Download PDF
Review  |  Open Access  |  8 Aug 2024

On the thermal expansion of the tetragonal phase of MAPbI3 and MAPbBr3

Views: 281 |  Downloads: 45 |  Cited:  0
Microstructures 2024;4:2024047.
10.20517/microstructures.2024.33 |  © The Author(s) 2024.
Author Information
Article Notes
Cite This Article

Abstract

Based on previously published research, the structural response of the tetragonal hybrid perovskite crystal structure of MAPbX3 [MA: [CH3NH3]+, methylammonium; X = I, Br] to thermal expansion is reviewed here. From an averaged crystal structure perspective, the tetragonal perovskite structure of MAPbI3 and MAPbBr3, based on diffraction data, shows apparent Pb-X bond length shortening and apparent shrinkage of the [PbX6] octahedra with increasing temperature. At the same time, these apparent observations, and hence the thermal expansion, are related to the progressive phase transformation towards the cubic structure, as the lattice parameters respond to a shear stress that couples to the order parameters, and this coupling is predicted by group theory and thus aims to explain precisely the apparent negative thermal expansion-like effects. A different picture emerges for the thermal expansion when considering the very localized structure, since neither a shortening of the Pb-X bond lengths nor a shrinking of the [PbX6] octahedra is observed with pair distribution function analysis, and the presence of orthorhombic short-range order in the tetragonal and cubic perovskite structures is assumed in published studies. The compared extended X-ray absorption fine structure studies, which also map the local structure and provide the “true” bond distance, show no lead-halide bond length shortening with temperature. The perpendicular mean square relative displacement has been determined. Therefore, a comparison of the tension and bond expansion effects in both perovskites can be made. In the orthorhombic phase of MAPbI3 and MAPbBr3, positive expansion and negative tension of the lead-halide bond are almost balanced. After transitioning to the tetragonal phase, the equilibrium shifts toward negative tension. This suggests that both hybrid perovskites have tighter lead-halide bonds and less rigid [PbX6] octahedra in the tetragonal phase than in the low temperature perovskite crystal structure.

Keywords

Hybrid perovskites, thermal expansion, local structure, pair distribution function (PDF) analysis, extended X-ray absorption fine structure (EXAFS), tension effect

INTRODUCTION

Due to their optoelectronic properties (high optical absorption coefficients, narrow-band bright photoluminescence, tuneable bandgaps, low exciton binding energies, and long-range carrier diffusion)[1-3], hybrid perovskites are promising materials for applications in photovoltaics[4,5]. Hybrid perovskites of the general form MAPbX3 (MA: [CH3NH3]+, methylammonium; X = I, Br) such as MAPbI3 and MAPbBr3, undergo phase transformations from the cubic high-temperature aristo-type structure (space group Pm$$\overline{3}$$m) first to a tetragonal structure (space group I4/mcm, Tc, MAPbI3 = 330 K and Tc, MAPbBr3 = 236 K) and finally to an orthorhombic low-temperature structure (space group Pnma, Tc, MAPbI3 = 160 K and Tc, MAPbBr3 = 149 K) in which the MA molecule is less disordered[6,7]. In the cubic and tetragonal phase, only disordered positions for the MA molecule can be observed with diffraction methods due to the fast jump-rotation of the MA molecule (ps relaxation times and THz relaxation frequencies)[8]. In the low-temperature phase, the MA molecule also jump-rotates, but much slower (ns relaxation times and GHz relaxation frequencies)[8]. Due to the interaction between the strong MA molecule dynamics and the rigidity of the [PbX6] octahedra framework structure, it is not surprising that liquid-crystal duality is increasingly employed in the description of hybrid perovskites[9,10]. In a new classification, the existing dynamic aspects are added to the crystal structure description[11]. Thus, the tetragonal I4/mcm room-temperature structure might now be described with the extended Glazer notation adadb- and the cubic Pm$$\overline{3}$$m aristo-type structure with adadad (in this new notation, d corresponds to the dynamic tilt, which indicates that the corresponding octahedra oscillate between two positions (positive and negative amplitude)). In this context, the quantitative analysis of the structural dynamics in time and space from molecular dynamics (MD) simulations of perovskite crystals by Liang et al. is noteworthy[12]. Recent experimental work by Weadock et al. also attempts to describe the nature of the dynamic local ordering of MAPbI3 and MAPbBr3 in the cubic phase by a combination of single-crystal diffuse scattering, neutron inelastic spectroscopy, and MD simulations[13]. In this investigation, a remarkable collective dynamics consisting of a network of local, two-dimensional, circular regions of dynamically tilting lead halide octahedra (lower symmetry), giving rise to long-range intermolecular CH3NH3+ correlations, has been identified[13]. The presence of orthorhombic short-range order in the tetragonal and cubic structures of MAPbX3 has also been experimentally confirmed by studies of the pair distribution function (PDF)[14-17]. As stated by Simenas et al. in their comprehensive review of hybrid perovskites[18], referring to the work of Weadock et al.[13], “the presence of such low symmetry correlated octahedral distortions at the nanoscale could be related to the observed ferroelectric effects in this compound”, but we will not discuss the ferroic properties in our selective review and refer to existing experimental work[19-21] and review publications[22-25] on this controversial topic. Due to the temperature-dependent changes in the crystal structures, local ordering phenomena and structural dynamics, hybrid perovskites exhibit a number of remarkable thermal properties that are particularly important for their application in photovoltaics[26]. For example, the thermal expansion of the tetragonal room temperature structure of MAPbI3 is relatively high and strongly anisotropic. In the temperature range from 308 to 324 K, Jacobsson et al. observed a negative linear thermal expansion coefficient α[001], tetragonal = -106 × 10-6 K-1 in the [001] direction and a positive linear thermal expansion coefficient α[100], tetragonal = 132 × 10-6 K-1 in the [100] direction[27]. Thus, the thermal expansion corresponds more to a soft material, for instance, a polymer such as polypropylene (α = 118 × 10-6 K-1)[28], than, for example, typical representatives of semiconductor technology such as silicon (α = 2.6 × 10-6 K-1)[29]. In keeping with soft materials, the thermal conductivity of hybrid perovskites is very low, with Ge et al. observing an extremely low thermal conductivity of κ = 0.3 Wm-1K-1 for MAPbI3[30]. The extreme thermal properties of hybrid perovskites cause difficulties because photovoltaics require close coupling with many different materials that have completely different thermal properties than hybrid perovskites. This affects both manufacturing and application stability and longevity of the thin film photovoltaic devices, making strain engineering important[31]. The observed thermal properties of hybrid perovskites indicate strong anharmonic interactions, with increasing evidence that the anharmonic properties of the lead-halide bond in particular[32] play a prominent role when wanting to better understand the extremely soft hybrid perovskite lattice. A very well-suited method to investigate the anharmonic properties of the lead-halide bond is extended X-ray absorption fine structure (EXAFS) analysis in combination with crystallographic structure determination[33,34]. Since EXAFS can be used to determine the relative atomic vibrations, including the correlation of atomic motion, in the form of the parallel and perpendicular mean square relative displacement (MSRD), the bond-bending and bond-stretching effective force constants and frequencies of the Pb-X bonds are directly accessible[33,34]. The perpendicular MSRD can be determined by comparing the “true” distance of certain atomic pairs, in this case, lead-halide, obtained by fitting the EXAFS spectra with a simulation, with the “apparent” distance measured by diffractometric means[34].

