

research papers
Direct derivation of anisotropic atomic displacement parameters from
simulations in extended solids with substitutional disorder using a neural network potentialaDepartment of Energy and Environment, National Institute of Advanced Industrial Science
and Technology (AIST), 1-8-31, Midorigaoka, Ikeda, Osaka 563-8577, Japan
*Correspondence e-mail: [email protected]
Atomic displacement parameters (ADPs) are crystallographic information describing the statistical distribution of atoms around an atom site. Anisotropic ADPs by atom were directly derived from classical 8SnSe6, Na2In2Sn4 and BaCu1.14In0.86P2. Unlike the very frequently used lattice dynamics approach, the MD approach can obtain ADPs in crystals with substitutional disorder and explicitly at finite temperature, but not under conditions where atoms migrate in the crystal. The calculated ADP approaches 0 when the temperature approaches 0, and the ADP is proportional to the temperature when the atom is in a harmonic potential and the sole contribution to the actual non-zero ADP is from the zero-point motion. The zero-point motion contribution can be estimated from the proportionality constant assuming this Einstein model. ADPs from MD simulations could act as a tool complementing experimental efforts to understand the including the distribution of atoms around atom sites.
(MD) simulations using a universal machine-learned potential. The (co)valences of atom positions were taken over recordings at different time steps in a single MD simulation. The procedure is demonstrated on extended solids, namely rocksalt structure MgO and three thermoelectric materials, AgKeywords: anisotropic atomic displacement parameters; ADPs; neural network potentials; molecular dynamics simulations; thermoelectric materials.
1. Introduction
Atomic displacement parameters (ADPs) are often provided as part of crystallographic
structural data and may represent atomic motion, possible static displacive disorder
and thermal vibration (Trueblood et al., 1996). In the past, there was discussion of `thermal vibrations' (Willis & Pryor, 1975
), but the International Union of Crystallography now recommends using the term `ADP'
(Trueblood et al., 1996
). ADPs are generally anisotropic. Calculation of ADPs is important in particular
to model light elements such as H, where obtaining ADPs from X-ray diffraction is
difficult, and to provide clues for obtaining a better picture of the actual through refinement.
Theoretical derivations of ADPs are typically performed indirectly through lattice
dynamics analysis, where the dynamical matrix is obtained and the vibrational, or
phonon, frequency of each mode is calculated. Examples of formalisms may be found
in the work of Erba et al. (2013), Madsen et al. (2013
), Lane et al. (2012
).
The lattice dynamics approach is difficult to apply in systems where disorder of elements on a (sub)lattice plays a critical role. Models with large supercells and/or a very short range order on the (sub)lattice are required and the symmetry is different from the experimental model. Ordering may significantly affect the final results in a small
with short-range ordering.Split-site systems are also problematic with lattice dynamics. One example of a split site is a double-well potential with minima at two very close positions. The average position of an atom at this site will be the middle of the two positions at elevated temperatures, while the atom will be at one of the minima at the 0 K limit. Placing the atom at the elevated temperature position leads to an imaginary mode at the Γ point.
In contrast, direct derivation of anisotropic ADPs, assuming a normal (or Gaussian)
distribution, from ΔξCiΔξCj〉, are simply the (co)variances of atom positions, for instance Cov(ΔξCi + ξCi, ΔξCj + ξCj). Here, ξCi and ΔξCi are the i component of the mean atom position vector and the displacement vector, respectively,
in Cartesian coordinates for a given atom. A large number of displacement values for
a given crystal can be obtained from MD simulations, and derivation of ADPs from these
values is very straightforward. This approach was used to obtain the anisotropic ADPs
of ND3 (Reilly et al., 2007) and benzophenone (Reilly et al., 2013
).
The ADP of an individual atom and that of an atom site must be clearly distinguished. The (co)variance of positions for a single atom is obtained in the former, while the (co)variance in the latter is taken over positions of multiple atoms that are each affine transformed such that the images of atoms are at almost the same position. The latter is experimentally observed, while the former is readily obtained using MD.
The direct method can, in principle, derive ADPs for any system. One notable exception is when atoms migrate during the simulations and are not trapped around an equilibrium point. The drawback of the direct method is that each MD simulation gives different results and obtaining very reliable values requires a long simulation time and/or a large
which could be computationally expensive using first-principles calculations.Verifying the accuracy of ADPs is a difficult issue to address. The ADPs in 3)4(NO2)2 showed average differences of ADPs for non-hydrogen atoms on the order of a few 0.0001 Å2, or less than 10%, between 9 (1) K X-ray diffraction (XRD) and 13 (1) K neutron diffraction
measurements (Iversen et al., 1996). The ADPs from classical MD simulations suffer some uncertainty from statistical
processing in addition to systematic deviations from the true value from the use of
approximations and the difference in classical and quantum dynamics.
This paper demonstrates the use of universal machine-learned neural network potential
(NNP) MD simulations to obtain the ADP. The calculated ADPs were compared with experimentally
reported values for rocksalt structure MgO and three thermoelectric materials, namely
argyrodite structure Ag8SnSe6 (Takahashi et al., 2024), Na2In2Sn4 (Yamada et al., 2023
) with a helical tunnel framework structure, and ThCr2Si2-type phosphide BaCu1.14In0.86P2 (Sarkar et al., 2024b
). Some have intrinsic substitutional disorder. The crystal structures of the thermoelectric
materials are visualized in Fig. 1
. The ADPs are especially interesting in the thermoelectric materials because rattling
atoms with large ADPs, which may be very anisotropic, could be effective in reducing
the and improving the thermoelectric performance.
