2021
[1,2]\fnmNicholas J \surWeadock \equalcontThese authors contributed equally to this work. \equalcontThese authors contributed equally to this work.
[3]\fnmDmitry \surReznik
[1,2,14]\fnmMichael F \surToney
1]\orgdivMaterials Science and Engineering, \orgnameUniversity of Colorado, Boulder, \orgaddress\cityBoulder, \stateCO, \postcode80309, \countryUSA
2]\orgdivDepartment of Chemical and Biological Engineering, \orgnameUniversity of Colorado, Boulder, \orgaddress\cityBoulder, \stateCO, \postcode80309, \countryUSA
3]\orgdivDepartment of Physics, \orgnameUniversity of Colorado, Boulder, \orgaddress\cityBoulder, \stateCO, \postcode80309, \countryUSA
4]\orgdivDepartment of Chemical Engineering, \orgnameStanford University, \orgaddress\cityStanford, \stateCA, \postcode94305, \countryUSA
5]\orgdivDepartment of Chemistry, \orgnameStanford University, \orgaddress\cityStanford, \stateCA, \postcode94305, \countryUSA
6]\orgdivDepartment of Mechanical Science & Engineering and Materials Resesarch Laboratory, \orgnameUniversity of Illinois at Urbana-Champaign, \orgaddress\cityUrbana, \stateIL, \postcode61801, \countryUSA
7]\orgdivAdvanced Photon Source, \orgnameArgonne National Lab, \orgaddress\cityLemont, \stateIL, \postcode60439, \countryUSA
8]\orgdivNeutron Scattering Division, \orgnameOak Ridge National Laboratory, \orgaddress\cityOak Ridge, \stateTN, \postcode37830, \countryUSA
9]\orgdivISIS Facility, \orgnameRutherford Appleton Laboratory, \orgaddress\cityChilton, \stateDidcot, \postcodeOxon OX11 0QX, \countryUnited Kingdom
10]\orgdivDepartment of Physics, \orgnameRoyal Holloway University of London, \orgaddress\postcodeEgham TW20 0EX, \countryUnited Kingdom
11]\orgdivNIST Center for Neutron Research, \orgnameNational Institute of Standards and Technology, \orgaddress\cityGaithersburg, \stateMD, \postcode20899, \countryUSA
12]\orgdivDepartment Chemie, \orgnameUniverstät Paderborn, \orgaddress\cityPaderborn, \countryGermany
13]\orgdivStanford Institute for Materials and Energy Sciences, \orgnameSLAC National Accelerator Laboratory, \orgaddress\cityMenlo Park, \stateCA, \postcode94025, \countryUSA
14]\orgdivRenewable and Sustainable Energy Institute (RASEI), \orgnameUniversity of Colorado, Boulder, \orgaddress\cityBoulder, \stateCO, \postcode80303, \countryUSA
The nature of dynamic local order in CH3NH3PbI3 and CH3NH3PbBr3
Abstract
Hybrid lead halide perovskites (LHPs) are a class of semiconductor with novel properties that are distinctively governed by structural fluctuations. Diffraction experiments sensitive to average, long-range order reveal a cubic structure in the device-relevant, high-temperature phase. Local probes find additional short-range order with lower symmetry that may govern the structure-function relationships of LHPs. However, the dimensionality, participating atoms, and dynamics of this short-range order are unresolved, impeding our understanding of technologically relevant properties including long carrier lifetimes and facile halide migration. Here, we determine the true structure of two prototypical hybrid LHPs, CH3NH3PbI3 and CH3NH3PbBr3, using a combination of single-crystal X-ray and neutron diffuse scattering, neutron inelastic spectroscopy, and molecular dynamics simulations. The remarkable collective dynamics we found are not suggested by previous studies and consist of a network of local two-dimensional, circular pancake-like regions of dynamically tilting lead halide octahedra (lower symmetry) that induce longer range intermolecular correlations within the CH3NH sublattice. The dynamic local structure can introduce transient ferroelectric or antiferroelectric domains that increase charge carrier lifetimes, and strongly affect the halide migration, a poorly understood degradation mechanism. Our approach of co-analyzing single-crystal X-ray and neutron diffuse scattering data with MD simulations will provide unparalleled insights into the structure of hybrid materials and materials with engineered disorder.
keywords:
Metal halide perovskites, Diffuse scattering, Local structure, Ion migrationThe crystal structure and associated symmetry of a material are key determinants of mechanical, electronic, optical, and thermal properties. One has to look no further than seminal condensed matter physics textbooks for derivations of these properties made possible by consideration of the long-range or average order determined by the crystal lattice and translational symmetry ashcroftSolidStatePhysics2000 ; kittelIntroductionSolidState2019 . However, in many materials including disordered rocksalts and intercalation compounds used for battery cathodes, relaxor ferroelectrics, thermoelectrics, and oxide and halide perovskites, the important properties are not well described by the long-range structure. Instead it is short-range order that dominates aspects of the structure-function relationship clementCationdisorderedRocksaltTransition2020 ; krogstadReciprocalSpaceImaging2020a ; simonovHiddenDiversityVacancy2020a ; xuThreedimensionalMappingDiffuse2004 ; krogstadRelationLocalOrder2018 ; rothSimpleModelVacancy2020a ; lanigan-atkinsTwodimensionalOverdampedFluctuations2021 . In scattering experiments, short-range or local order manifests as diffuse scattering; a result of static and dynamic deviations from the average structure. Fixed chemical or local structural correlations result in static diffuse scattering, whereas thermal diffuse scattering arises from dynamic displacements due to lattice dynamics welberryOneHundredYears2016 . Resolving structural correlations in disordered materials has recently become more feasible with the development of high flux, single crystal X-ray and neutron diffuse scattering instruments and sophisticated modeling algorithms rosenkranzCorelliEfficientSingle2008 ; yeImplementationCrossCorrelation2018 ; krogstadReciprocalSpaceImaging2020a ; proffenAdvancesTotalScattering2009 ; simonovYellComputerProgram2014 ; morganRmcdiscordReverseMonte2021 , opening up enormous opportunities to understand how local order impacts materials properties.