In a selective short review, the thermal expansion of the tetragonal phase of MAPbI3 and MAPbBr3 is discussed here on the basis of published work in which results from X-ray diffraction (XRD) and EXAFS are combined[6,7]. Based on the results of the structural investigations on the temperature-dependent XRD data of MAPbI3 and MAPbBr3, the thermal expansion in the tetragonal phases is analyzed and described by linear thermal expansion coefficients. The temperature dependence of these thermal expansion coefficients is then quantified for MAPbI3. For both components, the phase transformation from the tetragonal to the cubic perovskite structure is described by the corresponding order and strain parameters. Based on the PDF analysis of MAPbBr3 by Bernasconi et al., the thermal expansion is discussed from a local order perspective[16]. Finally, the results of temperature-dependent EXAFS analyses of MAPbI3 and MAPbBr3 are compared and discussed.

DISCUSSION

Apparent shrinkage of [PbX6] octahedra and Pb-X bond lengths changes in the tetragonal phases of MAPbI3 and MAPbBr3

Two previous temperature-dependent crystallographic investigations of MAPbI3 from Schuck et al.[6] and of MAPbBr3 from Weadock et al.[7] are compared and discussed in relation to the observed volume changes of the [PbX6] octahedra and the changes in Pb-X bond lengths. In the following, we will focus on the tetragonal phase [Figure 1] and the transition to the cubic structure by looking at the behavior of the [PbX6] octahedra network in dependence on temperature [Figure 2]. Powder samples of MAPbI3 were measured by Schuck et al. using synchrotron XRD as a function of temperature, and the resulting diffractograms were refined using the Rietveld method. The MA molecule geometry was taken from the previous density functional theory (DFT) optimization and treated as rigid[35]. The results of the structure refinements of MAPbI3[6] agree with previous temperature-dependent studies[36-38].

On the thermal expansion of the tetragonal phase of MAPbI<sub>3</sub> and MAPbBr<sub>3</sub>

Figure 1. Detail of the tetragonal room-temperature structure of MAPbI3 with its large iodine anisotropic temperature factors (ADPs) (their flat shape indicates a transverse displacement perpendicular to the Pb-I-Pb bonds and in the direction of the MA molecule), the coordination polyhedra of the MA molecule [MAI12], a cuboctahedron (light yellow), and one [PbI6] octahedra (light gray)[37]. The anisotropic shape and orientation of the halide ADPs are partly due to the strongly anharmonic metal-halide bond (see EXAFS section) and not only to the hydrogen bond between the MA and the halide atoms (formed between the jump rotations). In this representation of a single MA molecule, neither the 4-fold rotational axis nor the 2-fold rotational axis of the I4/mcm space group symmetry seems to exist for the molecule. However, since the MA molecule jump rotates[8] and there are, on average, eight MA molecule orientations in the cage, an average structure with I4/mcm symmetry is observed in diffraction.

On the thermal expansion of the tetragonal phase of MAPbI<sub>3</sub> and MAPbBr<sub>3</sub>

Figure 2. (A) The [PbI6] octahedra network of MAPbI3 for the cubic high-temperature structure at 350 K[36] (top) and the tetragonal room-temperature structure at 300 K[37] (bottom). (B) Pb-Xa-Pb bond angle (octahedra tilt) of MAPbI3 (red) and MAPbBr3 (blue) in the tetragonal (solid symbols) and cubic phase (open symbols). Besides the data from Schuck et al.[6] and Weadock et al.[7], values for the cubic phase at 350 K of MAPbI3 from Whitfield et al. are also given[36]. Adapted with permission from Schuck et al.[6]. Copyright 2022 American Chemical Society.

The volume of MAPbI3 increases with temperature over the whole temperature range (volumetric thermal expansion coefficient αV, orthorhombic = 90 × 10-6 K-1 and αV, tetragonal = 119 × 10-6 K-1). Based on the lattice constants for a pseudo-cubic cell from Schuck et al., a negative linear thermal expansion coefficient of α[001], tetragonal = -23.3(3) × 10-6 K-1 in the [001] direction and a positive linear thermal expansion coefficient of α[100], tetragonal = 75(3) × 10-6 K-1 in the [100] direction can be determined for MAPbI3 (temperature range 220-260 K) [Figure 3A][6]. The differences between these linear thermal expansion coefficients of Schuck et al. and those of Jacobsson et al. are due to the different temperature ranges used to determine the gradient[6,27]. Jacobsson et al. used a higher temperature range of 308 to 324 K. Figure 4 shows the percentage changes in Pb-X bond length expansion [Figure 4A] and the percentage changes in volume expansion [Figure 4B] from the onset of the tetragonal phase to higher temperatures[27]. In the tetragonal phase, the maximum thermal expansion at 300 K is +1.68% (compared to 160 K) for the whole volume and +2.14% for the [MAI12] cuboctahedra. In contrast, the negative thermal expansion (NTE) of the [PbI6] octahedra is -0.5%, corresponding to the negative change in average bond length, which is -0.17% smaller at 300 K compared to the value at 160 K. Based on the tetragonal lattice constants for a pseudo-cubic cell from Weadock et al., which are available from their temperature-dependent single-crystal structure investigations [crystallographic information files (CIF) were made available][7], a NTE of α[001], tetragonal = -66(3) × 10-6 K-1 in the [001] direction and a positive thermal expansion of α[100], tetragonal = 95(2) × 10-6 K-1 in the [100] direction can be determined for MAPbBr3 (for the temperature range 200-230 K) [Figure 3B]. As shown in Figure 4, a maximum thermal expansion at 240 K is observed for MAPbBr3 in the tetragonal phase, analogous to the behavior of MAPbI3. While the total volume of MAPbBr3 and the [MABr12] cuboctahedra volume increase by +0.78% and +1.04%, respectively, the Pb-Br bond length decreases by -0.19% with a concomitant NTE of -0.53% for the [PbBr6] octahedra. In both MAPbI3 and MAPbBr3, the temperature-dependent negative changes in the partial volumes of the [PbX6] octahedra, on the one hand, and positive partial volume changes of the [MAX12] cuboctahedra, on the other hand, differ significantly from the change in the total volume in the tetragonal phase. While the [MAX12] cuboctahedra volume expands almost linearly with increasing temperature (greater expansion than total volume), the [PbX6] octahedra volume shrinks towards the cubic phase [Figure 4B]. A comparison with pressure-dependent data in the tetragonal phase would certainly also be interesting, but for MAPbBr3 such data are not yet available[39] in the corresponding temperature and pressure range, and for MAPbI3 a clearly different behavior of the tetragonal structure already occurs at 300 K in the range of 0.05 GPa (the [PbI6] octahedra become 0.15% smaller, and Pb-I shortens by 0.05%)[40], so that it cannot be excluded that phase transformations also occur at minimal pressures and a direct comparison of the mechanochemical sensitive[41] MAPbI3 data is not readily possible.

