamics and Structural Stabilization in Formamidinium Lead Iodide Perovskites. J. Phys. Chem. Lett. 2021, 12, 3503–3508. 10.1021/acs.jpclett.1c00616. [DOI] [PubMed] [Google Scholar]
- Ghosh D.; Walsh Atkins P.; Islam M. S.; Walker A. B.; Eames C. Good Vibrations: Locking of Octahedral Tilting in Mixed-Cation Iodide Perovskites for Solar Cells. ACS Energy Lett. 2017, 2, 2424–2429. 10.1021/acsenergylett.7b00729. [DOI] [Google Scholar]
- Simenas M.; Balciunas S.; Maczka M.; Banys J.; Tornau E. E. Exploring the Antipolar Nature of Methylammonium Lead Halides: A Monte Carlo and Pyrocurrent Study. J. Phys. Chem. Lett. 2017, 8, 4906–4911. 10.1021/acs.jpclett.7b02239. [DOI] [PubMed] [Google Scholar]
- Simenas M.; Balciunas S.; Maczka M.; Banys J. Phase Transition Model of FA Cation Ordering in FAPbX3 (X = Br, I) Hybrid Perovskites. J. Mater. Chem. C 2022, 10, 5210–5217. 10.1039/D2TC00559J. [DOI] [Google Scholar]
- Tan L. Z.; Zheng F.; Rappe A. M. Intermolecular Interactions in Hybrid Perovskites Understood from a Combined Density Functional Theory and Effective Hamiltonian Approach. ACS Energy Lett. 2017, 2, 937–942. 10.1021/acsenergylett.7b00159. [DOI] [Google Scholar]
- Lahnsteiner J.; Jinnouchi R.; Bokdam M. Long-Range Order Imposed by Short-Range Interactions in Methylammonium Lead Iodide: Comparing Point-Dipole Models to Machine-Learning Force Fields. Phys. Rev. B 2019, 100, 094106 10.1103/PhysRevB.100.094106. [DOI] [Google Scholar]
- Park H.; Ali A.; Mall R.; Bensmail H.; Sanvito S.; El-Mellouhi F. Data-Driven Enhancement of Cubic Phase Stability in Mixed-Cation Perovskites. Mach. Learn.: Sci. Technol. 2021, 2, 025030 10.1088/2632-2153/abdaf9. [DOI] [Google Scholar]
- Yang J.; Wang Y.; Wu T.; Li S. Correlating the Composition-Dependent Structural and Electronic Dynamics of Inorganic Mixed Halide Perovskites. Chem. Mater. 2020, 32, 2470–2481. 10.1021/acs.chemmater.9b04995. [DOI] [Google Scholar]
- Anand D. V.; Xu Q.; Wee J.; Xia K.; Sum T. C. Topological Feature Engineering for Machine Learning Based Halide Perovskite Materials Design. npj Comput. Mater. 2022, 8, 203. 10.1038/s41524-022-00883-8. [DOI] [Google Scholar]
- Ahmadi M.; Ziatdinov M.; Zhou Y.; Lass E. A.; Kalinin S. V. Machine Learning for High-Throughput Experimental Exploration of Metal Halide Perovskites. Joule 2021, 5, 2797–2822. 10.1016/j.joule.2021.10.001. [DOI] [Google Scholar]
- Tao Q.; Xu P.; Li M.; Lu W. Machine Learning for Perovskite Materials Design and Discovery. npj Comput. Mater. 2021, 7, 23. 10.1038/s41524-021-00495-8. [DOI] [Google Scholar]
- Wiktor J.; Fransson E.; Kubicki D.; Erhart P. Quantifying Dynamic Tilting in Halide Perovskites: Chemical Trends and Local Correlations. Chem. Mater. 2023, 35, 6737–6744. 10.1021/acs.chemmater.3c00933. [DOI] [Google Scholar]
- Maczka M.; Ptak M.; Gagor A.; Zareba J. K.; Liang X.; Balciunas S.; Semenikhin O. A.; Kucheriv O. I.; Gural’skiy I.’y. A.; Shova S.; Walsh A.; Banys J.; Simenas M.. Phase Transitions, Dielectric Response, and Nonlinear Optical Properties of Aziridinium Lead Halide Perovskites. Chem. Mater. 2023.359725. 10.1021/acs.chemmater.3c02200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jauch W.; Palmer A. Anomalous Zero-Point Motion in SrTiO3: Results from γ-Ray Diffraction. Phys. Rev. B 1999, 60, 2961–2963. 10.1103/PhysRevB.60.2961. [DOI] [Google Scholar]
- Glazer A. M. Simple Ways of Determining Perovskite Structures. Acta Crystallogr., Sect. A 1975, 31, 756–762. 10.1107/S0567739475001635. [DOI] [Google Scholar]
- Poglitsch A.; Weber D. Dynamic Disorder in Methylammoniumtrihalogenoplumbates (II) Observed by Millimeter-Wave Spectroscopy. J. Chem. Phys. 1987, 87, 6373–6378. 10.1063/1.453467. [DOI] [Google Scholar]
- Whitfield P. S.; Herron N.; Guise W. E.; Page K.; Cheng Y. Q.; Milas I.; Crawford M. K. Structures, Phase Transitions and Tricritical Behavior of the Hybrid Perovskite Methyl Ammonium Lead Iodide. Sci. Rep. 2016, 6, 35685 10.1038/srep35685. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yamada Y.; Yamada T.; Phuong L. Q.; Maruyama N.; Nishimura H.; Wakamiya A.; Murata Y.; Kanemitsu Y. Dynamic Optical Properties of CH3NH3PbI3 Single Crystals As Revealed by One- and Two-Photon Excited Photoluminescence Measurements. J. Am. Chem. Soc. 2015, 137, 10456–10459. 10.1021/jacs.5b04503. [DOI] [PubMed] [Google Scholar]
- Szafrański M.; Katrusiak A. Mechanism of Pressure-Induced Phase Transitions, Amorphization, and Absorption-Edge Shift in Photovoltaic Methylammonium Lead Iodide. J. Phys. Chem. Lett. 2016, 7, 3458–3466. 10.1021/acs.jpclett.6b01648. [DOI] [PubMed] [Google Scholar]
- Frost J. M.; Butler K. T.; Walsh A. Molecular Ferroelectric Contributions to Anomalous Hysteresis in Hybrid Perovskite Solar Cells. APL Mater. 2014, 2, 081506 10.1063/1.4890246. [DOI] [Google Scholar]
- Kutes Y.; Ye L.; Zhou Y.; Pang S.; Huey B. D.; Padture N. P. Direct Observation of Ferroelectric Domains in Solution-Processed CH3NH3PbI3 Perovskite Thin Films. J. Phys. Chem. Lett. 2014, 5, 3335–3339. 10.1021/jz501697b. [DOI] [PubMed] [Google Scholar]
- Chen H.-W.; Sakai N.; Ikegami M.; Miyasaka T. Emergence of Hysteresis and Transient Ferroelectric Response in Organo-Lead Halide Perovskite Solar Cells. J. Phys. Chem. Lett. 2015, 6, 164–169. 10.1021/jz502429u. [DOI] [PubMed] [Google Scholar]
- Kim H.-S.; Kim S. K.; Kim B. J.; Shin K.-S.; Gupta M. K.; Jung H. S.; Kim S.-W.; Park N.-G. Ferroelectric Polarization in CH3NH3PbI3 Perovskite. J. Phys. Chem. Lett. 2015, 6, 1729–1735. 10.1021/acs.jpclett.5b00695. [DOI] [PubMed] [Google Scholar]
- Rakita Y.; Bar-Elli O.; Meirzadeh E.; Kaslasi H.; Peleg Y.; Hodes G.; Lubomirsky I.; Oron D.; Ehre D.; Cahen D. Tetragonal CH3NH3PbI3 is Ferroelectric. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, E5504–E5512. 10.1073/pnas.1702429114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rakita Y.; Meirzadeh E.; Bendikov T.; Kalchenko V.; Lubomirsky I.; Hodes G.; Ehre D.; Cahen D. CH3NH3PbBr3 is Not Pyroelectric, Excluding Ferroelectric-Enhanced Photovoltaic Performance. APL Mater. 2016, 4, 051101 10.1063/1.4949760. [DOI] [Google Scholar]
- Hoque M. N. F.; Yang M.; Li Z.; Islam N.; Pan X.; Zhu K.; Fan Z. Polarization and Dielectric Study of Methylammonium Lead Iodide Thin Film to Reveal its Nonferroelectric Nature under Solar Cell Operating Conditions. ACS Energy Lett. 2016, 1, 142–149. 10.1021/acsenergylett.6b00093. [DOI] [Google Scholar]
- Gómez A.; Wang Q.; Goñi A. R.; Campoy-Quiles M.; Abate A. Ferroelectricity-Free Lead Halide Perovskites. Energy Environ. Sci. 2019, 12, 2537–2547. 10.1039/C9EE00884E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ambrosio F.; De Angelis F.; Goñi A. R. The Ferroelectric-Ferroelastic Debate about Metal Halide Perovskites. J. Phys. Chem. Lett. 2022, 13, 7731–7740. 10.1021/acs.jpclett.2c01945. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mayers M. Z.; Tan L. Z.; Egger D. A.; Rappe A. M.; Reichman D. R. How Lattice and Charge Fluctuations Control Carrier Dynamics in Halide Perovskites. Nano Lett. 2018, 18, 8041–8046. 10.1021/acs.nanolett.8b04276. [DOI] [PubMed] [Google Scholar]
- Zhu T.; Ertekin E. Mixed Phononic and Non-Phononic Transport in Hybrid Lead Halide Perovskites: Glass-Crystal Duality, Dynamical Disorder, and Anharmonicity. Energy Environ. Sci. 2019, 12, 216–229. 10.1039/C8EE02820F. [DOI] [Google Scholar]