Organic-inorganic metal halide perovskites are a recently re-invigorated class of semiconductors with remarkable optoelectronic performance that defies traditional intuition: Lead-based metal halide perovskites (LHPs) possess a mechanically soft, defect-tolerant crystal lattice with strong structural disorder and mobile ions at modest temperatures eggerWhatRemainsUnexplained2018 . Fluctuations in the orbital overlaps arising from large thermal displacements of iodide in methylammonium (CH3NH, CD3ND = MA) lead iodide directly influence the temperature dependence of the electron (or hole) mobility and optical bandgap mayersHowLatticeCharge2018a . X-ray and neutron diffraction measurements, which probe long-range order, report that the high-temperature phase is cubic with well-defined Bragg peaks wellerCompleteStructureCation2015 ; whitfieldStructuresPhaseTransitions2016 ; swainsonPhaseTransitionsPerovskite2003 . This cubic perovskite structure is shown in Figure 1a and consists of corner-sharing PbX6 (X = I-, Br-) octahedra surrounding a dynamically disordered MA+ cation within the cuboctahedral interstice. Measurements probing short-range order in the high-temperature phase, however, suggest the local structure is of lower symmetry beecherDirectObservationDynamic2016 ; cominLatticeDynamicsNature2016 ; lauritaChemicalTuningDynamic2017 ; zhaoPolymorphousNatureCubic2020 ; weadockTestDynamicDomainCritical2020 . The lack of consensus regarding the exact symmetry and dynamics of this enigmatic local structure limits our understanding and control of optoelectronic properties and ion migration in LHPs. Short-range order arising from dynamic two-dimensional correlations of lead bromide octahedra has recently been identified in CsPbBr3 lanigan-atkinsTwodimensionalOverdampedFluctuations2021 . These correlations are most prominent in the cubic phase above 433 K and may not play a significant role in CsPbBr3-based device operation. The significance and prevalence of such correlations in hybrid LHPs, including contributions from the organic cations, is not resolved.
Ion migration in LHPs, especially under illumination, is detrimental to device performance and stability yet the origin is not well understood. Consequences of ion migration include formation of space charge potentials at interfaces, accelerated degradation and loss of constituent elements, and light-induced phase segregation in mixed-halide compositions yuanIonMigrationOrganometal2016 . A complete picture of the high-temperature structure and dynamics is essential to model ion migration pathways accurately holekevichandrappaCorrelatedOctahedralRotation2021 . It is also important to characterize the short-range order on the organic sublattice, as the rotating MA+ molecular dipoles may screen band-edge charge carriers and extend carrier lifetimes chenOriginLongLifetime2017 .
We use single crystal X-ray and neutron diffuse scattering and neutron spectroscopy, combined with molecular dynamics (MD) simulations, to uncover the true structure and dynamics of the nominally simple cubic () phases of MAPbI3 (¿ 327 K) and MAPbBr3 (¿ 237 K). We find that the cubic phase (Fig. 1a, c) is comprised of dynamic, two-dimensional roughly circular “pancakes” of tilted lead halide octahedra which align along any of the three principal axes of the cubic structure and (Fig. 1b, d) induce additional structural correlations of the organic sublattice in the layers sandwiching the octahedra. Taken together, these regions of dynamic local order are several unit cells in diameter with lifetimes on the order of several picoseconds, resulting in a dynamic landscape for charge carriers and ion migration. This extended dynamic local order was neither observed nor predicted in previous studies of structural dynamics in hybrid LHPs ferreira_elastic_2018 ; gold-parkerAcousticPhononLifetimes2018 ; leguyDynamicsMethylammoniumIons2015 ; chenRotationalDynamicsOrganic2015 . Embedded within this dynamic structure we find additional static 3D droplets of the intermediate tetragonal phase, consistent with previous reports weadockTestDynamicDomainCritical2020 ; cominLatticeDynamicsNature2016 . The structural correlations on the MA+ sublattice are a unique characteristic in these hybrid systems with implications discussed below.

The experimental scattering function from the cubic phase of MAPbI3 at 345 K is shown in Fig. 2. Specifically, we present the L = 0.5, 1.5 planes measured with both X-ray (XDS, Fig. 2a,c, left panels) and neutron diffuse scattering (NDS, Fig 2b,d, left panels). calculated from atomic trajectories obtained with MD simulations of MAPbI3, see Methods, is plotted in the right-hand panels of a-d and both panels of g,h for comparison.
The experimental of the cubic phase of MAPbBr3 at 250 K is plotted in Fig. S1 and show diffuse scattering intensity profiles for XDS and NDS that are nearly identical to MAPbI3.

Our approach to solving the true structure of the nominally simple cubic MAPbI3 and MAPbBr3 involves decomposing the observed diffuse scattering profile into four components and analyzing their energy and Q-dependence. These components include: (1) rods of constant (XDS) or varied (NDS) intensity spanning the BZ edge (M-R direction in Fig. 1, ); (2) an additional contribution centered at the R-point []; (3) broad intensity centered at the X-point [] observed previously and discussed in the SI; and (4) contributions from the MA+ sublattice deduced from the alternating diffuse rod intensity profile observed in NDS but not XDS. This difference between XDS and NDS is highlighted with one-dimensional line cuts of XDS and NDS data shown in Figs. 2e and f. Specifically, the peaks at K = originating from the extended rod along H in XDS are not observed in the NDS data. Our analysis is complemented by MD simulations of MAPbI3 which reproduce the experimental diffuse scattering. The experimental and calculated XDS (Fig. 2) show remarkable agreement. Comparing the two results in a root-mean-square error of 3.9% (see Figs. S2,3 and related discussion) after adjusting only the background and scaling of the calculated . The agreement between the experimental and calculated NDS is very good, however there are small differences in intensities highlighted in Fig. S4 that likely arise from the classical potentials used here. Given the good agreement, we examine the real-space structure simulated with MD to determine the origin of the diffuse scattering.

The XDS rods of diffuse scattering have constant intensity along the entire BZ edge (Fig. S5) and a width larger than the instrument resolution (). This lineshape arises from two-dimensional structural correlations in real-space. We show these correlations in Fig. 3, which presents a simplified visualization of the MD simulation. These plots track PbI6 octahedral rotations, defined by azimuthal rotation angle about the [001] (Fig. 3a,b) or [010] (Fig. 3c,d) directions. Correlated regions of alternating large tilts are identified by neighboring red and blue pixels (representing octahedra) in Fig. 3a and d. Along the axis of rotation indicated in the corresponding schematic the tilts are uncorrelated between planes (Fig. 3b,c), revealing the existence of two-dimensional pancakes of tilted octahedra with a diameter on the order of 5 unit cells. These regions are reminiscent of tilts associated with the tetragonal structure in Fig. 1b. The tilts generate in-plane antiphase structural correlations with a periodicity of . The associated wave vector of the correlation is , which is the zone boundary for the simple cubic BZ. The tilted regions are confined to a single PbI6 sheet and therefore the diffuse scattering intensity manifests as rods.
The same structural correlations are present in the cubic phase of MAPbBr3, since the XDS and NDS are qualitatively identical to that of MAPbI3 (Fig. S1, 2). Extended diffuse rods observed in the all-inorganic CsPbBr3 lanigan-atkinsTwodimensionalOverdampedFluctuations2021 point to the existence of two-dimensional structural correlations of lead halide octahedra in all LHPs with a high temperature cubic structure.