On the thermal expansion of the tetragonal phase of MAPbI<sub>3</sub> and MAPbBr<sub>3</sub>

Figure 3. Temperature-dependent lattice constants for the tetragonal (pseudo-cubic) and cubic unit cell for (A) MAPbI3 (cubic at 350 K from Whitfield et al.[36]) and (B) MAPbBr3 based on the results of Schuck et al.[6] and Weadock et al.[7]. In both cases, the linear thermal expansion coefficients in direction [100] [positive thermal expansion (PTE)] and in direction [001] [negative thermal expansion (NTE)] are also given. Adapted with permission from Schuck et al.[6]. Copyright 2022 American Chemical Society.

On the thermal expansion of the tetragonal phase of MAPbI<sub>3</sub> and MAPbBr<sub>3</sub>

Figure 4. (A) Averaged Pb-X bond lengths expansion and (B) volumetric expansion of MAPbI3 (red) and MAPbBr3 (blue) in the tetragonal (solid symbols) and cubic phase (open and half-open symbols), the phase transformation temperature is indicated in each case (tetragonal to cubic). Besides the data from Schuck et al.[6] and Weadock et al.[7], values for the cubic phase of MAPbI3 are also given from Whitfield et al.[36] (left half-open symbols) and Stoumpos et al.[38] (right half-open symbols). (B) In addition to the volume changes for the [PbX6] octahedra (squares), the volume changes for the total volume (diamond symbols) and the volume changes for the [MAX12] cuboctahedra (circles) are also given. Note: Both Pb-X bond lengths, i.e., the one along tetragonal c (Pb-Xb in Figure 2A) and the one in the tetragonal a-b plane (Pb-Xa in Figure 2A), become smaller in the tetragonal phase with increasing temperature. Both bond lengths (Pb-Xa and Pb-Xb) for MAPbI3 and MAPbBr3 are within the error bars of the averaged value Pb-X. Adapted with permission from Schuck et al.[6]. Copyright 2022 American Chemical Society.

Temperature-dependent linear thermal expansion coefficients and the Grüneisen tensor

The linear thermal expansion coefficients calculated for MAPbI3 from the data of Schuck et al. differ significantly from the values of Jacobsson et al.[6,27]. This is not surprising since one linear fit obviously cannot describe the entire temperature range of the tetragonal phase. For this reason, the temperature dependence of the linear thermal expansion coefficients is investigated using the tetragonal lattice constants of Whitfield et al. [the lattice constants from temperature-dependent synchrotron powder XRD measurements of Whitfield et al. (figure 6C in Ref.[36]) have been digitized by us and are shown here in Figure 5A][36]. The temperature dependence of the linear thermal expansion coefficients [Figure 5B] can be approximately described for the PTE α[100] with a linear behavior; an increase towards the cubic phase up to 97 × 10-6 K-1 is obtained. The NTE α[001] can be described with a third-order polynomial and reaches a maximum value of -88 × 10-6 K-1 towards the cubic phase. Due to the temperature dependence of the two linear thermal expansion coefficients α[001] and α[100], the linear thermal expansion coefficient α[averaged] determined from the averaged tetragonal lattice constant aaveraged = ((2*atet)+ctet)/3 is more suitable for describing the thermal expansion as it has almost no temperature dependence. The value given by Whitfiled et al.[36] is to be multiplied by 1/L0 and then results in α[averaged] = 42.4(4) × 10-6 K-1 (dotted line in Figure 5B)[42]. It is common to describe thermal expansion properties of solids in Grüneisen parameters: NTE results in negative Grüneisen parameter γ.

On the thermal expansion of the tetragonal phase of MAPbI<sub>3</sub> and MAPbBr<sub>3</sub>

Figure 5. (A) Temperature-dependent tetragonal (pseudo-cubic) lattice constants according to Whitfield et al.[36] (black) and from Schuck et al.[6] (red). (B) Temperature dependence of the linear thermal expansion coefficients α[100], α[001], and α[averaged] based on the lattice constants of Whitfield et al. [displayed in (A)][36]. α[100]: linear behavior; α[001]: third-order polynomial (solid lines). Adapted with permission from Schuck et al.[6]. Copyright 2022 American Chemical Society.

On the thermal expansion of the tetragonal phase of MAPbI<sub>3</sub> and MAPbBr<sub>3</sub>

Figure 6. (A) Temperature-dependent values of $${\left[2 c_{13}^{S} \alpha_{[100]}+c_{33}^{S} \alpha_{[001]}\right] \sim \gamma_{[001]}}$$ and (B) temperature-dependent values of $${\left[\left(c_{11}^{S}+c_{13}^{S}\right) \alpha_{[100]}+c_{13}^{S} \alpha_{[001]}\right] \sim \gamma_{[100]}}$$. (A and B) for MAPbI3 using the tabulated cij values from Feng et al.[45].

In the case of the tetragonal crystal structure of hybrid perovskites, two independent Grüneisen parameters, γ[100] and γ[001], are related to the principal linear thermal expansion coefficients α[100] and α[001] via (according to Schorr et al.[43]):

$$ \begin{equation} \begin{aligned} \begin{array}{l} \gamma_{[100]}=\frac{V_{m}}{C_{p}}\left[\left(c_{11}^{S}+c_{13}^{S}\right) \alpha_{[100]}+c_{13}^{S} \alpha_{[001]}\right] \\ \gamma_{[001]}=\frac{V_{m}}{C_{p}}\left[2 c_{13}^{S} \alpha_{[100]}+c_{33}^{S} \alpha_{[001]}\right] . \\ \end{array} \end{aligned} \end{equation} $$

where Vm is the molar volume, Cp is the molar specific heat at constant pressure, and the cij are the adiabatic elastic stiffness coefficients in this Grüneisen tensor (according to Haussühl and Kitaigorodskii)[44]. Using the tabulated cij values[45], a negative Grüneisen parameter γ[001] can only be expected near the phase transition to the cubic phase above 307 K [Figure 6A]. At lower temperatures, γ[001] is positive and the Grüneisen parameter γ[100] is always positive for MAPbI3 [Figure 6B].

Tetragonal to the cubic phase transition

According to Whitfield et al., the phase transition from the tetragonal to the cubic crystal structure in MAPbI3 and MAPbBr3 can be described by the temperature-dependent change of three degrees of freedom in the tetragonal structure: the rotational distortion caused by the change of the Pb-Ia-Pb angle (R4+), the strain caused by the thermal expansion (GM1+), and the strain caused by the tetragonal distortion (GM3+)[36].