- Comin R.; Crawford M. K.; Said A. H.; Herron N.; Guise W. E.; Wang X.; Whitfield P. S.; Jain A.; Gong X.; McGaughey A. J. H.; et al. Lattice Dynamics and the Nature of Structural Transitions in Organolead Halide Perovskites. Phys. Rev. B 2016, 94, 094301 10.1103/PhysRevB.94.094301. [DOI] [Google Scholar]
- Whalley L. D.; Frost J. M.; Jung Y.-K.; Walsh A. Perspective: Theory and simulation of hybrid halide perovskites. J. Chem. Phys. 2017, 146, 220901 10.1063/1.4984964. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yaffe O.; Guo Y.; Tan L. Z.; Egger D. A.; Hull T.; Stoumpos C. C.; Zheng F.; Heinz T. F.; Kronik L.; Kanatzidis M. G.; et al. Local Polar Fluctuations in Lead Halide Perovskite Crystals. Phys. Rev. Lett. 2017, 118, 136001 10.1103/PhysRevLett.118.136001. [DOI] [PubMed] [Google Scholar]
- Songvilay M.; Giles-Donovan N.; Bari M.; Ye Z.-G.; Minns J. L.; Green M. A.; Xu G.; Gehring P. M.; Schmalzl K.; Ratcliff W. D.; et al. Common Acoustic Phonon Lifetimes in Inorganic and Hybrid Lead Halide Perovskites. Phys. Rev. Mater. 2019, 3, 093602 10.1103/PhysRevMaterials.3.093602. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Weadock N. J.; Gehring P. M.; Gold-Parker A.; Smith I. C.; Karunadasa H. I.; Toney M. F. Test of the Dynamic-Domain and Critical Scattering Hypotheses in Cubic Methylammonium Lead Triiodide. Phys. Rev. Lett. 2020, 125, 075701 10.1103/PhysRevLett.125.075701. [DOI] [Google Scholar]
- Guo Y.; Yaffe O.; Paley D. W.; Beecher A. N.; Hull T. D.; Szpak G.; Owen J. S.; Brus L. E.; Pimenta M. A. Interplay Between Organic Cations and Inorganic Framework and Incommensurability in Hybrid Lead-Halide Perovskite CH3NH3PbBr3. Phys. Rev. Mater. 2017, 1, 042401 10.1103/PhysRevMaterials.1.042401. [DOI] [Google Scholar]
- Kawamura Y.; Mashiyama H. Modulated Structure in Phase II of CH3NH3PbCl3. J. Korean Phys. Soc. 1999, 35, S1437–S1440. [Google Scholar]
- Chi L.; Swainson I.; Cranswick L.; Her J.-H.; Stephens P.; Knop O. The Ordered Phase of Methylammonium Lead Chloride CH3ND3PbCl3. J. Solid State Chem. 2005, 178, 1376–1385. 10.1016/j.jssc.2004.12.037. [DOI] [Google Scholar]
- Cordero F.; Craciun F.; Trequattrini F.; Generosi A.; Paci B.; Paoletti A. M.; Pennesi G. Stability of Cubic FAPbI3 from X-ray Diffraction, Anelastic, and Dielectric Measurements. J. Phys. Chem. Lett. 2019, 10, 2463–2469. 10.1021/acs.jpclett.9b00896. [DOI] [PubMed] [Google Scholar]
- Chen T.; Foley B. J.; Park C.; Brown C. M.; Harriger L. W.; Lee J.; Ruff J.; Yoon M.; Choi J. J.; Lee S.-H. Entropy-Driven Structural Transition and Kinetic Trapping in Formamidinium Lead Iodide Perovskite. Sci. Adv. 2016, 2, e1601650 10.1126/sciadv.1601650. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen T.; Chen W.-L.; Foley B. J.; Lee J.; Ruff J. P. C.; Ko J. Y. P.; Brown C. M.; Harriger L. W.; Zhang D.; Park C.; et al. Origin of Long Lifetime of Band-Edge Charge Carriers in Organic-Inorganic Lead Iodide Perovskites. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, 7519–7524. 10.1073/pnas.1704421114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kennedy B. J.; Prodjosantoso A. K.; Howard C. J. Powder Neutron Diffraction Study of the High Temperature Phase Transitions in NaTaO3. J. Phys.: Condens. Matter 1999, 11, 6319. 10.1088/0953-8984/11/33/302. [DOI] [Google Scholar]
- Sun S.; Deng Z.; Wu Y.; Wei F.; Halis Isikgor F.; Brivio F.; Gaultois M. W.; Ouyang J.; Bristowe P. D.; Cheetham A. K.; et al. Variable Temperature and High-Pressure Crystal Chemistry of Perovskite Formamidinium Lead Iodide: a Single Crystal X-Ray Diffraction and Computational Study. Chem. Commun. 2017, 53, 7537–7540. 10.1039/C7CC00995J. [DOI] [PubMed] [Google Scholar]
- Fabini D. H.; Stoumpos C. C.; Laurita G.; Kaltzoglou A.; Kontos A. G.; Falaras P.; Kanatzidis M. G.; Seshadri R. Reentrant Structural and Optical Properties and Large Positive Thermal Expansion in Perovskite Formamidinium Lead Iodide. Angew. Chem., Int. Ed. 2016, 55, 15392–15396. 10.1002/anie.201609538. [DOI] [PubMed] [Google Scholar]
- Franz A.; Többens D. M.; Lehmann F.; Kärgell M.; Schorr S. The Influence of Deuteration on the Crystal Structure of Hybrid Halide Perovskites: a Temperature-Dependent Neutron Diffraction Study of FAPbBr3. Acta Crystallogr., Sect. B 2020, 76, 267–274. 10.1107/S2052520620002620. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mozur E. M.; Trowbridge J. C.; Maughan A. E.; Gorman M. J.; Brown C. M.; Prisk T. R.; Neilson J. R. Dynamical Phase Transitions and Cation Orientation-Dependent Photoconductivity in CH(NH2)2PbBr3. ACS Mater. Lett. 2019, 1, 260–264. 10.1021/acsmaterialslett.9b00209. [DOI] [Google Scholar]
- Maczka M.; Ptak M. Temperature-Dependent Raman Studies of FAPbBr3 and MAPbBr3 Perovskites: Effect of Phase Transitions on Molecular Dynamics and Lattice Distortion. Solids 2022, 3, 111–121. 10.3390/solids3010008. [DOI] [Google Scholar]
- Dimitrovska-Lazova S.; Bukleski M.; Tzvetkov P.; Pecovska-Gjorgjevich M.; Kovacheva D.; Aleksovska S. The Mechanism of the Isostructural Phase Transition in C(NH2)3PbI3 as a Guide for Understanding the Properties of the New Phase. Mater. Chem. Phys. 2022, 275, 125240 10.1016/j.matchemphys.2021.125240. [DOI] [Google Scholar]
- Yesudhas S.; Burns R.; Lavina B.; Tkachev S. N.; Sun J.; Ullrich C. A.; Guha S. Coupling of Organic Cation and Inorganic Lattice in Methylammonium Lead Halide Perovskites: Insights into a Pressure-Induced Isostructural Phase Transition. Phys. Rev. Mater. 2020, 4, 105403 10.1103/PhysRevMaterials.4.105403. [DOI] [Google Scholar]
- Govinda S.; Kore B. P.; Swain D.; Hossain A.; De C.; Guru Row T. N.; Sarma D. D. Critical Comparison of FAPbX3 and MAPbX3 (X = Br and Cl): How Do They Differ?. J. Phys. Chem. C 2018, 122, 13758–13766. 10.1021/acs.jpcc.8b00602. [DOI] [Google Scholar]
- Sharma V. K.; Mukhopadhyay R.; Mohanty A.; García Sakai V.; Tyagi M.; Sarma D. D. Influence of the Halide Ion on the A-Site Dynamics in FAPbX3 (X = Br and Cl). J. Phys. Chem. C 2022, 126, 7158–7168. 10.1021/acs.jpcc.2c00968. [DOI] [Google Scholar]
- Møller C. K. Crystal Structure and Photoconductivity of Caesium Plumbohalides. Nature 1958, 182, 1436–1436. 10.1038/1821436a0. [DOI] [Google Scholar]
- Stoumpos C. C.; Kanatzidis M. G. The Renaissance of Halide Perovskites and Their Evolution as Emerging Semiconductors. Acc. Chem. Res. 2015, 48, 2791–2802. 10.1021/acs.accounts.5b00229. [DOI] [PubMed] [Google Scholar]
- Marronnier A.; Roma G.; Boyer-Richard S.; Pedesseau L.; Jancu J.-M.; Bonnassieux Y.; Katan C.; Stoumpos C. C.; Kanatzidis M. G.; Even J. Anharmonicity and Disorder in the Black Phases of Cesium Lead Iodide Used for Stable Inorganic Perovskite Solar Cells. ACS Nano 2018, 12, 3477–3486. 10.1021/acsnano.8b00267. [DOI] [PubMed] [Google Scholar]
- Trots D.; Myagkota S. High-Temperature Structural Evolution of Caesium and Rubidium Triiodoplumbates. J. Phys. Chem. Solids 2008, 69, 2520–2526. 10.1016/j.jpcs.2008.05.007. [DOI] [Google Scholar]
- Zhao B.; Jin S.-F.; Huang S.; Liu N.; Ma J.-Y.; Xue D.-J.; Han Q.; Ding J.; Ge Q.-Q.; Feng Y.; et al. Thermodynamically Stable Orthorhombic γ-CsPbI3 Thin Films for High-Performance Photovoltaics. J. Am. Chem. Soc. 2018, 140, 11716–11725. 10.1021/jacs.8b06050. [DOI] [PubMed] [Google Scholar]
- Sutton R. J.; Filip M. R.; Haghighirad A. A.; Sakai N.; Wenger B.; Giustino F.; Snaith H. J. Cubic or Orthorhombic? Revealing the Crystal Structure of Metastable Black-Phase CsPbI3 by Theory and Experiment. ACS Energy Lett. 2018, 3, 1787–1794. 10.1021/acsenergylett.8b00672. [DOI] [Google Scholar]