Intermolecular structural correlations in the MA+ sublattice are implied from the alternating intensity pattern of the diffuse rods present in NDS but not XDS. The MA+ correlations are explored in the calculated in Figs. 2g,h, and S6-8. Specifically, we set the neutron scattering lengths of Pb and I to zero in the calculation to isolate the contribution from the MA+ cation (Figs. 2h, S6c, g, k and S7), and C, N, and D to zero to isolate contributions from the inorganic octahedra (Figs. 2g, S6b, f, j and S7). Diffuse scattering from the inorganic framework manifests as extended diffuse rods along the zone edge, resembling the experimental XDS as the X-ray atomic form factors for C, N, D are small compared to those for Pb and I and contribute little intensity. The calculated from CD3ND shows remarkable behavior (Figs. 2h, S6, S7), namely well-defined diffuse rods along the zone edges in addition to broad, isotropic intensity. This broad, isotropic component is a result of uncorrelated MA+ dynamics leguyDynamicsMethylammoniumIons2015 ; chenRotationalDynamicsOrganic2015 commonly associated with the nominally cubic phases of LHPs. The presence of additional zone edge intensity, however, shows that previously unreported intermolecular structural correlations exist on the MA+ sublattice. These local correlations are not purely 2D as the intensity varies along individual diffuse rods in the MA+ (Fig. S6,7). Layers of PbX6 octahedra are sandwiched between layers of MA+ cations, therefore we expect the out-of-plane MA+ structural correlations to extend at least two unit cells. Finally, the MA+ and PbX6 correlations are connected as evidenced by a nonzero interference term, shown in Fig. S8, which results in the alternating diffuse intensity profile observed in NDS.
We propose that MA+ orientation is driven by PbX6 tilts; the MA+ molecules orient to favor the lowest energy configuration defined by the distorted cuboctahedral geometry of the tilted regions and electrostatic interactions between lead halide octahedra and polar MA+ molecules zhu2019mixed ; lahnsteiner2016room . The existence of 2D correlations on the PbBr6 sublattice in CsPbBr3 supports this prediction as the Cs+ cations are not expected to influence the PbBr6 octahedra through hydrogen bonding. lanigan-atkinsTwodimensionalOverdampedFluctuations2021 .
The lateral size of 2D pancakes, given by the correlation length, , is obtained from the -linewidth of the XDS and NDS diffuse rods. We fit the intensity of one-dimensional cuts in across various diffuse rods to a Lorentzian lineshape in accordance with Ornstein-Zernike theory for exponentially decaying spatial correlations:
(1) |
where is the half-width at half-maximum (HWHM) of the fitted Lorentzian and is the correlation length xuThreedimensionalMappingDiffuse2004 ; valeCriticalFluctuationsSpin2019 . Correlation lengths of 4-6 unit cells (3 nm) in diameter are obtained for MAPbBr3 and MAPbI3 as reported in Table S1. Similar are obtained from NDS and XDS, indicating the in-plane of the 2D pancakes are nearly equivalent for the PbX6 and MA+ sublattices. We also calculate directly from the MD simulations, see Figs. S9-12 and related discussion, and find values consistent with experiment.
Molecular dynamics simulations show that the 2D structural correlations are transient and diffusive, with dynamics of the octahedral correlations tracked in Fig. 3 and Supplementary Videos 1 and 2. We investigate the dynamics of the PbI6 and MA+ correlations by evaluating the energy dependence obtained from the dynamical structure factor measured with neutron inelastic spectroscopy (INS) on MAPbI3 at 340 K. The in the L = 0.5 plane, integrated from meV is shown in the right-hand panel of Fig. 4a and matches that observed with NDS (Fig. 2b) and that calculated from MD (left-hand panel of Fig. 4b). A slice of along the diffuse rod presented in Fig. 4b shows that the intensity is centered at meV with no inelastic component visible. Incoherent-background-subtracted energy scans (integrated along the length of the diffuse rod) show a finite, quasielastic component as exemplified in Fig. 4d. Quasielastic scattering is a result of diffusive or relaxational motions involving small energy transfers, manifesting as a peak centered at meV with a non-zero linewidth. The scattering is fit to a relaxational (Lorentzian) model, broadened by the resolution function, as shown in Fig. 4d. The average energy HWHM, plotted in Fig. 4e, is meV with a corresponding lifetime of ps. Lifetimes are obtained in the same manner from the calculated as shown in Fig. S13, and the average lifetime of 6.3 ps is reported in Fig. 4e. We note that these lifetimes are consistent with the mean residence time for rotations of the C-N bond in cubic MAPbI3, suggesting that MA+ reorientations are in part driven by the PbX6 correlations chenRotationalDynamicsOrganic2015 . Overall, the structural correlations are dynamic with finite lifetime of several picoseconds.

R-point scattering (, Fig. 4c) noted above has been investigated previously and is attributed to static droplets of the intermediate-temperature tetragonal phase embedded within the high-temperature cubic phase, possibly nucleating about defects weadockTestDynamicDomainCritical2020 . We observe no inelastic scattering at the R-points in both the INS or MD in Fig. 4b. Incoherent-background-subtracted energy scans at a constant Q corresponding to R-points are best fit by a Gaussian lineshape with resolution limited linewidth (Fig. 4c), hence this scattering is static with nm (Table S1), consistent with previous reports weadockTestDynamicDomainCritical2020 . These droplets are distinct from the 2D pancakes but we are not able to resolve the spatial distribution of these two components. We expect that both structures influence MA+ orientation.
The appearance of small regions of a lower temperature phase above the phase transition temperature is often a hallmark of critical scattering. For phase transitions with critical fluctuations, there is a marked jump in intensity of the associated low temperature phase as the system is cooled through the transition stirling_critical_1996 . We investigate the temperature dependence of the diffuse scattering in Figs. S13,14 through the cubic-tetragonal transition temperatures for MAPbI3 and MAPbBr3. The intensity of the diffuse rods increases continuously with decreasing temperature through the transition, however no significant and abrupt jump in intensity is observed. Furthermore, the two-dimensional nature of the dynamic structural correlations is not typical of critical behavior at phase transitions. For the R-point scattering, however, there is a large jump in intensity as the cubic R-point transforms to the -point of the tetragonal phase, giving Bragg scattering. The critical nature of the R-point scattering has been addressed previously weadockTestDynamicDomainCritical2020 ; cominLatticeDynamicsNature2016 .