The temperature-dependent changes of the three degrees of freedom were analyzed for MAPbI3 and MAPbBr3 using ISODISTORT (an internet-server tool for exploring structural phase transitions)[46] and are shown in Figure 7. The thermal expansion in the tetragonal phase discussed in the previous two sections is related to the progressive phase transformation towards the cubic structure, since the lattice parameters respond to a shear stress coupled to the order parameters (analyzed with ISODISTORT), and this coupling is predicted by a standard Landau theory treatment. The amplitude of the R4+ rotational mode, which is directly proportional to the cubic-tetragonal order parameter, can be fitted by the power law R4+~(Tc-T)β, where β is the critical exponent [Figure 7]. Both Whitfield et al.[36] and Liu et al.[47] conclude in their analysis that the tetragonal/cubic phase transition for MAPbI3 can be characterized by a single order parameter and that its behavior as a function of temperature and spontaneous strain follows the form of a weakly first-order, nearly tricritical phase transition, which is consistent with the description of a standard Landau free energy function. The standard Landau theory treatment should then also explain the temperature-dependent behavior of the Pb-X bond lengths and the [PbX6] octahedra in the tetragonal phase towards the cubic structure, although it is worth noting that there is some variance in the description of the order parameter in the case of MAPbI3 [Figure 7]. However, the differences between the R4+ data of Schuck et al.[6] and Weadock et al.[7] seem to be smaller.

On the thermal expansion of the tetragonal phase of MAPbI<sub>3</sub> and MAPbBr<sub>3</sub>

Figure 7. ISODISTORT analysis[46]. The phase transition from tetragonal to cubic, based on the data from Schuck et al.[6] and Liu et al.[47] for MAPbI3 and based on the data from Weadock et al.[7] for MAPbBr3 can be described by the temperature-dependent changes of the three degrees of freedom in the tetragonal structure: rotational distortion mode R4+ (blue color, legend on the right) and the two strain mode amplitudes (black color, legend on the left) GM1+ (thermal expansion) and GM3+ (tetragonal distortion (c-a)/c). Fit of the order parameter R4+ for MAPbI3 based on the data from Schuck et al.[6] (blue dashed line) and fit for the data from Liu et al.[47] (blue dotted line), additional also the R4+ fit from Whitfield et al.[36] (figure 8A in Ref.[36]) is reproduced (Tc = 329.95 K and β = 0.26, solid blue line). Fit of the order parameter R4+ for MAPbBr3 based on the data from Weadock et al. (blue dash-dotted line)[7].

Local structure of MAPbI3 and MAPbBr3 with atomic pair distribution function (PDF) analysis

The local atomic structure can be investigated by analyzing the total scattering from diffraction experiments[48,49], which, in contrast to XRD or neutron powder diffraction methods, use much shorter X-ray or neutron radiation wavelengths. In addition, total scattering measurements are generally made over a much larger scattering vector range (Qmax = 30 Å-1) than in conventional XRD experiments (Qmax = 5 Å-1). Besides Bragg scattering, total scattering includes diffusion scattering. The PDF can be used to analyze the total scattering, whereby a Fourier transformation of the measured total scattering must be carried out after extensive data reduction (including a thorough background correction)[48,49].

In the temperature-dependent pair correlation functions G(r) of MAPbI3, it is, first of all, noticeable, as already noted by Whitfield et al., that in a range up to about 4 Å, no discontinuous changes occur in the range from 10 to 350 K, but only a gradual broadening of G(r) towards high temperatures (see figure 4 in Ref.[36]), which is also especially the case in the range of the phase transformation from the orthorhombic to the tetragonal structure (o/t)[36]. This range up to 4 Å, which does not change discontinuously, corresponds to the Pb-X bond distances. The situation is different for the range from 4 Å to larger distances, where a discontinuous change in G(r) in the temperature range of the o/t phase transformation is observed.

A very similar behavior can also be observed in the data available for MAPbBr3. Here, there is also no discontinuous change at small distances in the region of the o/t phase transformation (see figure 1 in Ref.[17]). In the PDF analyses available in the literature, it was observed for both MAPbI3 and MAPbBr3 that in the temperature range of the tetragonal and cubic phases, the low range up to 10 Å, can be best described by a short-range octahedra distortion, which essentially corresponds to the orthorhombic low temperature perovskite crystals structure for both MAPbI3[14,17] and MAPbBr3[15-17]. To illustrate and qualitatively reproduce this effect, PDF fits in the range of 1-10 Å are shown in Figure 8, using the MAPbBr3 data at 200 K from Bird et al.[17] and the refinement results from Bernasconi et al.[16] (here, only the lead and bromine atoms were considered). In the PDF refinement only, the scaling and Uiso were refined. Similar to Bernasconi et al., the use of the tetragonal I4/mcm perovskite structure (Figure 8A, Rw = 0.3449) leads to a poorer refinement than the use of the orthorhombic Pnma perovskite structure (Figure 8B, Rw = 0.1740)[16]. Although the results of the PDF analysis in Figure 8 show deviations (in the regions of r = 7 Å and r = 9.5 Å, there are peaks in the simulation that are not present in the data, and the peak at r = 4 Å is significantly off in thickness), the same qualitative results can be obtained with the experimental data of Bird et al., and thus, the observations of Bernasconi et al. can be confirmed qualitatively[16,17]. The temperature-dependent lattice parameters, Pb-Br bond lengths and [PbBr6] volumes determined by Bernasconi et al. using the orthorhombic perovskite structure in their PDF refinement for the range of 1-10 Å are shown in Figure 9[16]. Obviously, the thermal expansion of the local structure of MAPbBr3 is fundamentally different from the thermal expansion of the averaged structure resulting from the analysis of the XRD data. The PDF refinement results from Bernasconi et al. show that all three orthorhombic lattice parameters increase with temperature, as do the Pb-Br bond lengths and [PbBr6] volumes[16]. More recent structural PDF studies using symmetry-adapted PDF analysis (SAPA)[50] have investigated the local structure of ScF3 and hybrid perovskites utilizing extended degrees of freedom (such as the “scissoring” mode X5+)[17,51].

On the thermal expansion of the tetragonal phase of MAPbI<sub>3</sub> and MAPbBr<sub>3</sub>

Figure 8. Short-range fit (r range = 1-10 Å) by using (A) the tetragonal MAPbBr3 perovskite structure (I4/mcm) and (B) the orthorhombic MAPbBr3 perovskite structure (Pnma). For the PDF fit (with PDFgui 1.0), the data from Bird et al.[17] at 200 K was used together with the fixed refinement results at 200 K from Bernasconi et al.[16]. In the PDF fit, only the scale and Uiso were refined.

On the thermal expansion of the tetragonal phase of MAPbI<sub>3</sub> and MAPbBr<sub>3</sub>

Figure 9. Results of the PDF analysis by Bernasconi et al.[16] using the G(r) range 1-10 Å with orthorhombic Pnma symmetry. (A) orthorhombic perovskite structure at 200 K; (B) orthorhombic lattice constants (pseudo-cubic); (C) averaged Pb-Br bond lengths expansion and (D) volumetric expansion of MAPbBr3. Adapted with permission from Bernasconi et al.[16]. Copyright 2017 American Chemical Society.