![]() |
Figure 1 Experimentally obtained crystal structures of (a, b) Ag8SnSe6 (Takahashi et al., 2024 ![]() ![]() ![]() ![]() |
2. Formalism
Various symbols are used in the literature to describe ADPs, including U, UC, B and β, and these have not always been used consistently. Table 1 summarizes the recommendations from a subcommittee on ADP nomenclature from the International
Union of Crystallography [from Section 1.3 of Trueblood et al. (1996
)]. Here, r is the mean atom position vector and u is the displacement vector from r. The basis vector lengths of the reciprocal axes are denoted as a*, b* and c* or a1, a2 and a3.
|
The U, while the Protein Data Bank (PDB) convention adopts UC (Grosse-Kunstleve & Adams, 2002). The U and UC are identical only when the basis vectors are orthogonal to each other. ADPs can
be visualized as Oak Ridge Thermal Ellipsoid Plot (ORTEP) ellipsoids, where the semi-axes are along the three eigenvectors of U and the lengths are proportional to the corresponding eigenvalue of U (Johnson, 1965
). Details of the transformation between different ADP definitions are summarized
in Appendix A
(Section A1).
The following is the proposed procedure to obtain ADPs:
(1) Conduct MD simulations.
(2) Calculate UC for each atom from positions obtained from MD simulations.
(3) Average UC for atoms related by translational symmetry.
(4) If necessary, convert to U, change the basis vectors of some sites, then average U over sites with the same Wyckoff position.
(5) Convert to whatever ADP quantity that is convenient.
The ADPs are already standard deviations of numerous atom positions. Providing a standard deviation, often denoted using parentheses, of a standard deviation (calculated ADP) from a single simulation is mathematically not sound.
Suppose there are 1 million data points from a single MD simulation. The data points can be split into 10 sets with 100 000 data points, and the standard deviation of U obtained from 10 sets of 100 000 data points can be obtained. This standard deviation would be different from 100 sets of 10 000 data points. The standard deviation depends on how the data points are split, and thus it is not a well defined value. However, the U of the 1 million data points is the same as both the average U of 10 sets of 100 000 data points and the average U of 100 sets of 10 000 data points; thus the U obtained from all data points is a well defined value.
The averages and standard deviations of U over different runs with different atom orderings, and those of U/T over different temperatures T are meaningful.
3. Methodology
3.1. MD procedure
MD simulations were conducted using the commercially available Matlantis package from Preferred Networks with their universal PreFerred Potential (PFP) (Takamoto
et al., 2022) version 7.0.0, a NNP trained on the Perdew–Burke–Ernzerhof (PBE) generalized gradient
approximation (GGA) to density functional theory (DFT) (Perdew et al., 1996
). Diverse compounds consisting of any of all 96 elements lighter than Cm inclusive
may be modeled using the same potential.
Adopting a potential model instead of first-principles calculations results in accumulation of atom positions at a pace that is orders of magnitude faster. Too much computational resource is necessary when using first-principles calculations to obtain a sufficient amount of atom position information for derivation of ADPs with reasonable precision. On the other hand, classical force fields or potentials are only available for a limited number of systems, and fine-tuning classical force fields or potentials for each crystal in consideration requires a high level of expertise. Additionally, much time is necessary for fitting and verification. Using an off-the-shelf universal potential applicable to a wide variety of systems is a very practical approach, at least as a first attempt to newly explore a crystal.
Various universal machine-learning NNPs have been proposed in recent years (Jacobs
et al., 2025). The PFP with the Matlantis package was chosen partly because it has a long history with continuous updates by
dedicated researchers. The developers trained the potential on a proprietary data
set exceeding 60 million configurations using the VASP code (Kresse & Furthmüller, 1996
; Kresse & Joubert, 1999
). It is not overfitted to a particular published data set such as the Materials Project.
The PFP is available as a `take it or leave it' potential; the user cannot modify
it.
The canonical, or constant number of atoms, volume and temperature (NVT), ensemble
was used with a Nosé–Hoover thermostat (Nosé, 1984; Hoover, 1985
). The objective of this paper is to demonstrate direct derivation of ADPs from MD
simulations using a reasonable energy and force calculator, and fine-tuning the NNP
for individual compounds is outside the scope.
Structural and calculation details of the considered crystals are described below.
3.2. MgO simulations
The experimental ADPs of MgO were compared with theoretical values obtained using the proposed procedure. MgO takes the rocksalt structure with space-group type Fm3m (No. 225). The Mg and O occupy one each (4a and 4b), both without variable parameters. The ADPs are isotropic.
The MgO conventional
contains four Mg and four O atoms. The computational lattice parameter at 0 K is 4.2567 Å, which was used for MD simulations. Supercells of the with sizes 3 × 3 × 3, 4 × 4 × 4, 5 × 5 × 5, 6 × 6 × 6 and 7 × 7 × 7 were obtained and used.The effect of the lattice expansion on the ADPs was also studied. The lattice parameter
at 300 K is 0.15% larger than at 20–100 K, according to the experimental volume per
mol data of White & Anderson (1966). Therefore, additional calculations were conducted on 5 × 5 × 5 supercells where
the lattice parameter was increased by 0.15%.
MD simulations were conducted with a time step of 2 fs, and the number of steps was 50000 (100 ps). Positions were recorded every 100 steps, but the positions for the first 10000 steps (20 ps, 100 position recordings, hereafter equilibration steps) were discarded to account for initial shifting of atoms to attain equilibration. The temperature T was varied between 50 and 500 K in 50 K intervals.
3.3. Ag8SnSe6 simulations
The 8SnSe6 contains 30 atoms and its space-group type is Pmn21 (No. 31). The experimental lattice parameters at 300 K are a = 7.91440, b = 7.82954 and c = 11.05912 Å, which were used for MD simulations. The Ag atoms rattle and are of
interest in this study. There are five symmetrically different Ag sites. The Wyckoff
positions of sites Ag1, Ag2 and Ag3 are 4b, and those of Ag4 and Ag5 are 2a (Takahashi et al., 2024).