Our results show that the simple cubic phase of MAPbI3 and MAPbBr3 is in fact a composite structure containing dynamic 2D structural correlations and static tetragonal droplets with of a few nanometers. MA+ correlations, induced by octahedral tilts, extend 2-3 unit cells in the normal direction. The simple cubic phase is only recovered when this composite is averaged over space and time (Fig. 1 a-d). Structural fluctuations of lead halide octahedra generate large variations in the optical bandgap and electron-phonon coupling lanigan-atkinsTwodimensionalOverdampedFluctuations2021 ; mayersHowLatticeCharge2018a ; zhaoPolymorphousNatureCubic2020 , therefore it is essential to incorporate the observed correlations when modeling optoelectronic properties. We expect the dynamic correlations and static droplets to introduce spatial anisotropy to the electronic potential landscape with a longer lifetime than variations resulting from uncorrelated dynamic disorder. Furthermore, we must consider the effect of local correlations of the organic sublattice on LHP properties. MA+ has a large electric dipole moment of 2.3 Debye along the C-N axis chenOriginLongLifetime2017 . Uncorrelated dynamic disorder of the MA+ sublattice, as previously reported, would generate a zero or negligible dipole moment. The MA+ correlations observed here, however, may generate a net dipole moment resulting in transient, local ferroelectric or antiferroelectric domains with lifetimes on the order of 3-6 ps. These domains influence electron-hole recombination and shift band alignment biEnhancedPhotovoltaicProperties2017 ; liuChemicalNatureFerroelastic2018 . The dynamic structural correlations also affect ion migration. In solid-state ion conductors, rotational dynamics of polyanions are known to strongly impact cation mobility smithLowtemperaturePaddlewheelEffect2020 . Since the lifetime of the dynamic structural correlations (3-6 ps) exceeds the calculated time between attempted halide jumps (1 ps) frostWhatMovingHybrid2016 , a prefactor in calculating diffusivity, the correlations appear static and must be considered in future calculations. These correlations may impose a barrier to halide migration and reduce diffusivity holekevichandrappaCorrelatedOctahedralRotation2021 .
The two-dimensional dynamic structural correlations we observe may arise from the strongly anharmonic lattice dynamics observed in LHPs zhu2019mixed ; gold-parkerAcousticPhononLifetimes2018 ; songvilayCommonAcousticPhonon2019 ; ferreira_elastic_2018 . Indeed, the structural correlations in CsPbBr3 are reportedly driven by soft, overdamped, anharmonic phonons at the BZ edge lanigan-atkinsTwodimensionalOverdampedFluctuations2021 . We do not observe dispersive phonons along the BZ edge M-R branch in Fig. 4b, therefore these modes may be overdamped for MAPbI3 and MAPbBr3 as well. The structural correlations we observe on the PbX6 sublattice have a longer lifetime than those in CsPbBr3, and are prevalent at device relevant temperatures. The additional intermolecular correlations between MA+ are not present in CsPbBr3 and contribute to the improved optoelectronic properties of hybrid LHPs.
1 Acknowledgements
The authors acknowledge helpful discussions with Xixi Qin, Volker Blum, and Alex Zunger. This work (X-ray and neutron scattering) was supported by the Center for Hybrid Organic Inorganic Semiconductors for Energy, and Energy Frontier Research Center funded by the Office of Basic Energy Sciences, an office of science within the US Department of Energy (DOE). J. A. V. acknowledges fellowship support from the Stanford University Office of the Vice Provost of Graduate Education and the National Science Foundation Graduate Research Fellowship Program under Grant No. DGE – 1656518. H. I. K. acknowledges funding through the DOE Office of Basic Energy Sciences, Division of Materials Science and Engineering, under Contract No. DE-AC02-76SF0051. T.C.S. and D.R. acknowledge funding by the DOE Office of Basic Energy Sciences, Office of Science, under Contract No. DE-SC0006939. A portion of this research used resources at the Spallation Neutron Source, a DOE Office of Science User Facility operated by the Oak Ridge National Laboratory. Use of the Advanced Photon Source at Argonne National Laboratory was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, under Contract No. DE-AC02-06CH11357. Experiments at the ISIS Pulsed Neutron and Muon Source were supported by a beamtime allocation from the Science and Technology Facilities Council. Any mention of commercial products here is for information only; it does not imply recommendation or endorsement by the National Institute of Standards and Technology.
2 Methods
2.1 Single crystal growth
2.1.1 Materials
Methyl-d3-amine (CD3NH2, 99 at. % D) was purchased from Sigma-Aldrich. Hydrobromic acid (concentrated HBr solution, 48 % in water) was purchased from Acros Organics. Hydroiodic acid (concentrated HI solution, 47 % in water with 1.5 % hypophosphorous acid) was purchased from Sigma-Aldrich. Methanol-d4 and ethanol-d6 were purchased from Acros Organics and Cambridge Isotope Laboratories, Inc., respectively, with 99 at. % D purity. PbI2 (99.999+ %-Pb) and PbBr2 (98+ %, extra pure) were purchased from Strem Chemicals, Inc. and Acros Organics, respectively. CH3NH3I and CH3NH3Br were purchased from Greatcell Solar Materials. All solvents were of reagent grade or higher purity and anhydrous solvents were stored over molecular sieves or used directly from a JC Meyer solvent purification system.
2.1.2 Synthesis of CD3ND3X3 (X = Br, I)
Methyl-d3-amine was bubbled through ethanol (in a round bottom flask with flowing N2 in the headspace) in an ice water bath under vigorous stirring for approximately 2 minutes. Excess concentrated HBr was subsequently added dropwise with stirring. After the solution warmed naturally to room temperature (RT), the solvent was removed under reduced pressure at 60 °C. The resulting oil was re-dissolved in anhydrous ethanol and the product, CD3NH3Br, was precipitated as a colorless solid upon addition of the solution to cold diethyl ether. The solid was isolated by vacuum filtration and subsequently re-dispersed, sonicated, and isolated from diethyl ether three times before drying the product overnight under reduced pressure.
Solid CD3NH3Br was dissolved in ethanol-d6 and stirred for 3 h at 60 °C before precipitating the product in anhydrous diethyl ether and removing excess solvent under reduced pressure. This hydrogen-deuterium exchange was repeated twice more before finally isolating and drying the product, CD3ND3Br, at 65 °C overnight.
The synthesis of CD3ND3I followed an analogous procedure as for CD3ND3Br, except for the substitutions of: ethanol (or ethanol-d6 during the exchange) for methanol (or methanol-d4), and concentrated HBr for concentrated HI.
2.1.3 Crystallization of CD3ND3PbBr3
Solid CD3ND3Br (0.235 g, 2 mmol) and PbBr2 (0.731 g, 2 mmol) were combined in 2 mL of anhydrous dimethylformamide and the mixture was sonicated for 20 minutes. Once dissolved, the solution was passed through a 0.22-m hydrophilic PTFE filter and transferred to an oil bath at 78 °C. Crystals of CD3ND3PbBr3 slowly nucleated at the interface between the solution and the surface of the vial as the temperature increased slowly (1-3 °C h-1) from 78 °C (up to 85-100 °C). Crystals were isolated from the hot solution and the residual precursor solution was quickly removed from the surface of the crystal. Large crystals (200 mg) were obtained by isolating small crystals (25-50 mg) as seeds and re-filtering the precursor solution into a fresh vial with one seed crystal and quickly transferring back to the oil bath at 78 °C. Repeating the temperature ramp described above resulted in large crystal growth after a period of minor re-dissolution from the seed.
2.1.4 Crystallization of CH3NH3PbBr3
A 1-M solution of CH3NH3Br and PbBr2 in anhydrous dimethylformamide was prepared and sonicated for 1 h. Then 1 mL of the precursor solution was passed through a glass microfiber filter and transferred to a shell vial. The shell vial was placed inside a larger vial containing 4 mL of dichloromethane, the anti-solvent for the vapor diffusion crystallization, and sealed. Millimeter-scale (CH3NH3)PbBr3 crystals formed over the course of hours to days at room temperature.