- Cohen A. V.; Egger D. A.; Rappe A. M.; Kronik L. Breakdown of the Static Picture of Defect Energetics in Halide Perovskites: The Case of the Br Vacancy in CsPbBr3. J. Phys. Chem. Lett. 2019, 10, 4490–4498. 10.1021/acs.jpclett.9b01855. [DOI] [PubMed] [Google Scholar]
- Li Z.-G.; Zacharias M.; Zhang Y.; Wei F.; Qin Y.; Yang Y.-Q.; An L.-C.; Gao F.-F.; Li W.; Even J.; et al. Origin of Phase Transitions in Inorganic Lead Halide Perovskites: Interplay between Harmonic and Anharmonic Vibrations. ACS Energy Lett. 2023, 8, 3016–3024. 10.1021/acsenergylett.3c00881. [DOI] [Google Scholar]
- Stoumpos C. C.; Malliakas C. D.; Peters J. A.; Liu Z.; Sebastian M.; Im J.; Chasapis T. C.; Wibowo A. C.; Chung D. Y.; Freeman A. J.; et al. Crystal Growth of the Perovskite Semiconductor CsPbBr3: A New Material for High-Energy Radiation Detection. Cryst. Growth Des. 2013, 13, 2722–2727. 10.1021/cg400645t. [DOI] [Google Scholar]
- Cottingham P.; Brutchey R. L. Depressed Phase Transitions and Thermally Persistent Local Distortions in CsPbBr3 Quantum Dots. Chem. Mater. 2018, 30, 6711–6716. 10.1021/acs.chemmater.8b02295. [DOI] [Google Scholar]
- Sakata M.; Nishiwaki T.; Harada J. Neutron Diffraction Study of the Structure of Cubic CsPbBr3. J. Phys. Soc. Jpn. 1979, 47, 232–233. 10.1143/JPSJ.47.232. [DOI] [Google Scholar]
- Rodová M.; Brožek J.; Knížek K.; Nitsch K. Phase Transitions in Ternary Caesium Lead Bromide. J. Therm. Anal. Calorim. 2003, 71, 667–673. 10.1023/A:1022836800820. [DOI] [Google Scholar]
- Linaburg M. R.; McClure E. T.; Majher J. D.; Woodward P. M. Cs1–xRbxPbCl3 and Cs1–xRbxPbBr3 Solid Solutions: Understanding Octahedral Tilting in Lead Halide Perovskites. Chem. Mater. 2017, 29, 3507–3514. 10.1021/acs.chemmater.6b05372. [DOI] [Google Scholar]
- Harada J.; Sakata M.; Hoshino S.; Hirotsu S. Neutron Diffraction Study of the Structure in Cubic CsPbCl3. J. Phys. Soc. Jpn. 1976, 40, 212–218. 10.1143/JPSJ.40.212. [DOI] [Google Scholar]
- Gesi K.; Ozawa K.; Hirotsu S. Effect of Hydrostatic Pressure on the Structural Phase Transitions in CsPbCl3 and CsPbBr3. J. Phys. Soc. Jpn. 1975, 38, 463–466. 10.1143/JPSJ.38.463. [DOI] [Google Scholar]
- Hua G. L. Normal Vibration Modes and the Structural Phase Transitions in Caesium Trichloroplumbate CsPbCl3. J. Phys.: Condens. Matter 1991, 3, 1371. 10.1088/0953-8984/3/11/002. [DOI] [Google Scholar]
- Hirotsu S.; Sawada S. Crystal Growth and Phase Transitions of CsPbCl3. Phys. Lett. A 1969, 28, 762–763. 10.1016/0375-9601(69)90608-2. [DOI] [Google Scholar]
- Hirotsu S. Experimental Studies of Structural Phase Transitions in CsPbCl3. J. Phys. Soc. Jpn. 1971, 31, 552–560. 10.1143/JPSJ.31.552. [DOI] [Google Scholar]
- Zhang L.; Wang L.; Wang K.; Zou B. Pressure-Induced Structural Evolution and Optical Properties of Metal-Halide Perovskite CsPbCl3. J. Phys. Chem. C 2018, 122, 15220–15225. 10.1021/acs.jpcc.8b05397. [DOI] [Google Scholar]
- Szafrański M.; Katrusiak A.; Ståhl K. Time-Dependent Transformation Routes of Perovskites CsPbBr3 and CsPbCl3 Under High Pressure. J. Mater. Chem. A 2021, 9, 10769–10779. 10.1039/D1TA01875B. [DOI] [Google Scholar]
- Sharma S.; Weiden N.; Weiss A. Phase Transitions in CsSnCl3 and CsPbBr3. An NMR and NQR Study. Z. Naturforsch. A 1991, 46, 329–336. 10.1515/zna-1991-0406. [DOI] [Google Scholar]
- Kubicki D. J.; Prochowicz D.; Salager E.; Rakhmatullin A.; Grey C. P.; Emsley L.; Stranks S. D. Local Structure and Dynamics in Methylammonium, Formamidinium, and Cesium Tin(II) Mixed-Halide Perovskites from 119Sn Solid-State NMR. J. Am. Chem. Soc. 2020, 142, 7813–7826. 10.1021/jacs.0c00647. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Petrosova H. R.; Kucheriv O. I.; Shova S.; Gural’skiy I. A. Aziridinium Cation Templating 3D Lead Halide Hybrid Perovskites. Chem. Commun. 2022, 58, 5745–5748. 10.1039/D2CC01364A. [DOI] [PubMed] [Google Scholar]
- Cortecchia D.; Neutzner S.; Yin J.; Salim T.; Srimath Kandada A. R.; Bruno A.; Lam Y. M.; Martí-Rujas J.; Petrozza A.; Soci C. Structure-Controlled Optical Thermoresponse in Ruddlesden-Popper Layered Perovskites. APL Mater. 2018, 6, 114207 10.1063/1.5045782. [DOI] [Google Scholar]
- Li L.; Liu X.; Li Y.; Xu Z.; Wu Z.; Han S.; Tao K.; Hong M.; Luo J.; Sun Z. Two-Dimensional Hybrid Perovskite-Type Ferroelectric for Highly Polarization-Sensitive Shortwave Photodetection. J. Am. Chem. Soc. 2019, 141, 2623–2629. 10.1021/jacs.8b12948. [DOI] [PubMed] [Google Scholar]
- Li L.; Shang X.; Wang S.; Dong N.; Ji C.; Chen X.; Zhao S.; Wang J.; Sun Z.; Hong M.; et al. Bilayered Hybrid Perovskite Ferroelectric with Giant Two-Photon Absorption. J. Am. Chem. Soc. 2018, 140, 6806–6809. 10.1021/jacs.8b04014. [DOI] [PubMed] [Google Scholar]
- Koegel A. A.; Oswald I. W. H.; Rivera C.; Miller S. L.; Fallon M. J.; Prisk T. R.; Brown C. M.; Neilson J. R. Influence of Inorganic Layer Thickness on Methylammonium Dynamics in Hybrid Perovskite Derivatives. Chem. Mater. 2022, 34, 8316–8323. 10.1021/acs.chemmater.2c01868. [DOI] [Google Scholar]
- Li L.; Sun Z.; Wang P.; Hu W.; Wang S.; Ji C.; Hong M.; Luo J. Tailored Engineering of an Unusual (C4H9NH3)2(CH3NH3)2Pb3Br10 Two-Dimensional Multilayered Perovskite Ferroelectric for a High-Performance Photodetector. Angew. Chem., Int. Ed. 2017, 56, 12150–12154. 10.1002/anie.201705836. [DOI] [PubMed] [Google Scholar]
- Wu Z.; Liu X.; Ji C.; Li L.; Wang S.; Peng Y.; Tao K.; Sun Z.; Hong M.; Luo J. Discovery of an Above-Room-Temperature Antiferroelectric in Two-Dimensional Hybrid Perovskite. J. Am. Chem. Soc. 2019, 141, 3812–3816. 10.1021/jacs.8b13827. [DOI] [PubMed] [Google Scholar]
- Ji C.; Wang S.; Wang Y.; Chen H.; Li L.; Sun Z.; Sui Y.; Wang S.; Luo J. 2D Hybrid Perovskite Ferroelectric Enables Highly Sensitive X-Ray Detection with Low Driving Voltage. Adv. Funct. Mater. 2020, 30, 1905529 10.1002/adfm.201905529. [DOI] [Google Scholar]
- Wang S.; Liu X.; Li L.; Ji C.; Sun Z.; Wu Z.; Hong M.; Luo J. An Unprecedented Biaxial Trilayered Hybrid Perovskite Ferroelectric with Directionally Tunable Photovoltaic Effects. J. Am. Chem. Soc. 2019, 141, 7693–7697. 10.1021/jacs.9b02558. [DOI] [PubMed] [Google Scholar]
- Ma Y.; Wang J.; Guo W.; Han S.; Xu J.; Liu Y.; Lu L.; Xie Z.; Luo J.; Sun Z. The First Improper Ferroelectric of 2D Multilayered Hybrid Perovskite Enabling Strong Tunable Polarization-Directed Second Harmonic Generation Effect. Adv. Funct. Mater. 2021, 31, 2103012 10.1002/adfm.202103012. [DOI] [Google Scholar]
- Liu Y.; Han S.; Wang J.; Ma Y.; Guo W.; Huang X.-Y.; Luo J.-H.; Hong M.; Sun Z. Spacer Cation Alloying of a Homoconformational Carboxylate trans Isomer to Boost in-Plane Ferroelectricity in a 2D Hybrid Perovskite. J. Am. Chem. Soc. 2021, 143, 2130–2137. 10.1021/jacs.0c12513. [DOI] [PubMed] [Google Scholar]
- Han S.; Li M.; Liu Y.; Guo W.; Hong M.-C.; Sun Z.; Luo J. Tailoring of a Visible-Light-Absorbing Biaxial Ferroelectric Towards Broadband Self-Driven Photodetection. Nat. Commun. 2021, 12, 284. 10.1038/s41467-020-20530-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang S.; Li L.; Weng W.; Ji C.; Liu X.; Sun Z.; Lin W.; Hong M.; Luo J. Trilayered Lead Chloride Perovskite Ferroelectric Affording Self-Powered Visible-Blind Ultraviolet Photodetection with Large Zero-Bias Photocurrent. J. Am. Chem. Soc. 2020, 142, 55–59. 10.1021/jacs.9b10919. [DOI] [PubMed] [Google Scholar]