2 × 2 × 2 and 3 × 3 × 3 supercells were built for MD simulations of Ag8SnSe6. The time step was 2 fs, and the number of steps was 155000 and 60000, respectively (310 and 120 ps, respectively). Positions were recorded every 100 steps, but the positions for the first 20000 steps (40 ps, 200 position recordings) were discarded as equilibration steps. The number of Na position data is the same in the two calculations (16 × 2 × 2 × 2 Ag atoms × 1350 position recordings and 16 × 3 × 3 × 3 Ag atoms × 400 position recordings). The temperature T was varied between 50 and 300 K in 50 K intervals.
The U for each Ag atom in a MD simulation was obtained by taking the (co)variances of atom
positions from position recordings. The symmetrically equivalent coordinate triplets
for 4a sites are (x, y, z), (−x + 1/2, −y, z + 1/2), (x + 1/2, −y, −z + 1/2) and (−x, y, z). The U for the second, third and fourth types of atoms can be transformed to U for the first type, which is reported in this paper, using a matrix R [equation (16)] which is a diagonal matrix with diagonal components (−1, −1, 1), (1, −1, 1) and
(−1, 1, 1), respectively. For 2a sites with equivalent coordinate triplets (0, y, z) and (0, −y, z + 1/2), the diagonal matrix R of the latter has diagonal components (1, −1, 1). The final U value of each of Ag was derived by averaging the U, for each atom, over all atoms in all simulations.
3.4. Na2In2Sn4 simulations
The 2In2Sn4 contains 16 atoms and its space-group type is P212121 (No. 19). The experimental lattice parameters at 300 K are a = 6.3091, b = 6.5632 and c = 11.3917 Å, which were used for MD simulations. There are four 4a sites, where one is fully occupied by Na and the other three are shared by In and
Sn with a 1:2 ratio. Na has high anisotropy and rattles in this compound (Yamada et al., 2023).
For Na2In2Sn4, 4 × 4 × 2 supercells were built containing 128 Na sites and 384 In/Sn sites. Ten supercells, where 128 In and 256 Sn atoms were randomly assigned to the 384 In/Sn sites, were built as initial structures. The time step was 2 fs, positions were recorded every 100 steps, and 55000 steps (110 ps, 500 position recordings) were considered. The first 5000 steps (10 ps, 50 position recordings) were discarded as equilibration steps. The temperature T was varied between 50 and 300 K in 50 K intervals.
The final U was obtained similarly as in Ag8SnSe6. The symmetrically equivalent coordinate triplets for 4a sites are (x, y, z), (−x + 1/2, −y, z + 1/2), (−x, y + 1/2, −z + 1/2) and (x + 1/2, −y + 1/2, −z). The U for the second, third and fourth types of atoms can be transformed to U for the first type, which is reported in this paper, using a matrix R [equation (16)] which is a diagonal matrix with diagonal components (−1, −1, 1), (−1, 1, −1), and
(1, −1, −1), respectively.
3.5. BaCu1.14In0.86P2 simulations
The 1.14In0.86P2 contains 10 atoms and its space-group type is I4/mmm (No. 139). The experimental lattice parameters at 175 K are a = b = 4.0073 and c = 13.451 Å, which were used for MD calculations. Cu and In share a 4d site and P occupies a 4e site. Ba mainly resides at the 2a site, but the Ba site is reported as triple-split: 82% of Ba is at the 2a site, while 18% of Ba are at a 4e site very close to the 2a site (at z = ±0.02 compared with z = 0 at the 2a site).
of BaCu6 × 6 × 2 supercells were used for MD simulations, which contain 144, 164, 124 and 288 Ba, Cu, In and P atoms, respectively, or 72(BaCu1.139In0.861P2). Ten supercells were prepared as initial structures. The Cu and In atoms were randomly assigned to the 288 Cu/In sites. All Ba were initially positioned at the center of the triple-split sites, namely the 2a site, assuming that Ba can move to the 4e site, as necessary, during the initial equilibration steps. The time step was 2 fs, positions were recorded every 100 steps, and 55000 steps (110 ps, 500 position recordings) were considered. The first 5000 steps (10 ps, 50 position recordings) were discarded as equilibration steps. The temperature T was varied between 25 and 300 K in 25 K intervals.
The final U was obtained similarly as in Ag8SnSe6. The I4/mmm symmetry forces U11 = U22 and U23 = U13 = U12 = 0, although this is not exactly attained with statistical handling of atom positions. The quantities (U11 + U22)/2 (simply denoted as U11 for brevity), U33 and the isotropic Uiso = (U11 + U22 + U33)/3 were evaluated in this study.
3.6. Similarity of ADPs
An obvious way to discuss the similarities of isotropic ADPs is simply by comparing the U values. However, comparing anisotropic U with different principal axis directions is not straightforward, especially in degenerate cases where the principal axes can be taken differently (an extreme case is isotropic U).
According to Whitten & Spackman (2006), a measure of the overlap between probability density functions from two ADPs, U1 and U2, is
and the similarity index is defined as
For isotropic U with Uiso1 and Uiso2,
which is the ratio of the geometric to arithmetic mean raised to the power of 3/2.
4. Results and discussion
The computational ADPs are those of individual atoms (ADP by atom) unless noted otherwise, while the experimental ADPs are as-reported values of ADP by site.
4.1. MgO
Fig. 2 shows the Uiso versus T for Mg and O (the subscript `iso' is dropped for brevity hereon in this section).
The lines are linear regressions of
, and this trend is found over all temperature ranges and for all supercells.
![]() |
Figure 2 ADP Uiso versus temperature T of (a) Mg and (b) O in MgO. The linear regressions pass through the origin. The dashed lines represent Uiso of the 5 × 5 × 5 corrected with the Einstein model. The Lp+0.15% points are from the 5 × 5 × 5 with the lattice parameter increased by 0.15% to account for the lattice expansion at 300 K. |
The convergence of Uiso with respect to may be judged by plotting the Uiso over certain time periods in a long MD run. There are 500 position recordings from
the 50000 step (100 ps) MD simulations, and the 10 Uiso obtained from the 50(n − 1) + 1-th to 50n-th position recordings are plotted for 1 ≤ n ≤ 10 in Fig. 3. The Uiso is converged when the Uiso is roughly the same value over different n. The Uiso for n = 1 is clearly larger than those from other n, and the n = 1 and 2 results are discarded as `still in equilibration and not converged with
respect to sampling time'.