2.1.5 Crystallization of CD3ND3PbI3 and CH3NH3PbI3
Solid CD3ND3I (0.594 g, 3.6 mmol) and PbI2 (1.66 g, 3.6 mmol) were combined in 3 mL of anhydrous -butyrolactone and dissolved at 70 °C with stirring. The solution was hot-filtered through a through a 0.22-m PTFE filter and immediately transferred to an oil bath at 70 °C. The temperature was increased to 110 °C over the course of approximately 2 h, followed by a slow ramp (1-3 °C h-1) up to 120-130 °C. Crystals were isolated and re-seeded as described for CD3ND3PbBr3 to reach the desired size.
The crystallization of CH3NH3PbI3 followed an analogous procedure as for CD3ND3PbI3, except for the substitution of CD3ND3I for CH3NH3I.
2.2 Diffuse X-ray scattering
Reciprocal space maps, including Bragg and diffuse scattering contributions, were collected in transmission geometry at the Advanced Photon Source Sector 6-ID-D using monochromatic 86.9 keV X-rays. This X-ray scattering is inherently energy integrated, incorporating contributions from both static and thermal diffuse disorder. Single crystal samples of (CD3ND3)PbI3, (CH3NH3)PbI3, and (CH3NH3)PbBr3, m in size were mounted on the tip of Kapton capillaries. At each temperature, the sample was rotated at with images collected every 0.1 s on a Pilatus 2M CdTe area detector. Sample temperatures were varied between 150 - 300 K with an Oxford N-Helix (check) Cryostream and 300 - 360 K with a hot nitrogen blower. The raw images are first processed with a peak finding algorithm to determine and refine an orientation matrix, and then rebinned into a reciprocal space volume 10 r.l.u. a side using the CCTW reduction workflowkrogstadReciprocalSpaceImaging2020a ; Cctw . The dataset is symmetrized using cubic point group operations to remove missing data resulting from gaps between detector pixel banks.
Sample damage is evident from changes in the scattering intensity with prolonged exposure. The damage thresholds were determined by identifying the onset of additional scattering intensity in subsequent measurements of the same sample at constant temperature. In MAPbI3, damage is apparent after 90 minutes of measurement time (including time when the shutter is off in between each rotation for data collection), therefore all diffuse scattering analyzed here was collected on a fresh sample within 90 minutes.
2.3 Diffuse neutron scattering
Neutron diffuse scattering measurements were performed on the CORELLI spectrometer at the Spallation Neutron Source at Oak Ridge National Lab in Oak Ridge, Tennessee, USArosenkranzCorelliEfficientSingle2008 ; yeImplementationCrossCorrelation2018 . Deuterated MAPbBr3 (300, 350 mg, rectangular prism) and MAPbI3 (220 mg, rhombic dodecahedron) crystals were used to reduce incoherent background contributions from hydrogen. The crystals were mounted in a CCR cryostat and oriented in the (HK0) scattering plane. The incident neutron energies range from 10 - 200 meV, and the 2 coverage spans -30 to 145∘. MAPbBr3 scattering data was collected from 150 - 300 K for 4-6 hours per temperature, and MAPbI3 scattering data was collected from 300 - 400 K at 12 hours per temperature to account for smaller sample size and increased background from the thermal shielding. Diffuse scattering volumes 6 r.l.u. a side were obtained by reducing the raw data using defined workflows implemented in MANTIDarnoldMantidDataAnalysis2014 . The contributions from inelastic scattering are removed by implementing the cross-correlation technique in the MANTID reduction workflow. The cross-correlation chopper selects elastically scattered neutrons within a resolution bandwidth of 0.9 meV full-width at half maximum yeImplementationCrossCorrelation2018 . Background artefacts including the aluminum sample enclosure are removed by subtracting the highest temperature datasets (300 K for MAPbBr3, 400 K for MAPbI3), normalized by the Bragg peaks, where little to no diffuse scattering is detected.
2.4 Neutron inelastic scattering
Neutron inelastic scattering measurements were performed on the Merlin direct geometry chopper spectrometer at the ISIS Neutron and Muon source at the Rutherford Appleton Laboratory in Didcot, UKbewleyMERLINNewHigh2006 . A deuterated single crystal of MAPbI3 weighing 698 mg was oriented in the (HKK) scattering plane and mounted in a Brookhaven-style aluminum sample can. Temperature control was performed using a closed-cycle refrigerator with heaters attached directly to the sample can. Three incident energies of 11, 22, and 65 meV were selected by utilizing the repetition rate multiplication mode with the Fermi chopper set to 250 Hz. The was collected at 340 K by rotating the sample 120∘ in 0.5∘ steps, and a radial collimator was used to reduce scattering from the aluminum sample can. Data reduction was performed using the HORACE data reduction suite and analyzed with Phonon Explorer and the National Institute of Standards and Technology (NIST) Center for Neutron Research (NCNR) Data Analysis and Visualization Environment (DAVE) software packages ewingsHoraceSoftwareAnalysis2016a ; reznikAutomatingAnalysisNeutron2020 ; phonon-explorer ; azuahDAVEComprehensiveSoftware2009 .
Energy scans for linewidth analysis were sliced from the 11 meV incident energy dataset. The resolution linewidth of this dataset was determined from constant- scans at values unique to the aluminum Debye-Scherrer rings. Several constant- scans obtained in this manner were fit to a Gaussian lineshape with constant background to obtain an average resolution linewidth of 0.5 meV FWHM. A representative constant- scan was used as the resolution function in subsequent analyses. The incoherent background was determined from constant- scans at values which contain only incoherent scattering with no contributions from diffuse rods, Bragg peaks, or aluminum Debye-Scherrer rings. Several background scans were averaged, then subtracted from the constant- scans along the diffuse rods and at the R-point to remove the incoherent background contribution and isolate the dynamics associated with these features.
Additional high-resolution neutron spectroscopy was performed on the cold neutron triple-axis spectrometer (SPINS) at the NCNR in Gaithersburg, MD, USA. The same deuterated single crystal of MAPbI3 weighing 698 mg was oriented in the (HKK) scattering plane and mounted in a Brookhaven-style aluminum sample can. Constant- scans, in which we vary the energy transfer while maintaining constant momentum transfer , were performed with a fixed final energy meV. We compare constant- scans taken along and at at 300 and 140 K, to constant- cuts calculated from the MD S() as a way to validate the calculations.
2.5 Molecular dynamics simulations
MD trajectories for deuterated MAPbI3 are generated using identical simulations to those in ref. zhu2019mixed with the mass of hydrogen set to that of deuterium. We used a super-cell that is based on the 12-atom pseudocubic unit cell with cell length . The simulations were done in the LAMMPS package plimpton1995fast . The inter-atomic potential is from ref. mattoni2015methylammonium . Trajectories were integrated using a 0.5 fs time step throughout. For the first 20 ps, the system was thermalized in the NPT ensemble at 300 K and 0 Pa, then it was equilibrated for another 20 ps in the NVE ensemble. After equilibration, the equations of motion were integrated for 20 ns and trajectories were written to a file every 50 fs.