Analysis of the anharmonic properties of the lead-halide bond with EXAFS

EXAFS spectroscopy is based on the ionization of an atom when an X-ray quantum is absorbed. This releases an electron whose kinetic energy depends on the energy of the X-rays. The released electron propagates as a matter wave and is scattered by neighboring atoms. Depending on the wavelength of the electron, constructive or destructive interference occurs between the outgoing wave and the backscattered waves. Because the energy of the X-rays is varied, the energy of the released electrons also changes and thus the corresponding wavelength of the electrons. This leads to alternating constructive and destructive interference and, thus, to a change in X-ray absorption depending on the energy, which corresponds to the fine structure of the X-ray absorption spectrum. These X-ray absorption changes can be measured in the energy range from just above the X-ray absorption edge up to a few hundred electron volts above it, i.e., in the upper range of the X-ray absorption edge (hence the term “extended” in EXAFS). From the shape and strength of the X-ray absorption changes, it is possible to deduce at what distance from the ionized atom it is scattered and to what extent, i.e., a so-called radial distribution function is obtained via Fourier transformation of the X-ray absorption “fine structure” data after normalization and background subtraction (for example with the Athena software)[52]. By comparing the scattering of electrons on and between neighboring atoms with simulations (for example, with the Artemis software)[6,52,53], the atomic distances of atom pairs REXAFS can be determined with high accuracy. However, the radial distribution function not only contains information about the atomic distances but also about the distance variations of the respective atom pairs, which corresponds to the parallel MSRD (EXAFS Debye-Waller factor or C2). Since the lifetime of the EXAFS photoelectron (~10-15 s) is very short compared to the atomic vibrations (~10-12 s), and, at the same time, the limited mean free path of the EXAFS photoelectron is restricted to the local environment of the investigated atom, the anharmonic properties of the investigated bond pair can be studied very well[34]. The relative ΔC2 (T) was fitted with an Einstein behavior by Schuck et al.[6] to determine parallel MSRD (T)[54] of MAPbI3:

$$ \begin{equation} \begin{aligned} \operatorname{parallel} \operatorname{MSRD}(\mathrm{T})=\frac{\hbar^{2}}{\mu k_{B}} \frac{1}{2 \theta_{E}} \operatorname{coth}\left(\frac{\theta_{E}}{2 T}\right) \end{aligned} \end{equation} $$

(µ: reduced mass, ΘE: Einstein temperature). The resulting parallel MSRD (T) is shown in Figure 10A. This figure also shows the parallel MSRD (T) for MAPbBr3, which was interpreted by Weadock et al. using both a correlated Debye model and an Einstein behavior [in the Debye model, a distinction was made between a low-temperature (ΘD,LT = 148 K) and a high-temperature range (ΘD,LT = 167 K)][7]. Weadock et al. noted that the Einstein temperatures for MAPbI3 and MAPbBr3 increase with decreasing reduced mass μ = m1m2/(m1 + m2), which corresponds to the relationship between the frequency and mass of an oscillator[7]. From ΘE, a direct derivation of the effective force constant k0 (bond-stretching effective force constant) could be made since the following applies:

$$ \begin{equation} \begin{aligned} k_{0}=\Theta_{E}^{2} \mu=\left(\mu * 1.660539 * 10^{-27} \mathrm{~kg} *\left(2 \pi * v_{E}\right)^{2}\right) / 16.022\left[\mathrm{eV}/\mathrm{Å}^{2}\right] \end{aligned} \end{equation} $$

On the thermal expansion of the tetragonal phase of MAPbI<sub>3</sub> and MAPbBr<sub>3</sub>

Figure 10. Temperature-dependent MSRDs: (A) parallel, (B) perpendicular and (C) γEXAFS. Red: MAPbI3 (Schuck et al.[6]); blue: MAPbBr3 (Weadock et al.[7]). Solid lines: Einstein fits. In (B) the parallel MSRD values for MAPbI3 were fitted with two Einstein temperatures, one for the orthorhombic low-temperature range up to 160 K, the other for the tetragonal high-temperature range up to 260 K (here with an additional static component of 0.05). This procedure also leads to a better explanation of the experimental values for the description of the γEXAFS dependence in (C). Adapted with permission from Schuck et al.[6]. Copyright 2022 American Chemical Society.

[1 N/m = 1/16.022 eV/Å2], [1 THz = 47.9924 K], and νE: Einstein frequency [THz] [Table 1].

Table 1

Einstein temperatures and frequencies for parallel (ΘE||, νE||) and perpendicular MSRD ($$\boldsymbol{\Theta}_{\mathbf{E} \perp}$$, $$\boldsymbol{V}_{\mathbf{E} \perp}$$), bond-stretching effective force constant k|| (corresponds to k0), and bond-bending effective force constant k

MAPbI3[6]MAPbBr3[7]
µ78.70257.666
Pb-X bond-stretching
ΘE|| [K]97.0 (3)117.3 (4)ΘD|| [K]LT: 148 (4); HT: 167 (5)
νE|| [THz]2.021 (5)2.444 (6)νD|| [THz]LT: 3.08 (8); HT: 3.48 (10)
k|| [eV/Å2]1.315 (3)1.410 (6)k|| [eV/Å2]LT: 1.43 (3); HT: 1.87 (5)
Pb-X bond-bending
$${\Theta}_{{E} \perp}$$ [K]LT: 46.9 (8); HT: 50 (1)HT: 46 (1)
$${V}_{{E} \perp}$$ [THz]LT: 0.98 (2); HT: 1.04 (2)HT: 0.96 (2)
k [eV/Å2]LT: 0.309 (6); HT: 0.348 (7)HT: 0.218 (5)

From the temperature-dependent relative interatomic distances ΔREXAFS and the averaged relative interatomic distances ΔRXRD resulting from the synchrotron XRD [Figure 11], Schuck et al. determine the relative “perpendicular MSRD” ($$\Delta < \Delta \mathrm{u}_{perp}^{2} >$$) for MAPbI3, which correspond to vibrations perpendicular to the lead-halide bonding direction (bond-bending vibrations)[6].

On the thermal expansion of the tetragonal phase of MAPbI<sub>3</sub> and MAPbBr<sub>3</sub>

Figure 11. Relative Pb-X expansion of MAPbI3 from Schuck et al.[6] (red) and MAPbBr3 (Weadock et al.[7]) (blue). Positive bond expansion ΔREXAFS (solid upward triangles) and negative tension effect (solid downward triangles) result in the overall crystallographic expansion ΔRXRD (open circles). Values for the cubic perovskite phase of MAPbI3 are also given from Whitfield et al.[36] (open red star symbol). Adapted with permission from Schuck et al.[6]. Copyright 2022 American Chemical Society.