![]() |
Figure 3 Uiso for MgO based on the 50(n − 1) + 1-th to 50n-th position recordings within the 500 position recordings from a MD simulation. |
Fig. 4 plots U/T versus T. Ideally, the proportionality factor should be the same over the entire temperature
and over all supercells with the same lattice parameters. U/T for T ≤ 150 K calculations tend to be larger than for T ≥ 200 K calculations in supercells other than 3 × 3 × 3; thus the U/T averaged over T ≥ 200 K points are considered from now on. There is a 11% difference between U/T of 3 × 3 × 3 and 7 × 7 × 7 supercells for both Mg and O, while this difference decreases
to 2% between 5 × 5 × 5 and 7 × 7 × 7 supercells for both Mg and O. Therefore, the
5 × 5 × 5 is reasonably converged with regard to size.
![]() |
Figure 4 ADP Uiso divided by temperature T versus T for (a) Mg and (b) O in MgO. The horizontal lines are the average value. The Lp+0.15% points are from the 5 × 5 × 5 with the lattice parameter increased by 0.15% to account for the lattice expansion at 300 K. The horizontal lines for 6 × 6 × 6 and 5 × 5 × 5 Lp+0.15% almost overlap. |
The standard deviation of U/T normalized by the average U/T over the seven points between 200 and 500 K is discussed next. With the exception of 5% for Mg and O in the 5 × 5 × 5 the value is less than 3% for both Mg and O in the other four sizes. Increasing the lattice parameter of the 5 × 5 × 5 by 0.15% results in a ∼1% increase in the U/T; thus, considering does not result in a substantial difference in U, especially near room temperature.
Experimental determination of ADPs of both Mg and O is possible using XRD partly because
their atomic numbers, Z, are close to each other (Z = 12 and 8 for Mg and O, respectively). The reported values of U from XRD or electron diffraction at room temperature are 0.0038–0.0040 and 0.0042–0.0046 Å2 for Mg and O, respectively (Lawrence, 1973; Sasaki et al., 1979
; Tsirelson et al., 1998
).
The fitted values of 0.0037 and 0.0034 Å2 for Mg and O, respectively, at 298 K with the 5 × 5 × 5 (the same value is obtained with or without lattice parameter expansion of 0.15%)
slightly underestimate experimental values. However, applying a correction accounting
for zero-point motion based on the Einstein model increases U to 0.0042 Å2 for both Mg and O, with or without lattice parameter expansion (the correction is
discussed in detail in Section 4.5). This corrected value for O is much closer to the experimental results.
4.2. Ag8InSe6
Figs. 5(a), 5
(b) show the ADPs for Ag obtained from 2 × 2 × 2 and 3 × 3 × 3 supercells, respectively.
The isotropic ADP (Uiso) is given for Ag2, Ag3 and Ag5 while the three eigenvalues of the ADP are provided,
as U1 < U2 < U3, to be consistent with Fig. 3 of Takahashi et al. (2024
), which is reproduced as Fig. 5
(c) with the same symbols and scales as in Figs. 5
(a), 5
(b). The 2 × 2 × 2 and 3 × 3 × 3 results are very similar except for 300 K, suggesting good convergence, and have
roughly the same values as the experimental results in Fig. 5
(c). However, there are minor differences. Ag3 U1 and U2 approach 0 as temperature T→0 in calculations, but this is not the case in experiment.
![]() |
Figure 5 (a, b) Computational and (c) experimental U values of Ag8InSe6. The displayed quantities and their symbols are the same as experimental data [shown in (c)] by Yamada et al. (2023 ![]() |
The calculated 2 × 2 × 2 show qualitatively similar trends, but the exact values can differ by more than 0.01 Å2, which could be considered a very large value. Computational values can be refined
by adding more data points by increasing the simulation time. The ADPs from the two
supercells in Fig. 5
are different by less than 10% except for Ag3 U3 at 300 K, and the systematic difference between calculations and experiments cannot
be addressed by calculating more data points. Developing and/or using an energy and
forces calculator other than PFP may result in a better agreement, but this is outside
the scope of this study.
The principal semi-axis lengths of the ellipsoid, which are the three eigenvalues
of the matrix U, help explain the shape of the ellipsoid. The calculated U values are U1 < U2 << U3 and U1 << U2 < U3 in Ag1 and Ag3, respectively, resulting in cigar-like (prolate) and saucer-like (oblate)
ellipsoids, respectively, as expected from experimental results. The computational
Ag ADPs at 300 K based on the 3 × 3 × 3 are shown in Fig. 1(b). The shapes of the Ag1 and Ag3 ellipsoids are consistent with the experimentally
obtained ellipsoids in Fig. 1
(a). The calculated Ag1 U1, Ag1 U2, Ag3 U2 and Ag5 Uiso are proportional to temperature over the entire temperature range shown, as is indicated
in the linear fit that passes through the origin in Figs. 5
(a), 5
(b). Other calculated U values appear to approach U→0 in the limit T→0 with the clear exception of Ag3 U3. The Ag3 U3 value for 50 K is surprisingly larger than the 100 K value in both 2 × 2 × 2 and
3 × 3 × 3 supercells.
The Ag1 and Ag3 sites are almost threefold trigonal planar coordinated. The experimental
Ag–Se distances for Ag1 are 2.653, 2.659 and 2.707 Å, while those for Ag3 are 2.541,
2.687 and 2.774 Å. The chemical pressure from the very short Ag3–Se distance of 2.541 Å
strongly motivates Ag to rattle in the direction out of the coordination plane (Takahashi
et al., 2024), which can effectively result in a split site. This rattling mechanism caused by
`retreat from stress' is found in tetrahedrites and tennantites (Cu,Zn)12(Sb,As)4S13 (Suekuni et al., 2018
). The chemical pressure is much weaker for Ag1; thus Ag1 can be a non-split site
while a similarly coordinated Ag3 may be a split site in the direction almost normal
to the threefold coordination plane. Experimentally, none of U1, U2 and U3 of Ag3 approach U→0 in the limit T→0, which is also a hint of site splitting.