2.6 S(Q,E) calculations
The dynamical structure factor, S(), can be calculated from MD. Given a set of suitably accurate trajectories, , S() can be evaluated directly as
(2) |
See the supplementary information supp_info and refs. therein dove1993introduction ; brown2006intensity ; hazemann2005high ; squires1996introduction ; harrelson2021computing ; zushi2015effect ; xiong2017native ; van1954correlations for a derivation of eq. 2. The theoretical calculations were done with a Python code developed by us sterling_pynamic . The calculated is weighted by X-ray atomic form factors or thermal neutron scattering lengths for direct comparison to experimental data hazemann2005high ; brown2006intensity . If the experiment uses neutrons, is the independent scattering length; if the experiment uses X-rays, is the dependent atomic form factor where is the magnitude of momentum transferred, , from the incident beam of particles into the material. is the energy transferred into the material and the sum runs over all atoms.
The reciprocal space resolution from MD is along the directions and along the direction. We do not integrate the calculation over reciprocal space unless otherwise specified, so the pixels in theoretical S(,) plots have this spacing. In order to ensure we are far from the transient thermalization regime in MD, we only sample the last nano-second of the simulation to calculate S(,). The last nano-second is split into 20 blocks that are 50 ps long; S(,) is then calculated from 10 non-consecutive blocks and averaged over them. For 50 ps blocks with a 50 fs sampling interval, the energy resolution is about meV and the maximum resolvable energy is 41.4 meV.
To compare to the experimental NDS data in Fig. 2, theoretical S(,) is integrated between meV, which is comparable to the resolution in the experiment. To compare to the experimental XDS data in Fig. 2, theoretical S(,) is integrated between meV. To compare with INS data in Fig. 4, both the experimental and theoretical S(,) are integrated meV. Anywhere that purely theoretical data is compared, the integration is meV. For excitations [e.g. Fig. 4(b)], theoretical is not integrated over energy; the MD energy resolution is set by the length of the trajectory sampled in the calculation.
The workflow and analysis outlined above was performed on MD simulations of protonated MAPbI3 and no significant differences were observed. The deuterated results are used in the main text for direct comparison to the experimental measurements.
2.7 Plotting
The unit cells in Fig. 1a, b were plotted in vesta momma2011vesta . The volumetric plot in Fig. 1e was made in mayavi ramachandran2011mayavi . Most other plots were made using the matplotlib package for Python matplotlib .
3 Data and Software Availability
The experimental and computational datasets used in the analysis here will be made available upon reasonable request. INS data associated with this experiment will be made publicly available by the Science and Technology Facilities Council at DOI: 10.5286/ISIS.E.RB2010431 The code used to calculate is available for free online sterling_pynamic .
4 Author Contributions
N.J.W., H.-G.S., and M.F.T. conceived the project. N.J.W., J.A.V., H.-G.S., F.Y., and M.J.K. performed the XDS and NDS experiments and data reduction and N.J.W., T.C.S., D.R., and M.F.T. performed the analysis. J.A.V., A.G.-P., I.C.S., and H.I.K. synthesized and mounted the single crystals. T.C.S. performed and analyzed the calculations from the MD datasets provided by B.A. and E.E. N.J.W., T.C.S., D.V., A.G-.P., I.C.S., P.M.G., and D.R. performed the INS experiments and contributed to the analysis. N.J.W. and T.C.S. prepared the figures and wrote the manuscript with contributions from all authors.
References
- \bibcommenthead
- (1) Ashcroft, N.W., Mermin, N.D.: Solid State Physics. Harcourt, New York (2000)
- (2) Kittel, C., McEuen, P., John Wiley & Sons.: Introduction to Solid State Physics. John Wiley & Sons, Hoboken, NJ (2019)
- (3) Clément, R.J., Lun, Z., Ceder, G.: Cation-disordered rocksalt transition metal oxides and oxyfluorides for high energy lithium-ion cathodes. Energy & Environmental Science 13(2), 345–373 (2020). https://doi.org/10.1039/C9EE02803J
- (4) Krogstad, M.J., Rosenkranz, S., Wozniak, J.M., Jennings, G., Ruff, J.P.C., Vaughey, J.T., Osborn, R.: Reciprocal space imaging of ionic correlations in intercalation compounds. Nature Materials 19(1), 63–68 (2020). https://doi.org/10.1038/s41563-019-0500-7
- (5) Simonov, A., De Baerdemaeker, T., Boström, H.L.B., Ríos Gómez, M.L., Gray, H.J., Chernyshov, D., Bosak, A., Bürgi, H.-B., Goodwin, A.L.: Hidden diversity of vacancy networks in Prussian blue analogues. Nature 578(7794), 256–260 (2020). https://doi.org/10.1038/s41586-020-1980-y
- (6) Xu, G., Zhong, Z., Hiraka, H., Shirane, G.: Three-dimensional mapping of diffuse scattering in . Physical Review B 70(17), 174109 (2004). https://doi.org/10.1103/PhysRevB.70.174109
- (7) 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., Chmaissem, O., Ye, Z.-G., Phelan, D.: The relation of local order to material properties in relaxor ferroelectrics. Nature Materials 17(8), 718–724 (2018). https://doi.org/10.1038/s41563-018-0112-7
- (8) Roth, N., Zhu, T., Iversen, B.B.: A simple model for vacancy order and disorder in defective half-Heusler systems. IUCrJ 7(4), 673–680 (2020). https://doi.org/10.1107/S2052252520005977
- (9) 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., Osborn, R., Delaire, O.: Two-dimensional overdamped fluctuations of the soft perovskite lattice in . Nature Materials, 1–7 (2021). https://doi.org/10.1038/s41563-021-00947-y