The following relationship applied:[55]

$$ \begin{equation} \begin{aligned} \Delta < \Delta \mathrm{u}_{perp}^{2} > =2 * \mathrm{R}_{\mathrm{XRD}}\left(\Delta \mathrm{R}_{\mathrm{EXAFS}}-\Delta \mathrm{R}_{\mathrm{XRD}}\right) . \end{aligned} \end{equation} $$

The absolute values of the perpendicular MSRD for MAPbI3 were then obtained, analogously to the parallel MSRD, through fitting of $$\Delta < \Delta \mathrm{u}_{perp}^{2} >$$ with a correlated Einstein model[34]. The parallel MSRD values were fitted with two Einstein temperatures: one for the orthorhombic low-temperature range up to 160 K (LT), the other for the tetragonal high-temperature range up to 260 K (HT). This procedure leads to a solid explanation of the experimental values (also for the description of the γEXAFS-dependence [Figure 10B and C, Table 1]). The differing dimensionality ($$\Delta \mathrm{u}_{perp}^{2}=\Delta \mathrm{u}_{x}^{2}+\Delta \mathrm{u}_{y}^{2}$$) compared to the parallel MSRD makes it necessary to use the following term:[33,34]

$$ \begin{equation} \begin{aligned} <\Delta \mathrm{u}_{perp}^{2}>(T)=\frac{\hbar^{2}}{\mu k_{B}} \frac{1}{\theta_{E(perp)}} \operatorname{coth}\left(\frac{\theta_{E(perp)}}{2 T}\right) \end{aligned} \end{equation} $$

Analogous to k0, the corresponding perpendicular effective force constant k (bond-bending effective force constant) from the correlated Einstein model could be calculated for MAPbI3 [Table 1][6]. Analogous to the procedure for MAPbI3, the relative perpendicular MSRD and the bond-bending effective force constant were also determined for MAPbBr3 from the available data[7]. For MAPbBr3, this was possible only for the tetragonal and cubic phases since the values determined for ΔREXAFS and ΔRXRD in the orthorhombic phase [Figure 11], based on the results of Weadock et al., lead only to unphysical values for the relative perpendicular MSRD (here, the digitization of the data from Weadock et al. could not be carried out with sufficient accuracy for the orthorhombic phase)[7]. However, since the phase transformation from tetragonal to cubic is being discussed here, this is not limiting. The ratio γEXAFS = perpendicular MSRD/parallel MSRD describes the degree of anisotropy of the relative vibrations [Figure 10C][34]. The uncertainties pointed out by Schuck et al. may be of a similar (if not larger) magnitude in the behavior of MAPbBr3 shown here, so that the anisotropy of the relative vibrations and the perpendicular MSRD of MAPbI3 and MAPbBr3 are unlikely to differ significantly. However, perpendicular MSRD $$< \Delta \mathrm{u}_{perp}^{2} > (T)$$ reveals the negative tension effect in the lead-halide bond [Figure 11][6]. Tension is defined as $$-< \Delta \mathrm{u}_{perp}^{2} >$$/2RXRD[56]. In the tetragonal/cubic phases of MAPbI3 and MAPbBr3, strong negative tension effects result in a shortening of RXRD, which is accompanied by the observed negative changes in the partial volume of the [PbX6] octahedra. Tenison effects are often used to explain the structural mechanism of NTE[57], and these tension effects are also associated with the behavior of optical phonons[58,59]. Although Dalba et al. point out that “in general the Einstein frequencies cannot be correlated in a simple way to the phonon density of states”[60], the same authors draw parallels between the relative atomic vibrations of EXAFS and optical phonons[60,61].

Also, an attempt is made here to establish correlations between the relative atomic vibrations obtained from EXAFS studies on one hand and the behavior of optical phonons on the other. In addition, the Grüneisen parameters of individual phonon modes (also known as phonon deformation potentials) are of interest as a probe of the anharmonicity associated with phonon-phonon scattering and heat transport[59,62,63]. Two recent publications investigated the temperature dependence of optical phonons using infrared (IR) and terahertz measurements of MAPbI3 and MAPbBr3[64,65]. Two optical phonons at around 0.9 and 1.8 THz have been reported from Boldyrev et al. in MAPbI3 single crystals using temperature-dependent (figure 5 in Ref.[64]), high-resolution terahertz reflection spectroscopy[64]. The authors of the study attribute both modes to the inorganic [PbI6] framework structure, with the low-frequency mode dominated by bending modes, while stretching processes play a larger role in the higher-frequency mode. Both modes observed from Boldyrev et al. show parallels to the bond-stretching frequencies of νE|| = 2 THz and to the bond-bending frequency of $${\nu}_{{E} \perp}$$ = 1 THz for MAPbI3 both observed with EXAFS[64]. In a similar temperature-dependent time-domain THz transmission and far IR reflectance investigation on optical phonons in MAPbBr3 single crystals, the higher frequency mode measured by Železný et al. shows an increase in the frequency in the tetragonal phase from approx[65]. 2.07 THz to approx. 2.12 THz (figure 5 in Ref.[65]), which corresponds to a negative mode Grüneisen parameter for this mode. In contrast, the low-frequency mode at 1.4 THz observed by Železný et al. shows hardly any temperature dependence[65]. The similarities to the EXAFS results of MAPbBr3 are also present here as the bond-stretching frequencies of νE|| = 2.4 THz and the bond-bending frequency of $${\nu}_{{E} \perp}$$ = 1 THz for MAPbBr3 are observed with EXAFS. Note the frequency increase in the EXAFS results; for MAPbBr3, the bond-stretching frequencies determined with a Debye model are lower in the low-temperature range (3.1 THz) than in the high-temperature range (3.5 THz). This frequency increase can also be observed, although to a lesser extent, in the bond-stretching frequencies for MAPbI3.

CONCLUSION

The review of the results from the analysis of the averaged crystal structure with XRD diffraction methods shows temperature-dependent changes in the tetragonal phase with an apparent shortening of the Pb-X bonds and an apparent shrinkage of the [PbX6] octahedra in both MAPbI3 and MAPbBr3 with increasing temperature up to the cubic phase. The thermal expansion can best be described with the linear thermal expansion coefficient α[averaged] for the averaged tetragonal lattice constant aaveraged = ((2*atet)+ctet)/3 since the lowest temperature dependence occurs for this expansion coefficient. Using the Grüneisen tensor, positive Grüneisen parameters γ[001] and γ[100] can be expected for nearly the complete temperature range of the tetragonal phase of MAPbI3 (with the exception of γ[001] in a small temperature range close to the cubic phase transition above 307 K). The analysis of the tetragonal/cubic phase transformation using the standard Landau theory treatment seems to show a certain variance in the determination of the order parameters, but this can probably largely be explained by the quality of the underlying structure refinements. From the point of view of the local structure, it is remarkable that the thermal expansion of the very local arrangement of the lead halide octahedra is apparently unaffected by the tetragonal/cubic phase transformation. The temperature-dependent changes in the range of 150-300 K of the Pb-Br bond distances determined for MAPbBr3 by Bernasconi et al. from PDF refinements in the range of 1-10 Å using the orthorhombic Pnma perovskite crystal structure are of approximately the same order of magnitude as the bond expansion values ΔREXAFS determined by Weadock et al. with EXAFS[7,16]. Tension effects are observed with EXAFS in the tetragonal phase for both MAPbI3 and MAPbBr3. The positive bond extension and the negative tension effects of the lead-halide bond are approximately the same in MAPbI3 (and presumably also in MAPbBr3) in the orthorhombic phase. After the transition to the tetragonal phase, the equilibrium shifted in favor of the negative tension effects in both MAPbI3 and MAPbBr3. Further analysis of similarities between EXAFS relative atomic vibrations and optical phonon studies could be a rewarding way to better understand the thermal expansion in the tetragonal perovskite structure of MAPbI3 and MAPbBr3. This brief overview of the thermal expansion of MAPbI3 and MAPbBr3 can hopefully contribute to a more comprehensive description that includes all relevant inorganic halide perovskites and the discussion of NTE coefficients for ferroelectric oxide perovskites.