Fig. 6 is a schematic of a double-split site for qualitative discussion on the temperature
(T) dependence of U by atom. The energy is proportional to T. At very low T (T1), the atom is trapped in one of the wells, resulting in a small U. The U is very large at higher T (T2) which allows atoms to move between the wells, for example by thermal fluctuation
or tunneling, but is sufficiently low that atoms still reside in one of the wells.
This is because the atom is typically located far from the average position between
the wells. Further increasing T such that the atom is effectively in a single well (T3) results in a substantial decrease in U because the atom can now be at the center of the well. Gradually increasing T (T4) results in a gradually increasing U. The U by atom and U by site should be almost the same for T ≥ T2, but, at T1, the U by site is expected to be much larger than U by atom because atoms can occupy both wells of the site.
![]() |
Figure 6 Schematic of atom distribution in a double-well potential. |
Tables 2 and 3
show the elements of U, the eigenvalues U1, U2 and U3, U3/U1 and Uiso for all Ag sites at 200 K and 300 K, respectively, from the 3 × 3 × 3 simulations. Takahashi et al. (2024
) used anisotropic U for Ag1 and Ag3 because this dramatically improved the Rwp of experimental data at 300 K. Calculations can provide anisotropic U for all Ag simultaneously and independently without the need for repeated attempts. The calculated value of a measure of anisotropy, U3/U1, at 300 K is roughly 4 for Ag1, Ag3 and Ag5 (Table 3
). The Uiso of Ag3 is roughly double that of Ag1 and Ag5, while the number of Ag5 atoms is half
that of Ag1. Therefore, considering anisotropy of Ag5 might result in a smaller improvement
of Rwp compared with Ag1 and Ag3. Ag2 and Ag4 are less anisotropic than Ag1, Ag3 and Ag5
because of their smaller U3/U1; thus using anisotropic U would not improve Rwp significantly. The trends are similar for both 200 K (Table 2
) and 300 K (Table 3
).
|
|
The convergence of U in the 3 × 3 × 3 simulation was checked by comparing the similarity index between U from position recordings 201 to 400 and 401 to 600 (the initial 200 are discarded). The similarity index was 0.02, 0.07, 0.06, 0.04 and 0.16 for Ag1 to Ag5, respectively. This is about one order of magnitude smaller than the similarity index for Ag1 and Ag3 between the experimental and calculated values (position recordings between 201 and 600), which are 0.67 and 0.43, respectively. Therefore, the U values are regarded as converged with respect to sampling time.
4.3. Na2In2Sn4
Fig. 7 shows the eigenvalues of experimental (Yamada et al., 2023
) and calculated U, denoted as U1 < U2 < U3, for Na and In/Sn1 sites. (Uaniso_a is used instead of U3 in the original reference.) The results for In/Sn2 and In/Sn3 sites are very similar
to those of the In/Sn1 sites and are not shown. The In and Sn U are calculated separately, although one U for In and Sn combined is obtained experimentally.
![]() |
Figure 7 The eigenvalues of experimental (Yamada et al., 2023 ![]() |
The experimental and computational U values of Na are comparable with each other, and U3 of Na is significantly larger than U1 and U2 for all temperatures, implying an almost cigar-shaped spheroid [Fig. 7(a)]. The experimental U1, U2 and U3 are roughly proportional to T (linear regressions passing through the origin are shown as black lines). The calculated
U is roughly the same as the experimental U, but the details are slightly different. All of U1, U2 and U3 are almost proportional to T up to about 150 K, but consistently become larger than the linear regression of 50,
100 and 150 K values that pass through the origin (shown as red solid lines below
175 K, extrapolations to higher temperature shown with dashed lines above 175 K).
This deviation from proportionality in calculations is even more profound in In/Sn1
sites. The calculated U3 of In1 is comparable with the U3 of Na at 300 K, which is very different from the experimental results [Fig. 7(b)]. However, the experimental and computational U values are relatively close to each other at 50 and 100 K [Fig. 7
(c), which is an enlargement of Fig. 7
(b) at low U].
A large U value results in a large displacement from the equilibrium position. The standard deviation of the atom position, σ, is simply the square root of U. A U of 0.09 Å2 along a certain direction corresponds to a σ of 0.3 Å. Assuming a normal distribution, the atom is more than 3σ = 0.9 Å away from the average position for 0.3% of the time. This 0.9 Å is roughly one-third of the In/Sn–In/Sn bond length (∼2.87 Å).
In the author's previous study on LaH2.75O0.125 (Hinuma, 2025), the isotropic ADP of O, Uiso (denoted as 〈Δr2〉 in the reference), is roughly proportional to temperature below ∼0.02 Å2 but becomes much larger than what is expected from the proportionality trend above
this threshold Uiso value. The Uiso of La is proportional up to ∼0.03 Å2, which covers the entire considered range of T. This LaH2.75O0.125 is known as a very good H ion conductor. The H and O, together with vacancies, share
the same in LaH2.75O0.125; thus O may easily move away from the original site after a very small displacement
on the order of ∼0.1 Å from the equilibrium position.