- (10) Welberry, T.R., Weber, T.: One hundred years of diffuse scattering. Crystallography Reviews 22(1), 2–78 (2016). https://doi.org/10.1080/0889311X.2015.1046853
- (11) Rosenkranz, S., Osborn, R.: Corelli: Efficient single crystal diffraction with elastic discrimination. Pramana 71(4), 705–711 (2008). https://doi.org/10.1007/s12043-008-0259-x
- (12) Ye, F., Liu, Y., Whitfield, R., Osborn, R., Rosenkranz, S.: Implementation of cross correlation for energy discrimination on the time-of-flight spectrometer CORELLI. Journal of Applied Crystallography 51(2), 315–322 (2018). https://doi.org/10.1107/S160057671800403X
- (13) Proffen, T., Kim, H.: Advances in total scattering analysis. Journal of Materials Chemistry 19(29), 5078–5088 (2009). https://doi.org/10.1039/B821178G
- (14) Simonov, A., Weber, T., Steurer, W.: Yell: A computer program for diffuse scattering analysis via three-dimensional delta pair distribution function refinement. Journal of Applied Crystallography 47(3), 1146–1152 (2014). https://doi.org/10.1107/S1600576714008668
- (15) Morgan, Z.J., Zhou, H.D., Chakoumakos, B.C., Ye, F.: Rmc-discord: Reverse Monte Carlo refinement of diffuse scattering and correlated disorder from single crystals. Journal of Applied Crystallography 54(6), 1867–1885 (2021). https://doi.org/10.1107/S1600576721010141
- (16) Egger, D.A., Bera, A., Cahen, D., Hodes, G., Kirchartz, T., Kronik, L., Lovrincic, R., Rappe, A.M., Reichman, D.R., Yaffe, O.: What Remains Unexplained about the Properties of Halide Perovskites? Advanced Materials 30(20), 1800691 (2018). https://doi.org/10.1002/adma.201800691
- (17) 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 Letters 18(12), 8041–8046 (2018). https://doi.org/10.1021/acs.nanolett.8b04276
- (18) Weller, M.T., Weber, O.J., Henry, P.F., Pumpo, A.M.D., Hansen, T.C.: Complete structure and cation orientation in the perovskite photovoltaic methylammonium lead iodide between 100 and 352 K. Chemical Communications 51(20), 4180–4183 (2015). https://doi.org/10.1039/C4CC09944C
- (19) 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. Scientific Reports 6, 35685 (2016). https://doi.org/10.1038/srep35685
- (20) Swainson, I.P., Hammond, R.P., Soullière, C., Knop, O., Massa, W.: Phase transitions in the perovskite methylammonium lead bromide, CH3ND3PbBr3. Journal of Solid State Chemistry 176(1), 97–104 (2003). https://doi.org/10.1016/S0022-4596(03)00352-9
- (21) 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 Letters 1(4), 880–887 (2016). https://doi.org/10.1021/acsenergylett.6b00381
- (22) 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., Sargent, E.H.: Lattice dynamics and the nature of structural transitions in organolead halide perovskites. Physical Review B 94(9), 094301 (2016). https://doi.org/10.1103/PhysRevB.94.094301
- (23) 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. Chemical Science 8(8), 5628–5635 (2017). https://doi.org/10.1039/C7SC01429E
- (24) Zhao, X.-G., Dalpian, G.M., Wang, Z., Zunger, A.: Polymorphous nature of cubic halide perovskites. Physical Review B 101(15), 155137 (2020). https://doi.org/10.1103/PhysRevB.101.155137
- (25) 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. Physical Review Letters 125(7), 075701 (2020). https://doi.org/10.1103/PhysRevLett.125.075701
- (26) Yuan, Y., Huang, J.: Ion Migration in Organometal Trihalide Perovskite and Its Impact on Photovoltaic Efficiency and Stability. Accounts of Chemical Research 49(2), 286–293 (2016). https://doi.org/10.1021/acs.accounts.5b00420
- (27) Holekevi Chandrappa, M.L., Zhu, Z., Fenning, D.P., Ong, S.P.: Correlated Octahedral Rotation and Organic Cation Reorientation Assist Halide Ion Migration in Lead Halide Perovskites. Chemistry of Materials 33(12), 4672–4678 (2021). https://doi.org/10.1021/acs.chemmater.1c01175
- (28) 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., Yoon, M., Chang, Y.-M., Choi, J.J., Lee, S.-H.: Origin of long lifetime of band-edge charge carriers in organic–inorganic lead iodide perovskites. Proceedings of the National Academy of Sciences 114(29), 7519–7524 (2017). https://doi.org/10.1073/pnas.1704421114
- (29) Ferreira, A.C., Létoublon, A., Paofai, S., Raymond, S., Ecolivet, C., Rufflé, B., Cordier, S., Katan, C., Saidaminov, M.I., Zhumekenov, A.A., Bakr, O.M., Even, J., Bourges, P.: Elastic Softness of Hybrid Lead Halide Perovskites. Physical Review Letters 121(8), 085502 (2018). https://doi.org/10.1103/PhysRevLett.121.085502
- (30) Gold-Parker, A., Gehring, P.M., Skelton, J.M., Smith, I.C., Parshall, D., Frost, J.M., Karunadasa, H.I., Walsh, A., Toney, M.F.: Acoustic phonon lifetimes limit thermal transport in methylammonium lead iodide. Proceedings of the National Academy of Sciences 115(47), 11905–11910 (2018). https://doi.org/10.1073/pnas.1812227115
- (31) 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., Nelson, J., Cabral, J.T., Barnes, P.R.F.: The dynamics of methylammonium ions in hybrid organic–inorganic perovskite solar cells. Nature Communications 6, 7124 (2015). https://doi.org/10.1038/ncomms8124
- (32) 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 perovskite. Physical Chemistry Chemical Physics 17(46), 31278–31286 (2015). https://doi.org/10.1039/C5CP05348J
- (33) 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 . Physical Review Materials 1(4), 042401 (2017). https://doi.org/10.1103/PhysRevMaterials.1.042401
- (34) See Supplemental Material at [URL will be inserted by publisher] for derivation of S(Q,E) from MD as well as other supporting information.