DECLARATIONS

Authors’ contributions

Contributed equally to this article: Schuck G, Többens DM, Schorr S

Availability of data and materials

Not applicable.

Financial support and sponsorship

None.

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author (s) 2024.

REFERENCES

1. Chouhan L, Ghimire S, Subrahmanyam C, Miyasaka T, Biju V. Synthesis, optoelectronic properties and applications of halide perovskites. Chem Soc Rev 2020;49:2869-85.

2. Pérez-Fidalgo L, Xu K, Charles BL, et al. Anomalous electron-phonon coupling in cesium-substituted methylammonium lead iodide perovskites. J Phys Chem C 2023;127:22817-26.

3. Spera EL, Pereyra CJ, Gau DL, Berruet M, Marotti RE. Excitonic optical properties of CH3NH3PbI3 perovskite and its dependence with temperature. MRS Adv 2024;9:39-44.

4. Jošt M, Kegelmann L, Korte L, Albrecht S. Monolithic perovskite tandem solar cells: a review of the present status and advanced characterization methods toward 30% efficiency. Adv Energy Mater 2020;10:1904102.

5. Duan L, Walter D, Chang N, et al. Stability challenges for the commercialization of perovskite-silicon tandem solar cells. Nat Rev Mater 2023;8:261-81.

6. Schuck G, Többens DM, Wallacher D, Grimm N, Tien TS, 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;126:5388-402.

7. Weadock NJ, Mackeen C, Qin X, et al. Thermal contributions to the local and long-range structural disorder in CH3NH3PbBr3. PRX Energy 2023;2:033004.

8. 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;123:11436-46.

9. Miyata K, Atallah TL, Zhu XY. Lead halide perovskites: crystal-liquid duality, phonon glass electron crystals, and large polaron formation. Sci Adv 2017;3:e1701469.

10. Tailor NK, Satapathi S. Crystalline-liquid duality of specific heat in halide perovskite semiconductor. Scr Mater 2023;223:115061.

11. Adams DJ, Churakov SV. Classification of perovskite structural types with dynamical octahedral tilting. IUCrJ 2023;10:309-20.

12. Liang X, Klarbring J, Baldwin WJ, Li Z, Csányi G, Walsh A. Structural dynamics descriptors for metal halide perovskites. J Phys Chem C Nanomater Interfaces 2023;127:19141-51.

13. Weadock NJ, Sterling TC, Vigil JA, et al. The nature of dynamic local order in CH3NH3PbI3 and CH3NH3PbBr3. Joule 2023;7:1051-66.

14. Beecher AN, Semonin OE, Skelton JM, et al. Direct observation of dynamic symmetry breaking above room temperature in methylammonium lead iodide perovskite. ACS Energy Lett 2016;1:880-7.

15. Page K, Siewenie JE, Quadrelli P, Malavasi L. Short-range order of methylammonium and persistence of distortion at the local scale in MAPbBr3 hybrid perovskite. Angew Chem Int Ed 2016;55:14320-4.

16. Bernasconi A, Malavasi L. Direct evidence of permanent octahedra distortion in MAPbBr3 hybrid perovskite. ACS Energy Lett 2017;2:863-8.

17. Bird TA, Chen J, Songvilay M, et al. Large dynamic scissoring mode displacements coupled to band gap opening in hybrid perovskites. arXiv 2021. Available from: https://arxiv.org/abs/2108.05751 [Last accessed on 7 Aug 2024].

18. Simenas M, Gagor A, Banys J, Maczka M. Phase transitions and dynamics in mixed three- and low-dimensional lead halide perovskites. Chem Rev 2024;124:2281-326.

19. Kutes Y, Ye L, Zhou Y, Pang S, Huey BD, Padture NP. Direct observation of ferroelectric domains in solution-processed CH3NH3PbI3 perovskite thin films. J Phys Chem Lett 2014;5:3335-9.

20. Bari M, Bokov AA, Ye Z. Ferroelastic domains and phase transitions in organic-inorganic hybrid perovskite CH3NH3PbBr3. J Mater Chem C 2021;9:3096-107.

21. Bari M, Bokov AA, Leach GW, Ye Z. Ferroelastic domains and effects of spontaneous strain in lead halide perovskite CsPbBr3. Chem Mater 2023;35:6659-70.

22. Wilson JN, Frost JM, Wallace SK, Walsh A. Dielectric and ferroic properties of metal halide perovskites. APL Mater 2019;7:010901.

23. Breternitz J. The “ferros” of MAPbI3: ferroelectricity, ferroelasticity and its crystallographic foundations in hybrid halide perovskites. Cryst Mater 2022;237:135-40.

24. Ambrosio F, De Angelis F, Goñi AR. The ferroelectric-ferroelastic debate about metal halide perovskites. J Phys Chem Lett 2022;13:7731-40.

25. Zheng W, Wang X, Zhang X, et al. Emerging halide perovskite ferroelectrics. Adv Mater 2023;35:e2205410.

26. Haeger T, Heiderhoff R, Riedl T. Thermal properties of metal-halide perovskites. J Mater Chem C 2020;8:14289-311.

27. Jacobsson TJ, Schwan LJ, Ottosson M, Hagfeldt A, Edvinsson T. Determination of thermal expansion coefficients and locating the temperature-induced phase transition in methylammonium lead perovskites using X-ray diffraction. Inorg Chem 2015;54:10678-85.

28. Bozec Y, Kaang S, Hine P, Ward I. The thermal-expansion behaviour of hot-compacted polypropylene and polyethylene composites. Composit Sci Technol 2000;60:333-44.

29. Becker P, Scyfried P, Siegert H. The lattice parameter of highly pure silicon single crystals. Z Physik B Condens Matter 1982;48:17-21.

30. Ge C, Hu M, Wu P, et al. Ultralow thermal conductivity and ultrahigh thermal expansion of single-crystal organic-inorganic hybrid perovskite CH3NH3PbX3 (X = Cl, Br, I). J Phys Chem C 2018;122:15973-8.

31. Zhou Y, Guo Z, Qaid SMH, Xu Z, Zhou Y, Zang Z. Strain engineering toward high-performance formamidinium-based perovskite solar cells. Solar RRL 2023;7:2300438.

32. Katan C, Mohite AD, Even J. Entropy in halide perovskites. Nat Mater 2018;17:377-9.

33. Fornasini P, Grisenti R. On EXAFS debye-waller factor and recent advances. J Synchrotron Rad 2015;22:1242-57.

34. Sanson A. EXAFS spectroscopy: a powerful tool for the study of local vibrational dynamics. Microstructures 2021;1:2021004.