The mean square displacement (MSD) of an atom trapped near the equilibrium point becomes a constant regardless of time. This MSD is the ADP when the initial position of the atom is the equilibrium point. However, the MSD is proportional to time in a diffusing atom under U values, and thereby the proposed algorithm concedes that it is not applicable in this material. Assigning atoms to the nearest site might be possible, for example by Gaussian mixture modeling or Voronoi tessellation, and atom sites must be correctly identified or provided explicitly.
and the proportionality factor is two times the dimension times the For very slowly moving atoms, the moving of some atoms away from the original equilibrium position can be detected but the cannot be derived with reasonable precision with a realistic simulation time. MD simulations find that In/Sn atoms can move away from the equilibrium point at above 175 K, based on MSD increasing with time and/or too largeInspection of atom movements during MD simulations of Na2In2Sn4 revealed significant displacements of atoms at T ≥ 200 K. The increase in U above the proportionality trend in calculations, but not in experiment, is caused by the difference in how U is obtained. Experimentally, the displacement of atoms is the distance to the nearest atom site, and the same diffusing atom may be assigned to different sites as time progresses. In contrast, the displacement in calculations is always the distance to the original atom site. The calculated U overestimates the actual U when atoms can move away from the original site.
For the record, Table 4 shows calculated U values of Na together with reported experimental values at 200, 250 and 300 K (Yamada
et al., 2023
). The calculated and experimental values for 200, 250 and 300 K are consistent with
each other, although the sign is different in some off-diagonal U coefficients.
|
4.4. BaCu1.14Cu0.86P2
Table 5 shows the calculated U of BaCu1.14In0.86P2 at 175 K together with experimental results (Sarkar et al., 2024b
) at 173 K. In the reference, the same Uiso was provided for each split Ba site, and anisotropic U was not given for Ba. Anisotropic U was provided for the Cu/In and P sites. Experimentally, the Cu/In site has the largest
U and is moderately anisotropic with U33/U11 = 1.38, while the U of P is very anisotropic with U33/U11 = 2.24 and is slightly smaller than that of Cu/In. The calculations underestimate
experimentally determined U. Notably, the calculated U33 of Cu/In and P are roughly one-half and one-third of the experimental U33, respectively.
|
Fig. 8 shows the calculated anisotropic U of Ba and Cu and isotropic Uiso of In and P. The Uiso of In and P are almost the same value for all temperatures T. All atoms are in an almost harmonic potential, with the linear regressions passing
very close to the origin although not required to do so. There were no migrating atoms.
The ±1 standard deviation over the 10 MD runs is shown. The standard deviation of
Cu is large because of the large U value and relatively large (standard deviation)/(mean average) ratio.
![]() |
Figure 8 Calculated U values by atom for BaCu1.14In0.86P2. The anisotropic U11 and U33 are given for Ba and Cu because of their large anisotropy, and isotropic Uiso is plotted for In and P with low anisotropy. The linear regressions in solid lines are not forced to pass through the origin. The range of ±1 standard deviation of U over 10 calculations is shown. |
Assuming over the temperature range, the (standard deviation)/(mean average) over all considered
temperatures was 4.6% or less for all of U11, U33 and Uiso of all elements (the mean averages and standard deviations are given in Table 6
). The correction to U from zero-point motion based on the Einstein model [details in Appendix A
(Section A2), UlowT and Tc in equations (29
) and (30
), respectively] was calculated. The correction is at most 11% for P and 4% within
other elements. The corrected U is given in Table 5
.
|
The difference in U between experiment and calculations in Table 5 arises from how the values were derived. The experimental ADPs are by site, while
the calculated ADPs in Table 5
and Fig. 8
are the averages of ADPs by atom. Therefore, the U33 by site was additionally derived using a histogram of z coordinates. Atoms with coordinates outside of −0.1 < z < 0.4 were translated to this range in integer multiples of 0.5; note that BaCu1.14In0.86P2 is a body-centered crystal.
Figs. 9(a)–9
(d) show the histograms for Ba, Cu, In and P, respectively, for T = 50, 175 and 300 K. The bin size of the z coordinate is 0.001. There are 144 atoms × 500 position recordings × 10 supercells
= 720000 total positions for Ba. Only the z ≃ 0.14 peak is shown for P (there is another peak at z ≃ −0.14, which is a mirror image around z = 0, that is not shown). The normal distribution regressions and their standard deviations
are also given, which can be used to obtain the ADP by site. The histogram for Cu
in Fig. 9
(b) cannot be described well by a single normal distribution.
![]() |
Figure 9 Distribution of z coordinates of (a) Ba, (b, e) Cu, (c) In, (d) P and (f) Cu+In combined in BaCu1.14In0.86P2 derived from the atom positions obtained from MD simulations. The atom positions were binned with 0.001 intervals of z. The points are fitted to (a–d) normal distributions (solid lines) and (e, f) superposition of three normal distributions in equation (4 ![]() |
The curve for Ba has a single peak, and the experimentally suggested triple-well scenario
with small peaks at Δz = ±0.02 (Sarkar et al., 2024b) is clearly denied. In contrast, Cu, but not In, shows a triple peak with additional
peaks at Δz ≃ ±0.025. Figs. 9
(e), 9
(f) show the histograms of Cu and Cu/In combined, respectively, for T = 50 and 175 K and the regressions
The values to be fitted are a scaling constant A, the occupancy ratio of the main peak p, the standard deviation of the peaks σ and the position of the additional peaks Δz. Equation (4) provides a very good fit, and the parameters used in fitting are given in Table
7
. Very interestingly, the p for Cu/In at 175 K is 0.811, which is the experimental occupancy of the main peak
of Ba. The Δz of Ba in the experiment is 0.021, while the Δz of Cu/In in the calculation is a very similar value of 0.026.
|
Fig. 10 plots the calculated U33 ADP by atom and by site. The site ADP is obtained as
where c is the lattice parameter and σ is the standard deviation of the histogram. The ADP was obtained for Cu only, In
only, and Cu/In combined for the 4d site. A single normal distribution and a superposition of three normal distributions
with the same σ as in equation (4
), which models a triple-split site, were calculated with Cu and Cu/In sites.