- (35) Zhu, T., Ertekin, E.: Mixed phononic and non-phononic transport in hybrid lead halide perovskites: glass-crystal duality, dynamical disorder, and anharmonicity. Energy & Environmental Science 12(1), 216–229 (2019). https://doi.org/10.1039/C8EE02820F
- (36) Lahnsteiner, J., Kresse, G., Kumar, A., Sarma, D., Franchini, C., Bokdam, M.: Room-temperature dynamic correlation between methylammonium molecules in lead-iodine based perovskites: An ab initio molecular dynamics perspective. Physical Review B 94(21), 214114 (2016). https://doi.org/10.1103/PhysRevB.94.214114
- (37) Vale, J.G., Boseggia, S., Walker, H.C., Springell, R.S., Hunter, E.C., Perry, R.S., Collins, S.P., McMorrow, D.F.: Critical fluctuations in the spin–orbit Mott insulator . Journal of Physics: Condensed Matter 31(18), 185803 (2019). https://doi.org/10.1088/1361-648X/ab0471
- (38) Stirling, W.G., Perry, S.C.: Critical scattering. In: Introduction to Neutron Scattering, pp. 86–102 (1996)
- (39) Bi, F., Markov, S., Wang, R., Kwok, Y., Zhou, W., Liu, L., Zheng, X., Chen, G., Yam, C.: Enhanced Photovoltaic Properties Induced by Ferroelectric Domain Structures in Organometallic Halide Perovskites. The Journal of Physical Chemistry C 121(21), 11151–11158 (2017). https://doi.org/10.1021/acs.jpcc.7b03091
- (40) Liu, Y., Collins, L., Proksch, R., Kim, S., Watson, B.R., Doughty, B., Calhoun, T.R., Ahmadi, M., Ievlev, A.V., Jesse, S., Retterer, S.T., Belianinov, A., Xiao, K., Huang, J., Sumpter, B.G., Kalinin, S.V., Hu, B., Ovchinnikova, O.S.: Chemical nature of ferroelastic twin domains in perovskite. Nature Materials 17(11), 1013–1019 (2018). https://doi.org/10.1038/s41563-018-0152-z
- (41) Smith, J.G., Siegel, D.J.: Low-temperature paddlewheel effect in glassy solid electrolytes. Nature Communications 11(1), 1483 (2020). https://doi.org/10.1038/s41467-020-15245-5
- (42) Frost, J.M., Walsh, A.: What Is Moving in Hybrid Halide Perovskite Solar Cells? Accounts of Chemical Research 49(3), 528–535 (2016). https://doi.org/10.1021/acs.accounts.5b00431
- (43) 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., Brown, C.M., Chernyshov, D., van Beek, W., Cochran, S., Stock, C.: Common acoustic phonon lifetimes in inorganic and hybrid lead halide perovskites. Physical Review Materials 3(9), 093602 (2019). https://doi.org/10.1103/PhysRevMaterials.3.093602
- (44) Cctw. https://sourceforge.net/projects/cctw/
- (45) Arnold, O., Bilheux, J.C., Borreguero, J.M., Buts, A., Campbell, S.I., Chapon, L., Doucet, M., Draper, N., Ferraz Leal, R., Gigg, M.A., Lynch, V.E., Markvardsen, A., Mikkelson, D.J., Mikkelson, R.L., Miller, R., Palmen, K., Parker, P., Passos, G., Perring, T.G., Peterson, P.F., Ren, S., Reuter, M.A., Savici, A.T., Taylor, J.W., Taylor, R.J., Tolchenov, R., Zhou, W., Zikovsky, J.: Mantid—Data analysis and visualization package for neutron scattering and SR experiments. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 764, 156–166 (2014). https://doi.org/10.1016/j.nima.2014.07.029
- (46) Bewley, R.I., Eccleston, R.S., McEwen, K.A., Hayden, S.M., Dove, M.T., Bennington, S.M., Treadgold, J.R., Coleman, R.L.S.: MERLIN, a new high count rate spectrometer at ISIS. Physica B: Condensed Matter 385–386, 1029–1031 (2006). https://doi.org/10.1016/j.physb.2006.05.328
- (47) Ewings, R.A., Buts, A., Le, M.D., van Duijn, J., Bustinduy, I., Perring, T.G.: Horace: Software for the analysis of data from single crystal spectroscopy experiments at time-of-flight neutron instruments. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 834, 132–142 (2016). https://doi.org/10.1016/j.nima.2016.07.036
- (48) Reznik, D., Ahmadova, I.: Automating Analysis of Neutron Scattering Time-of-Flight Single Crystal Phonon Data. Quantum Beam Science 4(4), 41 (2020). https://doi.org/10.3390/qubs4040041
- (49) Reznik, D.: Phonon-Explorer. https://github.com/dmitryr1234/Phonon-Explorer
- (50) Azuah, R.T., Kneller, L.R., Qiu, Y., Tregenna-Piggott, P.L.W., Brown, C.M., Copley, J.R.D., Dimeo, R.M.: DAVE: A comprehensive software suite for the reduction, visualization, and analysis of low energy neutron spectroscopic data. Journal of Research of the National Institute of Standards and Technology 114, 341–358 (2009). https://doi.org/10.6028/jres.114.025
- (51) Plimpton, S.: Fast parallel algorithms for short-range molecular dynamics. Journal of computational physics 117(1), 1–19 (1995). https://doi.org/10.1006/jcph.1995.1039
- (52) Mattoni, A., Filippetti, A., Saba, M., Delugas, P.: Methylammonium rotational dynamics in lead halide perovskite by classical molecular dynamics: the role of temperature. The Journal of Physical Chemistry C 119(30), 17421–17428 (2015). https://doi.org/10.1021/acs.jpcc.5b04283
- (53) Dove, M.T., Dove, M.T.: Introduction to Lattice Dynamics. Cambridge university press, Cambridge, United Kingdom (1993). https://doi.org/10.1017/CBO9780511619885
- (54) Brown, P., Fox, A., Maslen, E., O’keefe, M., Willis, B.: Intensity of diffracted intensities. In: International Tables for Crystallography vol. C, pp. 554–595. Wiley Online Library, Hoboken, NJ, United States (2006). Chap. 6.1. https://doi.org/10.1107/97809553602060000600
- (55) Hazemann, I., Dauvergne, M., Blakeley, M., Meilleur, F., Haertlein, M., Van Dorsselaer, A., Mitschler, A., Myles, D.A., Podjarny, A.: High-resolution neutron protein crystallography with radically small crystal volumes: application of perdeuteration to human aldose reductase. Acta Crystallographica Section D: Biological Crystallography 61(10), 1413–1417 (2005). https://doi.org/10.1107/S0907444905024285
- (56) Squires, G.L.: Introduction to the Theory of Thermal Neutron Scattering. Cambridge University Press, Cambridge, United Kingdom (2012). https://doi.org/10.1017/CBO9781139107808
- (57) Harrelson, T.F., Dettmann, M., Scherer, C., Andrienko, D., Moulé, A.J., Faller, R.: Computing inelastic neutron scattering spectra from molecular dynamics trajectories. Scientific reports 11(1), 1–12 (2021). https://doi.org/10.1038/s41598-021-86771-5
- (58) Zushi, T., Ohmori, K., Yamada, K., Watanabe, T.: Effect of a sio2 layer on the thermal transport properties of si nanowires: A molecular dynamics study. Physical Review B 91(11), 115308 (2015). https://doi.org/10.1103/PhysRevB.91.115308
- (59) Xiong, S., Selli, D., Neogi, S., Donadio, D.: Native surface oxide turns alloyed silicon membranes into nanophononic metamaterials with ultralow thermal conductivity. Physical Review B 95(18), 180301 (2017). https://doi.org/10.1103/PhysRevB.95.180301
- (60) Van Hove, L.: Correlations in space and time and born approximation scattering in systems of interacting particles. Physical Review 95(1), 249 (1954). https://doi.org/10.1103/PhysRev.95.249
- (61) Sterling, T.C.: Pynamic Structure Factor. https://github.com/tyst3273/pynamic-structure-factor
- (62) Momma, K., Izumi, F.: Vesta 3 for three-dimensional visualization of crystal, volumetric and morphology data. Journal of applied crystallography 44(6), 1272–1276 (2011). https://doi.org/10.1107/S0021889811038970
- (63) Ramachandran, P., Varoquaux, G.: Mayavi: 3d visualization of scientific data. Computing in Science & Engineering 13(2), 40–51 (2011). https://doi.org/10.1109/MCSE.2011.35
- (64) Hunter, J.D.: Matplotlib: A 2d graphics environment. Computing in Science Engineering 9(3), 90–95 (2007). https://doi.org/10.1109/MCSE.2007.55