35. Schuck G, Többens DM, 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;122:5227-37.

36. Whitfield PS, Herron N, Guise WE, et al. Structures, phase transitions and tricritical behavior of the hybrid perovskite methyl ammonium lead iodide. Sci Rep 2016;6:35685.

37. Franz A, Többens DM, Schorr S. Interaction between cation orientation, octahedra tilting and hydrogen bonding in methylammonium lead triiodide. Cryst Res Technol 2016;51:534-40.

38. Stoumpos CC, Malliakas CD, Kanatzidis MG. Semiconducting tin and lead iodide perovskites with organic cations: phase transitions, high mobilities, and near-infrared photoluminescent properties. Inorg Chem 2013;52:9019-38.

39. Celeste A, Capitani F. Hybrid perovskites under pressure: present and future directions. J Appl Phys 2022;132:220903.

40. 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;7:3458-66.

41. Gil-González E, Pérez-Maqueda LA, Sánchez-Jiménez PE, Perejón A. Paving the way to establish protocols: modeling and predicting mechanochemical reactions. J Phys Chem Lett 2021;12:5540-6.

42. Whitfield PS, Herron N, Guise WE, et al. Correction: Corrigendum: structures, phase transitions and tricritical behavior of the hybrid perovskite methyl ammonium lead iodide. Sci Rep 2017;7:42831.

43. Schorr S, Sheptyakov D. Low-temperature thermal expansion in sphalerite-type and chalcopyrite-type multinary semiconductors. J Phys Condens Matter 2008;20:104245.

44. Haussühl S. Kristallphysik; Weinheim, Germany: Physic-Verlag; 1983.

45. Feng J. Mechanical properties of hybrid organic-inorganic CH3NH3BX3 (B = Sn, Pb; X = Br, I) perovskites for solar cell absorbers. APL Mater 2014;2:081801.

46. Campbell BJ, Stokes HT, Tanner DE, Hatch DM. ISODISPLACE: a web-based tool for exploring structural distortions. J Appl Cryst 2006;39:607-14.

47. Liu J, Du J, Phillips AE, Wyatt PB, Keen DA, Dove MT. Neutron powder diffraction study of the phase transitions in deuterated methylammonium lead iodide. J Phys Condens Matter 2022;34:145401.

48. Egami T, Billinge SJL. Underneath the bragg peaks: structural analysis of complex materials; Oxford: Elsevier; 2003.

49. Billinge SJ. Nanoscale structural order from the atomic pair distribution function (PDF): there’s plenty of room in the middle. J Solid State Chem 2008;181:1695-700.

50. Bird TA, Herlihy A, Senn MS. Symmetry-adapted pair distribution function analysis (SAPA): a novel approach to evaluating lattice dynamics and local distortions from total scattering data. J Appl Cryst 2021;54:1514-20.

51. Bird TA, Woodland-Scott J, Hu L, et al. Anharmonicity and scissoring modes in the negative thermal expansion materials ScF3 and CaZrF6. Phys Rev B 2020;101:064306.

52. Ravel B, Newville M. ATHENA, ARTEMIS, HEPHAESTUS: data analysis for X-ray absorption spectroscopy using IFEFFIT. J Synchrotron Rad 2005;12:537-41.

53. Bunker G. Application of the ratio method of EXAFS analysis to disordered systems. Nucl Instrum Methods Phys Res 1983;207:437-44.

54. Bunker G. Introduction to EXAFS; Cambridge: Cambridge University Press; 2010.

55. Fornasini P, a Beccara S, Dalba G, et al. Extended X-ray-absorption fine-structure measurements of copper: local dynamics, anharmonicity, and thermal expansion. Phys Rev B 2004;70:174301.

56. Fornasini P, Grisenti R. The coefficient of bond thermal expansion measured by extended X-ray absorption fine structure. J Chem Phys 2014;141:164503.

57. Fornasini P. Vibrational anisotropy. In: Schnohr CS, Ridgway MC, editors. X-ray absorption spectroscopy of semiconductors. Berlin: Springer; 2015. pp. 127-42.

58. Attfield JP. Mechanisms and materials for NTE. Front Chem 2018;6:371.

59. Dove MT, Fang H. Negative thermal expansion and associated anomalous physical properties: review of the lattice dynamics theoretical foundation. Rep Prog Phys 2016;79:066503.

60. Dalba G, Diop D, Fornasini P, Rocca F. An EXAFS study of thermal disorder in GaAs. J Phys Condens Matter 1994;6:3599-608.

61. Dalba G, Fornasini P, Kuzmin A, Purans J, Rocca F. X-ray absorption spectroscopy study of ReO3 lattice dynamics. J Phys Condens Matter 1995;7:1199-213.

62. Talit K, Strubbe DA. Stress effects on vibrational spectra of a cubic hybrid perovskite: a probe of local strain. J Phys Chem C 2020;124:27287-99.

63. Gava V, Martinotto AL, Perottoni CA. First-principles mode Gruneisen parameters and negative thermal expansion in α-ZrW2O8. Phys Rev Lett 2012;109:195503.

64. Boldyrev KN, Anikeeva VE, Semenova OI, Popova MN. Infrared spectra of the CH3NH3PbI3 hybrid perovskite: signatures of phase transitions and of organic cation dynamics. J Phys Chem C 2020;124:23307-16.

65. Železný V, Kadlec C, Kamba S, Repček D, Kundu S, Saidaminov MI. Infrared and terahertz studies of phase transitions in the CH3NH3PbBr3 perovskite. Phys Rev B 2023;107:174113.

Cite This Article

Review
Open Access
On the thermal expansion of the tetragonal phase of MAPbI3 and MAPbBr3
Götz SchuckGötz Schuck, ... Susan Schorr

How to Cite

Schuck, G.; Többens D. M.; Schorr S. On the thermal expansion of the tetragonal phase of MAPbI3 and MAPbBr3. Microstructures. 2024, 4, 2024047. http://dx.doi.org/10.20517/microstructures.2024.33

Download Citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click on download.

Export Citation File:

Type of Import

Tips on Downloading Citation

This feature enables you to download the bibliographic information (also called citation data, header data, or metadata) for the articles on our site.

Citation Manager File Format

Use the radio buttons to choose how to format the bibliographic data you're harvesting. Several citation manager formats are available, including EndNote and BibTex.

Type of Import

If you have citation management software installed on your computer your Web browser should be able to import metadata directly into your reference database.

Direct Import: When the Direct Import option is selected (the default state), a dialogue box will give you the option to Save or Open the downloaded citation data. Choosing Open will either launch your citation manager or give you a choice of applications with which to use the metadata. The Save option saves the file locally for later use.

Indirect Import: When the Indirect Import option is selected, the metadata is displayed and may be copied and pasted as needed.

About This Article

Special Issue

© The Author(s) 2024. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
281
Downloads
45
Citations
0
Comments
0
1

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Microstructures
ISSN 2770-2995 (Online)

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/