- Liu X.; Wang S.; Long P.; Li L.; Peng Y.; Xu Z.; Han S.; Sun Z.; Hong M.; Luo J. Polarization-Driven Self-Powered Photodetection in a Single-Phase Biaxial Hybrid Perovskite Ferroelectric. Angew. Chem., Int. Ed. 2019, 58, 14504–14508. 10.1002/anie.201907660. [DOI] [PubMed] [Google Scholar]
- Peng Y.; Liu X.; Sun Z.; Ji C.; Li L.; Wu Z.; Wang S.; Yao Y.; Hong M.; Luo J. Exploiting the Bulk Photovoltaic Effect in a 2D Trilayered Hybrid Ferroelectric for Highly Sensitive Polarized Light Detection. Angew. Chem., Int. Ed. 2020, 59, 3933–3937. 10.1002/anie.201915094. [DOI] [PubMed] [Google Scholar]
- Peng Y.; Bie J.; Liu X.; Li L.; Chen S.; Fa W.; Wang S.; Sun Z.; Luo J. Acquiring High-TC Layered Metal Halide Ferroelectrics via Cage-Confined Ethylamine Rotators. Angew. Chem., Int. Ed. 2021, 60, 2839–2843. 10.1002/anie.202011270. [DOI] [PubMed] [Google Scholar]
- Xu Z.; Dong X.; Wang L.; Wu H.; Liu Y.; Luo J.; Hong M.; Li L. Precisely Tailoring a FAPbI3-Derived Ferroelectric for Sensitive Self-Driven Broad-Spectrum Polarized Photodetection. J. Am. Chem. Soc. 2023, 145, 1524–1529. 10.1021/jacs.2c12300. [DOI] [PubMed] [Google Scholar]
- Liu Y.; Xu H.; Liu X.; Han S.; Guo W.; Ma Y.; Fan Q.; Hu X.; Sun Z.; Luo J. A Room-Temperature Antiferroelectric in Hybrid Perovskite Enables Highly Efficient Energy Storage at Low Electric Fields. Chem. Sci. 2022, 13, 13499–13506. 10.1039/D2SC05285G. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vasileiadou E. S.; Hadar I.; Kepenekian M.; Even J.; Tu Q.; Malliakas C. D.; Friedrich D.; Spanopoulos I.; Hoffman J. M.; Dravid V. P.; et al. Shedding Light on the Stability and Structure-Property Relationships of Two-Dimensional Hybrid Lead Bromide Perovskites. Chem. Mater. 2021, 33, 5085–5107. 10.1021/acs.chemmater.1c01129. [DOI] [Google Scholar]
- Li X.; Fu Y.; Pedesseau L.; Guo P.; Cuthriell S.; Hadar I.; Even J.; Katan C.; Stoumpos C. C.; Schaller R. D.; et al. Negative Pressure Engineering with Large Cage Cations in 2D Halide Perovskites Causes Lattice Softening. J. Am. Chem. Soc. 2020, 142, 11486–11496. 10.1021/jacs.0c03860. [DOI] [PubMed] [Google Scholar]
- Lyu F.; Zheng X.; Li Z.; Chen Z.; Shi R.; Wang Z.; Liu H.; Lin B.-L. Spatiodynamics, Photodynamics, and Their Correlation in Hybrid Perovskites. Chem. Mater. 2021, 33, 3524–3533. 10.1021/acs.chemmater.0c04500. [DOI] [Google Scholar]
- Jia Q.-Q.; Ni H.-F.; Lun M.-M.; Xie L.-Y.; Lu H.-F.; Fu D.-W.; Guo Q. Tunable Phase Transition Temperature and Nonlinear Optical Properties of Organic-Inorganic Hybrid Perovskites Enabled by Dimensional Engineering. J. Mater. Chem. C 2022, 10, 16330–16336. 10.1039/D2TC03497B. [DOI] [Google Scholar]
- Lu L.; Weng W.; Ma Y.; Liu Y.; Han S.; Liu X.; Xu H.; Lin W.; Sun Z.; Luo J. Anisotropy in a 2D Perovskite Ferroelectric Drives Self-Powered Polarization-Sensitive Photoresponse for Ultraviolet Solar-Blind Polarized-Light Detection. Angew. Chem., Int. Ed. 2022, 61, e202205030 10.1002/anie.202205030. [DOI] [PubMed] [Google Scholar]
- Xu Z.; Weng W.; Li Y.; Liu X.; Yang T.; Li M.; Huang X.; Luo J.; Sun Z. 3D-to-2D Dimensional Reduction for Exploiting a Multilayered Perovskite Ferroelectric toward Polarized-Light Detection in the Solar-Blind Ultraviolet Region. Angew. Chem., Int. Ed. 2020, 59, 21693–21697. 10.1002/anie.202009329. [DOI] [PubMed] [Google Scholar]
- Guo W.; Xu H.; Ma Y.; Liu Y.; Gao H.; Hu T.; Ren W.; Luo J.; Sun Z. Electrically Switchable Persistent Spin Texture in a Two-Dimensional Hybrid Perovskite Ferroelectric. Angew. Chem., Int. Ed. 2023, 62, e202300028 10.1002/anie.202300028. [DOI] [PubMed] [Google Scholar]
- Li M.; Xu Y.; Han S.; Xu J.; Xie Z.; Liu Y.; Xu Z.; Hong M.; Luo J.; Sun Z. Giant and Broadband Multiphoton Absorption Nonlinearities of a 2D Organometallic Perovskite Ferroelectric. Adv. Mater. 2020, 32, 2002972 10.1002/adma.202002972. [DOI] [PubMed] [Google Scholar]
- Hua L.; Wang J.; Liu Y.; Guo W.; Ma Y.; Xu H.; Han S.; Luo J.; Sun Z. Improper High-Tc Perovskite Ferroelectric with Dielectric Bistability Enables Broadband Ultraviolet-to-Infrared Photopyroelectric Effects. Adv. Sci. 2023, 10, 2301064 10.1002/advs.202301064. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang J.; Ma Y.; Wang Z.; Liu X.; Han S.; Liu Y.; Guo W.; Luo J.; Sun Z. Unusual Ferroelectric-Dependent Birefringence in 2D Trilayered Perovskite-Type Ferroelectric Exploited by Dimensional Tailoring. Matter 2022, 5, 194–205. 10.1016/j.matt.2021.10.020. [DOI] [Google Scholar]
- Guo W.; Xu H.; Ma Y.; Liu Y.; Wang B.; Tang L.; Hua L.; Luo J.; Sun Z. The Unprecedented Highest-Layer-Number Ferroelectric Semiconductor of 2D Homologous Single-Phase Perovskites Tailored by Regulating Thickness of Inorganic Frameworks. Adv. Funct. Mater. 2022, 32, 2207854 10.1002/adfm.202207854. [DOI] [Google Scholar]
- Zhuang J.-c.; Wei W.-j.; Song N.; Tang Y.-z.; Tan Y.-h.; Han D.-c.; Li Y.-k. A Narrow Bandgap of 2D Ruddlesden-Popper Bilayer Perovskite with Giant Entropy Change and Photoluminescence. Chem.—Eur. J. 2021, 27, 15716–15721. 10.1002/chem.202102550. [DOI] [PubMed] [Google Scholar]
- Li X.; Wu F.; Yao Y.; Wu W.; Ji C.; Li L.; Sun Z.; Luo J.; Liu X. Robust Spin-Dependent Anisotropy of Circularly Polarized Light Detection from Achiral Layered Hybrid Perovskite Ferroelectric Crystals. J. Am. Chem. Soc. 2022, 144, 14031–14036. 10.1021/jacs.2c06048. [DOI] [PubMed] [Google Scholar]
- Wang B.; Guo W.; Liu Y.; Xu H.; Fan Q.; Hua L.; Tang L.; Han S.; Luo J.; Sun Z. Exceptionally Strong Self-Powered Polarization Behaviors Involving with Bulk Photovoltaics in Highly Anisotropic 2D Hybrid Perovskite Ferroelectric Crystals. Chem. Mater. 2023, 35, 4007–4014. 10.1021/acs.chemmater.3c00277. [DOI] [Google Scholar]
- Han S.; Li L.; Ji C.; Liu X.; Wang G.-E.; Xu G.; Sun Z.; Luo J. Visible-Photoactive Perovskite Ferroelectric-Driven Self-Powered Gas Detection. J. Am. Chem. Soc. 2023, 145, 12853–12860. 10.1021/jacs.3c03719. [DOI] [PubMed] [Google Scholar]
- Wu Z.; Ji C.; Li L.; Kong J.; Sun Z.; Zhao S.; Wang S.; Hong M.; Luo J. Alloying n-Butylamine into CsPbBr3 To Give a Two-Dimensional Bilayered Perovskite Ferroelectric Material. Angew. Chem., Int. Ed. 2018, 57, 8140–8143. 10.1002/anie.201803716. [DOI] [PubMed] [Google Scholar]
- Ye H.; Peng Y.; Shang X.; Li L.; Yao Y.; Zhang X.; Zhu T.; Liu X.; Chen X.; Luo J. Self-Powered Visible-Infrared Polarization Photodetection Driven by Ferroelectric Photovoltaic Effect in a Dion-Jacobson Hybrid Perovskite. Adv. Funct. Mater. 2022, 32, 2200223 10.1002/adfm.202200223. [DOI] [Google Scholar]
- Drozdowski D.; Fedoruk K.; Kabanski A.; Maczka M.; Sieradzki A.; Gagor A. Broadband Yellow and White Emission from Large Octahedral Tilting in (110)-Oriented Layered Perovskites: Imidazolium-Methylhydrazinium Lead Halides. J. Mater. Chem. C 2023, 11, 4907–4915. 10.1039/D3TC00401E. [DOI] [Google Scholar]
- Lee J.-W.; Seo S.; Nandi P.; Jung H. S.; Park N.-G.; Shin H. Dynamic Structural Property of Organic-Inorganic Metal Halide Perovskite. iScience 2021, 24, 101959 10.1016/j.isci.2020.101959. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Byranvand M. M.; Otero-Martínez C.; Ye J.; Zuo W.; Manna L.; Saliba M.; Hoye R. L. Z.; Polavarapu L. Recent Progress in Mixed A-Site Cation Halide Perovskite Thin-Films and Nanocrystals for Solar Cells and Light-Emitting Diodes. Adv. Opt. Mater. 2022, 10, 2200423 10.1002/adom.202200423. [DOI] [Google Scholar]