![]() |
Figure 10 Calculated U33 per atom (empty symbols) and by site (filled symbols) in BaCu1.14In0.86P2. The U33 by site is the variance of the normal distribution fit as in Fig. 8 ![]() |
In Fig. 10, all shown U33 are well described with linear fits. There is no contribution to U except for the zero-point motion that is not reflected in the calculations. The non-zero
U33 by site but zero U33 by atom at T→0 implies a significant contribution to U by site from disorder of Cu/In sites, which causes atoms to move away from the average
atom sites defined as points. This is in addition to any contributions from zero-point
motion that is not reflected in the calculations.
The single-crystal XRD data of Sarkar et al. (2024a) [Cambridge Crystallographic Data Centre (CCDC) No. 2285845] were re-refined using
a model suggested by ADP calculations, which has a triple-split Cu site and no splitting
in Ba, In and P sites. The results are summarized in Table 8
. The R1 [I > 2σ(I)] and R1 (all data) of the re-refinement are 0.0298 and 0.0318, which are better than 0.031
and 0.033, respectively, in the original reference (Sarkar et al., 2024b
). However, the ADPs in the re-refinement are not consistent with the calculated values
in Table 5
.
|
4.5. Behavior of U at T→0
The Einstein model is considered first, where atoms are each isolated in a harmonic
potential and the atoms do not explicitly interact with each other. The relation holds between an ADP component U and temperature T at sufficiently high temperatures. On the other hand, U converges to a non-zero value, UlowT, at very low temperatures where the zero-point motion cannot be ignored. A crossover
temperature Tc is defined, which is the T corresponding to UlowT assuming that the
relation holds down to 0 K. This Tc can be obtained once a combination of U and T in the
regime is obtained. [See Appendix A
(Section A2) for the mathematical details for this section.] Using the U/T proportionality factor is preferred, if available, instead of a single combination.
Experimental results of U3 for In in Na2In2Sn4 are proportional to temperature down to 90 K [see Fig. 4(b) of Yamada et al. (2023)]. Using values of U3 = 0.15 Å2 at 250 K in the figure, UlowT = 0.008 Å2 and Tc = 13 K. Therefore, effects from the zero-point motion are not relevant in the temperature
range studied in this system.
The C in the Einstein model is not proportional to T3 at T→0, although experimentally is often found. In contrast, the Debye model gives
at T→0. In the Debye model, the U at T→0 converges to a non-zero U value, UlowT_D, and
at sufficiently high temperature, as in the Einstein model. The crossover temperature
Tc_D can be defined similarly, which is 1/4 of the Debye temperature in the simplest approximation.
The UlowT_D and Tc_D are
smaller than UlowT and Tc, respectively.
The Einstein and Debye models are totally different assumptions on atom vibrations in the crystal. Different atoms can have different characteristic frequencies in different directions in the Einstein model, while there is only one Debye temperature in a crystal.
An elastic neutron scattering study reports the Debye temperature of MgO as 743 ±
8 K (Beg, 1976), corresponding to Tc_D = 186 K. The corresponding Einstein model crossover temperature, Tc, would be roughly 214 K. The U based on the Einstein model and the
proportionality factor of the 5 × 5 × 5 model are shown as blue dashed lines in Fig. 2
. The Tc for Mg and O are 187 and 244 K, respectively, which are close to the 214 K inferred
from the experimental Debye temperature.
From another perspective, the NNP MD simulations in this study can clearly identify whether the experimental non-zero U at T→0 is solely the consequence of zero-point motion in an Einstein model or whether there are contributions from something else, such as or a split site. Therefore, ADPs estimated from NNP MD act as a probe to complement experimental observations.
Why does the proportionality persist at very low T in the NNP MD simulations, but not in experiment? This paradox can be resolved easily.
The NNP MD is based on classical dynamics although the parameters are fitted to non-classical
DFT, and there is no zero-point motion in classical dynamics.
The author raises an open question. Is it ever possible for a classical MD to give a non-zero ADP at T→0 in systems where atoms are in a harmonic potential, such as in MgO?
5. Summary
The anisotropic ADPs by atom were directly derived from MD simulations taking the (co)variances over atom positions at different time steps. A universal machine-learned NNP was used to accelerate the calculations. The proposed method is applicable to systems with substitutional disorder and split sites, unlike with conventional methods using the dynamical matrix from phonon modes calculated at 0 K. The ADPs can be obtained by atom, where the (co)variance of each atom is obtained and then averaged for atoms in an atom site, or by site, where the (co)variance is calculated over all atoms at the atom site. The zero-point motion is not reflected in an ADP from classical MD simulations, but it can be estimated by extrapolation assuming an Einstein model using the proportionality factor between the ADP and temperature, if proportionality holds. The proposed method gives ADPs of MgO that are consistent with experimental values. The experimentally obtained shapes of anisotropic displacement ellipsoids of rattling atoms in thermoelectric materials Ag8SnSe6 and Na2In2Sn4 are consistent with calculated results. The possibility of splitting of the Ag3 site of Ag8SnSe6 can be detected through calculation of the ADP over different temperatures. Migration of atoms was found in Na2In2Sn4 at T ≥ 200 K, resulting in a very large ADP by atom that disagrees with the experimentally obtained, much smaller ADPs by atom site. The ADPs by atom and by site are clearly different in the thermoelectric material BaCu1.14In0.86P2, which is caused by disorder in the shared Cu/In site. The non-zero ADPs by site at T→0 in BaCu1.14In0.86P2 are a good reason why ADPs should not be referred to as `temperature factors' or `thermal ellipsoids' because there are contributions to the 0 K ADPs from disorder of Cu/In atoms other than the zero-point motion. The investigations in this study suggest the effectiveness and limitations of direct ADP derivation from MD simulations and the use of calculated ADPs as a tool complementing experimental efforts to determine the including atom displacement around atom sites.
APPENDIX A
Mathematical details
A1. Relations between different ADP definitions
The U, β and B are related by [equation (38) of Trueblood et al. (1996)]
Using a scaling matrix
The Debye–Waller factor for a diffraction vector
which is a row vector in
is[from equations (21), (24), (34) and (36) of Trueblood et al. (1996)]. Here, (e1, e2, e3) is an orthonormal basis.