- Mohanty A.; Swain D.; Govinda S.; Row T. N. G.; Sarma D. D. Phase Diagram and Dielectric Properties of MA1–xFAxPbI3. ACS Energy Lett. 2019, 4, 2045–2051. 10.1021/acsenergylett.9b01291. [DOI] [Google Scholar]
- Zheng H.; Dai J.; Duan J.; Chen F.; Zhu G.; Wang F.; Xu C. Temperature-Dependent Photoluminescence Properties of Mixed-Cation Methylammonium-Formamidium Lead Iodide [HC(NH2)2]x[CH3NH3]1–xPbI3 Perovskite Nanostructures. J. Mater. Chem. C 2017, 5, 12057–12061. 10.1039/C7TC04146B. [DOI] [Google Scholar]
- Glinka Y. D.; Cai R.; Gao X.; Wu D.; Chen R.; Sun X. W. Structural Phase Transitions and Photoluminescence Mechanism in a Layer of 3D Hybrid Perovskite Nanocrystals. AIP Adv. 2020, 10, 065028 10.1063/5.0002171. [DOI] [Google Scholar]
- Ahmadi M.; Collins L.; Puretzky A.; Zhang J.; Keum J. K.; Lu W.; Ivanov I.; Kalinin S. V.; Hu B. Exploring Anomalous Polarization Dynamics in Organometallic Halide Perovskites. Adv. Mater. 2018, 30, 1705298 10.1002/adma.201705298. [DOI] [PubMed] [Google Scholar]
- Onoda-Yamamuro N.; Matsuo T.; Suga H. Dielectric Study of CH3NH3PbX3 (X = Cl, Br, I). J. Phys. Chem. Solids 1992, 53, 935–939. 10.1016/0022-3697(92)90121-S. [DOI] [Google Scholar]
- Govinda S.; Kore B. P.; Bokdam M.; Mahale P.; Kumar A.; Pal S.; Bhattacharyya B.; Lahnsteiner J.; Kresse G.; Franchini C.; et al. Behavior of Methylammonium Dipoles in MAPbX3 (X = Br and I). J. Phys. Chem. Lett. 2017, 8, 4113–4121. 10.1021/acs.jpclett.7b01740. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fabini D. H.; Seshadri R.; Kanatzidis M. G. The Underappreciated Lone Pair in Halide Perovskites Underpins their Unusual Properties. MRS Bull. 2020, 45, 467–477. 10.1557/mrs.2020.142. [DOI] [Google Scholar]
- Kalita D.; Sahu P.; Nandi P.; Madapu K. K.; Dhara S.; Schoekel A.; Mahanti S. D.; Topwal D.; Manju U.. Elucidating Phase Transitions and the Genesis of Broadband Emission in MA1–xFAxPbBr3 Perovskites. J. Phys. Chem. C 2023.12724608. 10.1021/acs.jpcc.3c05705 [DOI] [Google Scholar]
- Gallop N. P.; Ye J.; Greetham G. M.; Jansen T. L. C.; Dai L.; Zelewski S. J.; Arul R.; Baumberg J. J.; Hoye R. L. Z.; Bakulin A. A. The Effect of Caesium Alloying on the Ultrafast Structural Dynamics of Hybrid Organic-Inorganic Halide Perovskites. J. Mater. Chem. A 2022, 10, 22408–22418. 10.1039/D2TA05207E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- García-Fernández A.; Bermúdez-García J. M.; Castro-García S.; Llamas-Saiz A. L.; Artiaga R.; López-Beceiro J.; Hu S.; Ren W.; Stroppa A.; Sánchez-Andújar M.; et al. Phase Transition, Dielectric Properties, and Ionic Transport in the [(CH3)2NH2]PbI3 Organic-Inorganic Hybrid with 2H-Hexagonal Perovskite Structure. Inorg. Chem. 2017, 56, 4918–4927. 10.1021/acs.inorgchem.6b03095. [DOI] [PubMed] [Google Scholar]
- García-Fernández A.; Juarez-Perez E. J.; Bermúdez-García J. M.; Llamas-Saiz A. L.; Artiaga R.; López-Beceiro J. J.; Señarís-Rodríguez M. A.; Sánchez-Andújar M.; Castro-García S. Hybrid Lead Halide [(CH3)2NH2]PbX3 (X = Cl– and Br–) Hexagonal Perovskites with Multiple Functional Properties. J. Mater. Chem. C 2019, 7, 10008–10018. 10.1039/C9TC03543E. [DOI] [Google Scholar]
- Im J.-H.; Chung J.; Kim S.-J.; Park N.-G. Synthesis, Structure, and Photovoltaic Property of a Nanocrystalline 2H Perovskite-Type Novel Sensitizer (CH3CH2NH3)PbI3. Nanoscale Res. Lett. 2012, 7, 353. 10.1186/1556-276X-7-353. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lin C.-W.; Liu F.; Chen T.-Y.; Lee K.-H.; Chang C.-K.; He Y.; Leung T. L.; Ng A. M. C.; Hsu C.-H.; Popović J.; et al. Structure-Dependent Photoluminescence in Low-Dimensional Ethylammonium, Propylammonium, and Butylammonium Lead Iodide Perovskites. ACS Appl. Mater. Interfaces 2020, 12, 5008–5016. 10.1021/acsami.9b17881. [DOI] [PubMed] [Google Scholar]
- Jung M.-H. Exploration of Two-Dimensional Perovskites Incorporating Methylammonium for High Performance Solar Cells. CrystEngComm 2021, 23, 1181–1200. 10.1039/D0CE01469A. [DOI] [Google Scholar]
- Peng W.; Miao X.; Adinolfi V.; Alarousu E.; El Tall O.; Emwas A.-H.; Zhao C.; Walters G.; Liu J.; Ouellette O.; et al. Engineering of CH3NH3PbI3 Perovskite Crystals by Alloying Large Organic Cations for Enhanced Thermal Stability and Transport Properties. Angew. Chem., Int. Ed. 2016, 55, 10686–10690. 10.1002/anie.201604880. [DOI] [PubMed] [Google Scholar]
- Wu C.; Chen K.; Guo D. Y.; Wang S. L.; Li P. G. Cations Substitution Tuning Phase Stability in Hybrid Perovskite Single Crystals by Strain Relaxation. RSC Adv. 2018, 8, 2900–2905. 10.1039/C7RA12521F. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y.; Zhang T.; Li G.; Xu F.; Wang T.; Li Y.; Yang Y.; Zhao Y. A Mixed-Cation Lead Iodide MA1–xEAxPbI3 Absorber for Perovskite Solar Cells. J. Energy Chem. 2018, 27, 215–218. 10.1016/j.jechem.2017.09.027. [DOI] [Google Scholar]
- Vega E.; Mollar M.; Marí B. Effect of Guanidinium on the Optical Properties and Structure of the Methylammonium Lead Halide Perovskite. J. Alloys Compd. 2018, 739, 1059–1064. 10.1016/j.jallcom.2017.12.177. [DOI] [Google Scholar]
- Wang W.; Xu Q. Guanidinium Cation Doped (Gua)x(MA)1–xPbI3 Single Crystal for High Performance X-Ray Detector. Nucl. Instrum. Methods Phys. Res., Sect. A 2021, 1000, 165234 10.1016/j.nima.2021.165234. [DOI] [Google Scholar]
- Minussi F. B.; Bertoletti E. M.; Reis S. P.; Carvalho J. F.; Araújo E. B. Guanidinium Substitution-Dependent Phase Transitions, Ionic Conductivity, and Dielectric Properties of MAPbI3. Chem. Commun. 2022, 58, 2212–2215. 10.1039/D1CC06642K. [DOI] [PubMed] [Google Scholar]
- Minussi F. B.; Silva L. A.; Araújo E. B. Structure, Optoelectronic Properties and Thermal Stability of the Triple Organic Cation GAxFAxMA1– 2xPbI3 System Prepared by Mechanochemical Synthesis. Phys. Chem. Chem. Phys. 2022, 24, 4715–4728. 10.1039/D1CP04977A. [DOI] [PubMed] [Google Scholar]
- Minussi F. B.; Silva R. M. Jr.; Araújo E. B. Composition-Property Relations for GAxFAyMA1–x–yPbI3 Perovskites. Small 2023, 20, 2305054 10.1002/smll.202305054. [DOI] [PubMed] [Google Scholar]
- Minussi F. B.; da Silva R. M. J.; Araújo E. B. Differing Effects of Mixed A-Site Composition on the Properties of Hybrid Lead Iodide Perovskites. J. Phys. Chem. C 2023, 127, 8814–8824. 10.1021/acs.jpcc.3c01179. [DOI] [Google Scholar]
- Saliba M.; Matsui T.; Domanski K.; Seo J.-Y.; Ummadisingu A.; Zakeeruddin S. M.; Correa-Baena J.-P.; Tress W. R.; Abate A.; Hagfeldt A.; et al. Incorporation of Rubidium Cations into Perovskite Solar Cells Improves Photovoltaic Performance. Science 2016, 354, 206–209. 10.1126/science.aah5557. [DOI] [PubMed] [Google Scholar]
- Aebli M.; Porenta N.; Aregger N.; Kovalenko M. V. Local Structure of Multinary Hybrid Lead Halide Perovskites Investigated by Nuclear Quadrupole Resonance Spectroscopy. Chem. Mater. 2021, 33, 6965–6973. 10.1021/acs.chemmater.1c01945. [DOI] [Google Scholar]
- Mundt L. E.; Zhang F.; Palmstrom A. F.; Xu J.; Tirawat R.; Kelly L. L.; Stone K. H.; Zhu K.; Berry J. J.; Toney M. F.; et al. Mixing Matters: Nanoscale Heterogeneity and Stability in Metal Halide Perovskite Solar Cells. ACS Energy Lett. 2022, 7, 471–480. 10.1021/acsenergylett.1c02338. [DOI] [Google Scholar]
- Shannon R. D. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Crystallogr., Sect. A 1976, 32, 751–767. 10.1107/S0567739476001551. [DOI] [Google Scholar]
- Noh J. H.; Im S. H.; Heo J. H.; Mandal T. N.; Seok S. I. Chemical Management for Colorful, Efficient, and Stable Inorganic-Organic Hybrid Nanostructured Solar Cells. Nano Lett. 2013, 13, 1764–1769. 10.1021/nl400349b. [DOI] [PubMed] [Google Scholar]
- Niemann R. G.; Kontos A. G.; Palles D.; Kamitsos E. I.; Kaltzoglou A.; Brivio F.; Falaras P.; Cameron P. J. Halogen Effects on Ordering and Bonding of CH3NH3+ in CH3NH3PbX3 (X = Cl, Br, I) Hybrid Perovskites: A Vibrational Spectroscopic Study. J. Phys. Chem. C 2016, 120, 2509–2519. 10.1021/acs.jpcc.5b11256. [DOI] [Google Scholar]
- Shahrokhi S.; Dubajic M.; Dai Z.-Z.; Bhattacharyya S.; Mole R. A.; Rule K. C.; Bhadbhade M.; Tian R.; Mussakhanuly N.; Guan X.; Yin Y.; Nielsen M. P.; Hu L.; Lin C.-H.; Chang S. L. Y.; Wang D.; Kabakova I. V.; Conibeer G.; Bremner S.; Li X.-G.; Cazorla C.; Wu T. Anomalous Structural Evolution and Glassy Lattice in Mixed-Halide Hybrid Perovskites. Small 2022, 18, 2200847 10.1002/smll.202200847. [DOI] [PubMed] [Google Scholar]
- Brivio F.; Caetano C.; Walsh A. Thermodynamic Origin of Photoinstability in the CH3NH3Pb(I1–xBrx)3 Hybrid Halide Perovskite Alloy. J. Phys. Chem. Lett. 2016, 7, 1083–1087. 10.1021/acs.jpclett.6b00226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lehmann F.; Franz A.; Többens D. M.; Levcenco S.; Unold T.; Taubert A.; Schorr S. The Phase Diagram of a Mixed Halide (Br, I) Hybrid Perovskite Obtained by Synchrotron X-Ray Diffraction. RSC Adv. 2019, 9, 11151–11159. 10.1039/C8RA09398A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tang S.; Xiao X.; Hu J.; Gao B.; Chen H.; Peng Z.; Wen J.; Era M.; Zou D. Solvent-Free Mechanochemical Synthesis of a Systematic Series of Pure-Phase Mixed-Halide Perovskites MAPb(IxBr1–x)3 and MAPb(BrxCl1–x)3 for Continuous Composition and Band-Gap Tuning. ChemPlusChem. 2020, 85, 240–246. 10.1002/cplu.201900723. [DOI] [PubMed] [Google Scholar]
- Selig O.; Sadhanala A.; Müller C.; Lovrincic R.; Chen Z.; Rezus Y. L. A.; Frost J. M.; Jansen T. L. C.; Bakulin A. A. Organic Cation Rotation and Immobilization in Pure and Mixed Methylammonium Lead-Halide Perovskites. J. Am. Chem. Soc. 2017, 139, 4068–4074. 10.1021/jacs.6b12239. [DOI] [PubMed] [Google Scholar]
- Askar A. M.; Karmakar A.; Bernard G. M.; Ha M.; Terskikh V. V.; Wiltshire B. D.; Patel S.; Fleet J.; Shankar K.; Michaelis V. K. Composition-Tunable Formamidinium Lead Mixed Halide Perovskites via Solvent-Free Mechanochemical Synthesis: Decoding the Pb Environments Using Solid-State NMR Spectroscopy. J. Phys. Chem. Lett. 2018, 9, 2671–2677. 10.1021/acs.jpclett.8b01084. [DOI] [PubMed] [Google Scholar]
- López C. A.; Alvarez-Galván M. C.; Martínez-Huerta M. V.; Fauth F.; Alonso J. A. Crystal Structure Features of CH3NH3PbI3–xBrx Hybrid Perovskites Prepared by Ball Milling: a Route to More Stable Materials. CrystEngComm 2020, 22, 767–775. 10.1039/C9CE01461F. [DOI] [Google Scholar]
- Colella S.; Mosconi E.; Fedeli P.; Listorti A.; Gazza F.; Orlandi F.; Ferro P.; Besagni T.; Rizzo A.; Calestani G.; et al. MAPbI3–xClx Mixed Halide Perovskite for Hybrid Solar Cells: The Role of Chloride as Dopant on the Transport and Structural Properties. Chem. Mater. 2013, 25, 4613–4618. 10.1021/cm402919x. [DOI] [Google Scholar]
- Unger E. L.; Bowring A. R.; Tassone C. J.; Pool V. L.; Gold-Parker A.; Cheacharoen R.; Stone K. H.; Hoke E. T.; Toney M. F.; McGehee M. D. Chloride in Lead Chloride-Derived Organo-Metal Halides for Perovskite-Absorber Solar Cells. Chem. Mater. 2014, 26, 7158–7165. 10.1021/cm503828b. [DOI] [Google Scholar]
- Pistor P.; Borchert J.; Fränzel W.; Csuk R.; Scheer R. Monitoring the Phase Formation of Coevaporated Lead Halide Perovskite Thin Films by in Situ X-ray Diffraction. J. Phys. Chem. Lett. 2014, 5, 3308–3312. 10.1021/jz5017312. [DOI] [PubMed] [Google Scholar]
- Franz A.; Többens D. M.; Steckhan J.; Schorr S. Determination of the Miscibility Gap in the Solid Solutions Series of Methylammonium Lead Iodide/Chloride. Acta Crystallogr., Sect. B 2018, 74, 445–449. 10.1107/S2052520618010764. [DOI] [PubMed] [Google Scholar]
- Schuck G.; Lehmann F.; Ollivier J.; Mutka H.; Schorr S. Influence of Chloride Substitution on the Rotational Dynamics of Methylammonium in MAPbI3–xClx Perovskites. J. Phys. Chem. C 2019, 123, 11436–11446. 10.1021/acs.jpcc.9b01238. [DOI] [Google Scholar]
- Naqvi F. H.; Ko J.-H.; Kim T. H.; Ahn C. W.; Hwang Y. Disentangled Structural and Vibrational Characteristics of Methylammonium Halide Perovskite MAPbBr3–xClx with x = 0 ∼ 3 Studied by X-Ray Diffraction and Raman Scattering. Curr. Appl. Phys. 2022, 44, 150–157. 10.1016/j.cap.2022.10.006. [DOI] [Google Scholar]
- Lee A. Y.; Park D. Y.; Jeong M. S. Correlational Study of Halogen Tuning Effect in Hybrid Perovskite Single Crystals with Raman Scattering, X-Ray Diffraction, and Absorption Spectroscopy. J. Alloys Compd. 2018, 738, 239–245. 10.1016/j.jallcom.2017.12.149. [DOI] [Google Scholar]
- López C. A.; Álvarez Galván M. C.; Martínez-Huerta M. V.; Fernández-Díaz M. T.; Alonso J. A. Dynamic Disorder Restriction of Methylammonium (MA) Groups in Chloride-Doped MAPbBr3 Hybrid Perovskites: A Neutron Powder Diffraction Study. Chem.—Eur. J. 2019, 25, 4496–4500. 10.1002/chem.201806246. [DOI] [PubMed] [Google Scholar]
- Alvarez-Galván M. C.; Alonso J. A.; López C. A.; López-Linares E.; Contreras C.; Lázaro M. J.; Fauth F.; Martínez-Huerta M. V. Crystal Growth, Structural Phase Transitions, and Optical Gap Evolution of CH3NH3Pb(Br1–xClx)3 Perovskites. Cryst. Growth Des. 2019, 19, 918–924. 10.1021/acs.cgd.8b01463. [DOI] [Google Scholar]
- van de Goor T. W. J.; Liu Y.; Feldmann S.; Bourelle S. A.; Neumann T.; Winkler T.; Kelly N. D.; Liu C.; Jones M. A.; Emge S. P.; et al. Impact of Orientational Glass Formation and Local Strain on Photo-Induced Halide Segregation in Hybrid Metal-Halide Perovskites. J. Phys. Chem. C 2021, 125, 15025–15034. 10.1021/acs.jpcc.1c03169. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Karmakar A.; Askar A. M.; Bernard G. M.; Terskikh V. V.; Ha M.; Patel S.; Shankar K.; Michaelis V. K. Mechanochemical Synthesis of Methylammonium Lead Mixed-Halide Perovskites: Unraveling the Solid-Solution Behavior Using Solid-State NMR. Chem. Mater. 2018, 30, 2309–2321. 10.1021/acs.chemmater.7b05209. [DOI] [Google Scholar]
- Schuck G.; Többens D. M.; Wallacher D.; Grimm N.; Tien T. S.; Schorr S. Temperature-Dependent EXAFS Measurements of the Pb L3-Edge Allow Quantification of the Anharmonicity of the Lead-Halide Bond of Chlorine-Substituted Methylammonium (MA) Lead Triiodide. J. Phys. Chem. C 2022, 126, 5388–5402. 10.1021/acs.jpcc.1c05750. [DOI] [Google Scholar]
- Näsström H.; Becker P.; Márquez J. A.; Shargaieva O.; Mainz R.; Unger E.; Unold T. Dependence of Phase Transitions on Halide Ratio in Inorganic CsPb(BrxI1–x)3 Perovskite Thin Films Obtained from High-Throughput Experimentation. J. Mater. Chem. A 2020, 8, 22626–22631. 10.1039/D0TA08067E. [DOI] [Google Scholar]
- Ma J.; Qin M.; Li Y.; Wu X.; Qin Z.; Wu Y.; Fang G.; Lu X. Unraveling the Impact of Halide Mixing on Crystallization and Phase Evolution in CsPbX3 Perovskite Solar Cells. Matter 2021, 4, 313–327. 10.1016/j.matt.2020.10.023. [DOI] [Google Scholar]
- Yuan L.; Yuan M.; Xu H.; Hou C.; Meng X. Moisture-Stimulated Reversible Thermochromic CsPbI3–xBrx Films: In-Situ Spectroscopic-Resolved Structure and Optical Properties. Appl. Surf. Sci. 2022, 573, 151484 10.1016/j.apsusc.2021.151484. [DOI] [Google Scholar]
- Wang Y.; Guan X.; Chen W.; Yang J.; Hu L.; Yang J.; Li S.; Kalantar-Zadeh K.; Wen X.; Wu T. Illumination-Induced Phase Segregation and Suppressed Solubility Limit in Br-Rich Mixed-Halide Inorganic Perovskites. ACS Appl. Mater. Interfaces 2020, 12, 38376–38385. 10.1021/acsami.0c10363. [DOI] [PubMed] [Google Scholar]
- Breniaux E.; Dufour P.; Guillemet-Fritsch S.; Tenailleau C. Unraveling All-Inorganic CsPbI3 and CsPbI2Br Perovskite Thin Films Formation - Black Phase Stabilization by Cs2PbCl2I2 Addition and Flash-Annealing. Eur. J. Inorg. Chem. 2021, 2021, 3059–3073. 10.1002/ejic.202100304. [DOI] [Google Scholar]
- Jung J.; Yun Y.; Yang S. W.; Oh H. G.; Jeon A.-Y.; Nam Y.; Heo Y.-W.; Chae W.-S.; Lee S. Ternary Diagrams of Phase, Stability, and Optical Properties of Cesium Lead Mixed-Halide Perovskites. Acta Mater. 2023, 246, 118661 10.1016/j.actamat.2022.118661. [DOI] [Google Scholar]
- Chen K.; Deng X.; Goddard R.; Tüysüz H. Pseudomorphic Transformation of Organometal Halide Perovskite Using the Gaseous Hydrogen Halide Reaction. Chem. Mater. 2016, 28, 5530–5537. 10.1021/acs.chemmater.6b02233. [DOI] [Google Scholar]
- Maczka M.; Ptak M.; Gagor A.; Stefanska D.; Sieradzki A. Layered Lead Iodide of [Methylhydrazinium]2PbI4 with a Reduced Band Gap: Thermochromic Luminescence and Switchable Dielectric Properties Triggered by Structural Phase Transitions. Chem. Mater. 2019, 31, 8563–8575. 10.1021/acs.chemmater.9b03775. [DOI] [Google Scholar]
- Greenland C.; Shnier A.; Rajendran S. K.; Smith J. A.; Game O. S.; Wamwangi D.; Turnbull G. A.; Samuel I. D. W.; Billing D. G.; Lidzey D. G. Correlating Phase Behavior with Photophysical Properties in Mixed-Cation Mixed-Halide Perovskite Thin Films. Adv. Energy Mater. 2020, 10, 1901350 10.1002/aenm.201901350. [DOI] [Google Scholar]
- Menéndez-Proupin E.; Grover S.; Montero-Alejo A. L.; Midgley S. D.; Butler K. T.; Grau-Crespo R. Mixed-Anion Mixed-Cation Perovskite (FAPbI3)0.875(MAPbBr3)0.125: an Ab Initio Molecular Dynamics Study. J. Mater. Chem. A 2022, 10, 9592–9603. 10.1039/D1TA10860C. [DOI] [Google Scholar]
- Johnston A.; Walters G.; Saidaminov M. I.; Huang Z.; Bertens K.; Jalarvo N.; Sargent E. H. Bromine Incorporation and Suppressed Cation Rotation in Mixed-Halide Perovskites. ACS Nano 2020, 14, 15107–15118. 10.1021/acsnano.0c05179. [DOI] [PubMed] [Google Scholar]
- Xie L.-Q.; Chen L.; Nan Z.-A.; Lin H.-X.; Wang T.; Zhan D.-P.; Yan J.-W.; Mao B.-W.; Tian Z.-Q. Understanding the Cubic Phase Stabilization and Crystallization Kinetics in Mixed Cations and Halides Perovskite Single Crystals. J. Am. Chem. Soc. 2017, 139, 3320–3323. 10.1021/jacs.6b12432. [DOI] [PubMed] [Google Scholar]
- Barrier J.; Beal R. E.; Gold-Parker A.; Vigil J. A.; Wolf E.; Waquier L.; Weadock N. J.; Zhang Z.; Schelhas L. T.; Nogueira A. F.; et al. Compositional Heterogeneity in CsyFA1–yPb(BrxI1–x)3 Perovskite Films and Its Impact on Phase Behavior. Energy Environ. Sci. 2021, 14, 6394–6405. 10.1039/D1EE01184G. [DOI] [Google Scholar]
- Ye H.-Y.; Liao W.-Q.; Hu C.-L.; Zhang Y.; You Y.-M.; Mao J.-G.; Li P.-F.; Xiong R.-G. Bandgap Engineering of Lead-Halide Perovskite-Type Ferroelectrics. Adv. Mater. 2016, 28, 2579–2586. 10.1002/adma.201505224. [DOI] [PubMed] [Google Scholar]
- Abid H.; Trigui A.; Mlayah A.; Hlil E.; Abid Y. Phase Transition in Organic-Inorganic Perovskite (C9H19NH3)2PbI2Br2 of Long-Chain Alkylammonium. Results Phys. 2012, 2, 71–76. 10.1016/j.rinp.2012.04.003. [DOI] [Google Scholar]
- Wright N. E.; Qin X.; Xu J.; Kelly L. L.; Harvey S. P.; Toney M. F.; Blum V.; Stiff-Roberts A. D. Influence of Annealing and Composition on the Crystal Structure of Mixed-Halide, Ruddlesden-Popper Perovskites. Chem. Mater. 2022, 34, 3109–3122. 10.1021/acs.chemmater.1c04213. [DOI] [Google Scholar]
- Roy C. R.; Zhou Y.; Kohler D. D.; Zhu Z.; Wright J. C.; Jin S. Intrinsic Halide Immiscibility in 2D Mixed-Halide Ruddlesden-Popper Perovskites. ACS Energy Lett. 2022, 7, 3423–3431. 10.1021/acsenergylett.2c01631. [DOI] [Google Scholar]
- Calabrese J.; Jones N. L.; Harlow R. L.; Herron N.; Thorn D. L.; Wang Y. Preparation and Characterization of Layered Lead Halide Compounds. J. Am. Chem. Soc. 1991, 113, 2328–2330. 10.1021/ja00006a076. [DOI] [Google Scholar]
- Vasileiadou E. S.; Jiang X.; Kepenekian M.; Even J.; De Siena M. C.; Klepov V. V.; Friedrich D.; Spanopoulos I.; Tu Q.; Tajuddin I. S.; et al. Thick-Layer Lead Iodide Perovskites with Bifunctional Organic Spacers Allylammonium and Iodopropylammonium Exhibiting Trap-State Emission. J. Am. Chem. Soc. 2022, 144, 6390–6409. 10.1021/jacs.2c00571. [DOI] [PubMed] [Google Scholar]
- Ma Y.; Li W.; Liu Y.; Guo W.; Xu H.; Han S.; Tang L.; Fan Q.; Luo J.; Sun Z. Mixing Cage Cations in 2D Metal-Halide Ferroelectrics Enhances the Ferro-Pyro-Phototronic Effect for Self-Driven Photopyroelectric Detection. Chem. Sci. 2023, 14, 10347–10352. 10.1039/D3SC02946H. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mishra A.; Kubicki D. J.; Boziki A.; Chavan R. D.; Dankl M.; Mladenović M.; Prochowicz D.; Grey C. P.; Rothlisberger U.; Emsley L. Interplay of Kinetic and Thermodynamic Reaction Control Explains Incorporation of Dimethylammonium Iodide into CsPbI3. ACS Energy Lett. 2022, 7, 2745–2752. 10.1021/acsenergylett.2c00877. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang K.; Tang M.-C.; Dang H. X.; Munir R.; Barrit D.; De Bastiani M.; Aydin E.; Smilgies D.-M.; De Wolf S.; Amassian A. Kinetic Stabilization of the Sol-Gel State in Perovskites Enables Facile Processing of High-Efficiency Solar Cells. Adv. Mater. 2019, 31, 1808357 10.1002/adma.201808357. [DOI] [PubMed] [Google Scholar]
- Leblanc A.; Mercier N.; Allain M.; Dittmer J.; Fernandez V.; Pauporté T. Lead- and Iodide-Deficient (CH3NH3)PbI3 (d-MAPI): The Bridge between 2D and 3D Hybrid Perovskites. Angew. Chem., Int. Ed. 2017, 56, 16067–16072. 10.1002/anie.201710021. [DOI] [PubMed] [Google Scholar]
- Li N.; Zhu Z.; Chueh C.-C.; Liu H.; Peng B.; Petrone A.; Li X.; Wang L.; Jen A. K.-Y. Mixed Cation FAxPEA1–xPbI3 with Enhanced Phase and Ambient Stability toward High-Performance Perovskite Solar Cells. Adv. Energy Mater. 2017, 7, 1601307 10.1002/aenm.201601307. [DOI] [Google Scholar]
- Stefanska D.; Ptak M.; Maczka M. Synthesis, Photoluminescence and Vibrational Properties of Aziridinium Lead Halide Perovskites. Molecules 2022, 27, 7949. 10.3390/molecules27227949. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Semenikhin O. A.; Kucheriv O. I.; Sacarescu L.; Shova S.; Gural’skiy I. A. Quantum Dots Assembled from an Aziridinium Based Hybrid Perovskite Displaying Tunable Luminescence. Chem. Commun. 2023, 59, 3566–3569. 10.1039/D2CC06791A. [DOI] [PubMed] [Google Scholar]
- Laurita G.; Fabini D. H.; Stoumpos C. C.; Kanatzidis M. G.; Seshadri R. Chemical Tuning of Dynamic Cation Off-Centering in the Cubic Phases of Hybrid Tin and Lead Halide Perovskites. Chem. Sci. 2017, 8, 5628–5635. 10.1039/C7SC01429E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Marshall A. R.; Sansom H. C.; McCarthy M. M.; Warby J. H.; Ashton O. J.; Wenger B.; Snaith H. Dimethylammonium: An A-Site Cation for Modifying CsPbI3. Sol. RRL 2021, 5, 2000599 10.1002/solr.202000599. [DOI] [Google Scholar]
- Wang Z.; Zeng L.; Zhu T.; Chen H.; Chen B.; Kubicki D. J.; Balvanz A.; Li C.; Maxwell A.; Ugur E. Suppressed Phase Segregation for Triple-Junction Perovskite Solar Cells. Nature 2023, 618, 74–79. 10.1038/s41586-023-06006-7. [DOI] [PubMed] [Google Scholar]
- Duijnstee E. A.; Gallant B. M.; Holzhey P.; Kubicki D. J.; Collavini S.; Sturdza B. K.; Sansom H. C.; Smith J.; Gutmann M. J.; Saha S.; et al. Understanding the Degradation of Methylenediammonium and Its Role in Phase-Stabilizing Formamidinium Lead Triiodide. J. Am. Chem. Soc. 2023, 145, 10275–10284. 10.1021/jacs.3c01531. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Teri G.; Ni H.-F.; Luo Q.-F.; Wang X.-P.; Wang J.-Q.; Fu D.-W.; Guo Q. Tin-Based Organic-Inorganic Metal Halides with a Reversible Phase Transition and Thermochromic Response. Mater. Chem. Front. 2023, 7, 2235–2240. 10.1039/D3QM00105A. [DOI] [Google Scholar]
- Liu X.; Ji C.; Wu Z.; Li L.; Han S.; Wang Y.; Sun Z.; Luo J. [C5H12N]SnCl3: A Tin Halide Organic-Inorganic Hybrid as an Above-Room-Temperature Solid-State Nonlinear Optical Switch. Chem.—Eur. J. 2019, 25, 2610–2615. 10.1002/chem.201805390. [DOI] [PubMed] [Google Scholar]
- Zhang W.; Hong M.; Luo J. Centimeter-Sized Single Crystal of a One-Dimensional Lead-Free Mixed-Cation Perovskite Ferroelectric for Highly Polarization Sensitive Photodetection. J. Am. Chem. Soc. 2021, 143, 16758–16767. 10.1021/jacs.1c08281. [DOI] [PubMed] [Google Scholar]
- Jakubas R.; Rok M.; Mencel K.; Bator G.; Piecha-Bisiorek A. Correlation Between Crystal Structures and Polar (Ferroelectric) Properties of Hybrids of Haloantimonates(III) and Halobismuthates(III). Inorg. Chem. Front. 2020, 7, 2107–2128. 10.1039/D0QI00265H. [DOI] [Google Scholar]
- Zhang H.-Y.; Wei Z.; Li P.-F.; Tang Y.-Y.; Liao W.-Q.; Ye H.-Y.; Cai H.; Xiong R.-G. The Narrowest Band Gap Ever Observed in Molecular Ferroelectrics: Hexane-1,6-diammonium Pentaiodobismuth(III). Angew. Chem., Int. Ed. 2018, 57, 526–530. 10.1002/anie.201709588. [DOI] [PubMed] [Google Scholar]
- Szklarz P.; Gagor A.; Jakubas R.; Zielinski P.; Piecha-Bisiorek A.; Cichos J.; Karbowiak M.; Bator G.; Cizman A. Lead-Free Hybrid Ferroelectric Material Based on Formamidine: [NH2CHNH2]3Bi2I9. J. Mater. Chem. C 2019, 7, 3003–3014. 10.1039/C8TC06458J. [DOI] [Google Scholar]
- Wang B.; Ma D.; Zhao H.; Long L.; Zheng L. Room Temperature Lead-Free Multiaxial Inorganic-Organic Hybrid Ferroelectric. Inorg. Chem. 2019, 58, 13953–13959. 10.1021/acs.inorgchem.9b01793. [DOI] [PubMed] [Google Scholar]
- Wang Y.; Shi C.; Han X.-B. Mixed Bromine-Chlorine Induced Great Dielectric and Second-Order Nonlinear Optical Properties Changes in Phase Transitions Compounds [H2mdap][BiBr5(1– x)Cl5x](x = 0.00 – 1.00). J. Phys. Chem. C 2017, 121, 23039–23044. 10.1021/acs.jpcc.7b05126. [DOI] [Google Scholar]
- Ju D.; Zheng X.; Yin J.; Qiu Z.; Türedi B.; Liu X.; Dang Y.; Cao B.; Mohammed O. F.; Bakr O. M.; et al. Tellurium-Based Double Perovskites A2TeX6 with Tunable Band Gap and Long Carrier Diffusion Length for Optoelectronic Applications. ACS Energy Lett. 2019, 4, 228–234. 10.1021/acsenergylett.8b02113. [DOI] [Google Scholar]
- Wei F.; Deng Z.; Sun S.; Xie F.; Kieslich G.; Evans D. M.; Carpenter M. A.; Bristowe P. D.; Cheetham A. K. The Synthesis, Structure and Electronic Properties of a Lead-Free Hybrid Inorganic-Organic Double Perovskite (MA)2KBiCl6 (MA = methylammonium). Mater. Horiz. 2016, 3, 328–332. 10.1039/C6MH00053C. [DOI] [Google Scholar]
- Liu X.; Xu Z.; Long P.; Yao Y.; Ji C.; Li L.; Sun Z.; Hong M.; Luo J. A Multiaxial Layered Halide Double Perovskite Ferroelectric with Multiple Ferroic Orders. Chem. Mater. 2020, 32, 8965–8970. 10.1021/acs.chemmater.0c02966. [DOI] [Google Scholar]
- He L.; Shi P.-P.; Zhou L.; Liu Z.-B.; Zhang W.; Ye Q. Coexisting Ferroelectric and Ferroelastic Orders in Rare 3D Homochiral Hybrid Bimetal Halides. Chem. Mater. 2021, 33, 6233–6239. 10.1021/acs.chemmater.1c02084. [DOI] [Google Scholar]
- Zhou J.; Xie P.; Wang C.; Bian T.; Chen J.; Liu Y.; Guo Z.; Chen C.; Pan X.; Luo M.; Yin J.; Mao L. Hybrid Double Perovskite Derived Halides Based on Bi and Alkali Metals (K, Rb): Diverse Structures, Tunable Optical Properties and Second Harmonic Generation Responses. Angew. Chem., Int. Ed. 2023, 62, e202307646 10.1002/anie.202307646. [DOI] [PubMed] [Google Scholar]
- Wei F.; Deng Z.; Sun S.; Zhang F.; Evans D. M.; Kieslich G.; Tominaka S.; Carpenter M. A.; Zhang J.; Bristowe P. D.; et al. Synthesis and Properties of a Lead-Free Hybrid Double Perovskite: (CH3NH3)2AgBiBr6. Chem. Mater. 2017, 29, 1089–1094. 10.1021/acs.chemmater.6b03944. [DOI] [Google Scholar]
- Long G.; Sabatini R.; Saidaminov M. I.; Lakhwani G.; Rasmita A.; Liu X.; Sargent E. H.; Gao W. Chiral-Perovskite Optoelectronics. Nat. Rev. Mater. 2020, 5, 423–439. 10.1038/s41578-020-0181-5. [DOI] [Google Scholar]
- Ma J.; Wang H.; Li D. Recent Progress of Chiral Perovskites: Materials, Synthesis, and Properties. Adv. Mater. 2021, 33, 2008785 10.1002/adma.202008785. [DOI] [PubMed] [Google Scholar]
- Dong Y.; Zhang Y.; Li X.; Feng Y.; Zhang H.; Xu J. Chiral Perovskites: Promising Materials toward Next-Generation Optoelectronics. Small 2019, 15, 1902237 10.1002/smll.201970209. [DOI] [PubMed] [Google Scholar]
- Leng K.; Li R.; Lau S. P.; Loh K. P. Ferroelectricity and Rashba Effect in 2D Organic-Inorganic Hybrid Perovskites. Trends Chem. 2021, 3, 716–732. 10.1016/j.trechm.2021.05.003. [DOI] [Google Scholar]
- Huang P.-J.; Taniguchi K.; Miyasaka H. Bulk Photovoltaic Effect in a Pair of Chiral-Polar Layered Perovskite-Type Lead Iodides Altered by Chirality of Organic Cations. J. Am. Chem. Soc. 2019, 141, 14520–14523. 10.1021/jacs.9b06815. [DOI] [PubMed] [Google Scholar]
- Ahn J.; Ma S.; Kim J.-Y.; Kyhm J.; Yang W.; Lim J. A.; Kotov N. A.; Moon J. Chiral 2D Organic Inorganic Hybrid Perovskite with Circular Dichroism Tunable Over Wide Wavelength Range. J. Am. Chem. Soc. 2020, 142, 4206–4212. 10.1021/jacs.9b11453. [DOI] [PubMed] [Google Scholar]
- Kim Y.-H.; Zhai Y.; Lu H.; Pan X.; Xiao C.; Gaulding E. A.; Harvey S. P.; Berry J. J.; Vardeny Z. V.; Luther J. M.; et al. Chiral-Induced Spin Selectivity Enables a Room-Temperature Spin Light-Emitting Diode. Science 2021, 371, 1129–1133. 10.1126/science.abf5291. [DOI] [PubMed] [Google Scholar]
- Huang P.-J.; Taniguchi K.; Miyasaka H. Crucial Contribution of Polarity for the Bulk Photovoltaic Effect in a Series of Noncentrosymmetric Two-Dimensional Organic-Inorganic Hybrid Perovskites. Chem. Mater. 2022, 34, 4428–4436. 10.1021/acs.chemmater.2c00094. [DOI] [Google Scholar]
- Ma J.; Fang C.; Chen C.; Jin L.; Wang J.; Wang S.; Tang J.; Li D. Chiral 2D Perovskites with a High Degree of Circularly Polarized Photoluminescence. ACS Nano 2019, 13, 3659–3665. 10.1021/acsnano.9b00302. [DOI] [PubMed] [Google Scholar]
- Liu S.; Heindl M. W.; Fehn N.; Caicedo-Dávila S.; Eyre L.; Kronawitter S. M.; Zerhoch J.; Bodnar S.; Shcherbakov A.; Stadlbauer A.; et al. Optically Induced Long-Lived Chirality Memory in the Color-Tunable Chiral Lead-Free Semiconductor (R)/(S)-CHEA4Bi2BrxI10–x (x = 0–10). J. Am. Chem. Soc. 2022, 144, 14079–14089. 10.1021/jacs.2c01994. [DOI] [PubMed] [Google Scholar]
- Yao B.; Wei Q.; Yang Y.; Zhou W.; Jiang X.; Wang H.; Ma M.; Yu D.; Yang Y.; Ning Z. Symmetry-Broken 2D Lead-Tin Mixed Chiral Perovskite for High Asymmetry Factor Circularly Polarized Light Detection. Nano Lett. 2023, 23, 1938–1945. 10.1021/acs.nanolett.2c05085. [DOI] [PubMed] [Google Scholar]
- Ahart M.; Somayazulu M.; Cohen R. E.; Ganesh P.; Dera P.; Mao H.-k.; Hemley R. J.; Ren Y.; Liermann P.; Wu Z. Origin of Morphotropic Phase Boundaries in Ferroelectrics. Nature 2008, 451, 545–548. 10.1038/nature06459. [DOI] [PubMed] [Google Scholar]