One transformation matrix between the bases (a, b, c) and (e1, e2, e3) is [equation (50) of Trueblood et al. (1996)]
Coordinates of vector r in bases (a, b, c), (a*a, b*b, c*c) and (e1, e2, e3), which are denoted as (x, y, z), (ξ, η, ζ) and (ξC, ηC, ζC), respectively, transform as
[based on equation (41) of Trueblood et al. (1996)] and
[based on equations (46)–(48) of Trueblood et al. (1996)]. Matrix
was used by Trueblood et al. (1996).
The transformation between U with the basis (a*a, b*b, c*c) and UC with the orthonormal basis (e1, e2, e3) is
[based on equation (49) of Trueblood et al. (1996)].
The ADPs of an atom are often visualized as an ellipsoid. The principal semi-axes of the ellipsoid are simply the three eigenvalues of UC, and the directions of the semi-axes are those of the eigenvectors. The matrix of three eigenvectors P and the diagonal matrix with three corresponding eigenvalues Q are obtained as
The eigenvectors in the bases (a, b, c) and (a*a, b*b, c*c) are derived with equations (10) and (11
), respectively. Alternatively, principal component analysis may be used to obtain
the eigenvalues and eigenvectors of UC.
The conversion of U between two atoms with the same p and q, is given below. Let the mean positions of the two atoms p and q, and
, be related as
where R is a rotation matrix and t is a translation vector. The relation between displacement vectors and
is
The goal here is to change the basis vectors defining such that the transformed quantity
is directly comparable with U, which is accomplished by
This transformation is useful to increase the number of atoms with U values that can be easily averaged, thereby improving the reliability of results. The average U and UC of atoms that are directly comparable is simply the mean of U and UC of the atoms, respectively.
A2. Temperature dependence of the ADP in Einstein and Debye models
The ADP in the Einstein model is discussed first. A 1D harmonic potential is considered with the standard Hamiltonian
where x is the position, m is the particle mass and ω is the characteristic angular frequency. An atom is placed in this potential. Although this potential does not have terms describing explicit interactions with other atoms, this potential reflects the interactions between atoms.
The eigenvalues of this potential are
where n are non-negative integers. The corresponding variance of the position, which is also the same as the ADP, U, is
In the classical limit of continuous En, and when the equipartition theorem holds,
of the eigenvalues in equation (17and U for a certain T is
The limits of U at and
are
and
respectively, using a crossover temperature where UlowT = UhighT, namely
When a set of U and T in the regime, (U1, T1), is known,
resulting in
and
The crossover temperature tends to be higher in light atoms with small U/T atoms. Atoms with smaller U are more tightly bound to the equilibrium point.
The rightmost term in equations (29) and (30
) is when units of daltons for mass, K for temperature and Å2 for U are used. The proportionality factor between U and T may be used instead of U1/T1 if this can be obtained from fitting.
The correction factor to U arising from the zero-point motion may be written as
The function converges relatively slowly to 0 when x is increased. Its value when x = 1, 2, 3, 4, 5, 6 and 10 is 0.313, 0.082, 0.037, 0.021, 0.013, 0.019 and 0.003,
respectively. In other words, the correction factor is 30%, 10%, 3% and 1% when T/Tc is 1.023, 1.81, 3.3 and 5.8, respectively.
The Debye model is considered next. The simplest model of a 3D crystal is considered
with one atom per ). The U from the Debye model along a certain direction is, using a Debye temperature TD,
The low- and high-temperature approximations are
and
respectively. The crossover temperature where is
The UD_lowT can be obtained from a set of U and T in the regime, (U1, T1), as
The rightmost term is when units of daltons for mass, K for temperature and Å2 for U are used. According to equations (30) and (37
),
and
Note that the UD_lowT must be the same for all atoms and all directions in this simplest Debye model.
Acknowledgements
The author acknowledges Dr Naoya Ishida (AIST) for re-refinement of the published
single-crystal diffraction data of BaCu1.14In0.86P2 (CCDC No. 2285845) (Sarkar et al., 2024a) based on the suggested computational model. The author also thanks Dr Eiji Nishibori
for sharing experimental ADPs for Ag8SnSe6 used by Yamada et al. (2023), which were used to draw Figs. 1(a) and 5
(c) in this paper. The VESTA code (Momma & Izumi, 2011
) was used to draw Fig. 1
.
Funding information
This study was funded by a Kakenhi Grant-in-Aid (No. 24H00395) from the Japan Society for the Promotion of Science (JSPS).
References
Beg, M. M. (1976). Acta Cryst. A32, 154–156. CrossRef IUCr Journals Web of Science Google Scholar
Erba, A., Ferrabone, M., Orlando, R. & Dovesi, R. (2013). J. Comput. Chem. 34, 346–354. Web of Science CrossRef CAS PubMed Google Scholar
Grosse-Kunstleve, R. W. & Adams, P. D. (2002). J. Appl. Cryst. 35, 477–480. Web of Science CrossRef CAS IUCr Journals Google Scholar
Grosso, G. & Parravicini, G. P. (2014). Editors. Solid state physics, 2nd ed., pp. 391–436. Amsterdam: Academic Press. Google Scholar
Hinuma, Y. (2025). Comput. Mater. Sci. 246, 113368. CrossRef Google Scholar
Hoover, W. G. (1985). Phys. Rev. A 31, 1695–1697. CrossRef CAS Google Scholar
Iversen, B. B., Larsen, F. K., Figgis, B. N., Reynolds, P. A. & Schultz, A. J. (1996).
Acta Cryst. B52, 923–931. CrossRef ICSD CAS Web of Science IUCr Journals Google Scholar
Jacobs, R., Morgan, D., Attarian, S., Meng, J., Shen, C., Wu, Z., Xie, C. Y., Yang,
J. H., Artrith, N., Blaiszik, B., Ceder, G., Choudhary, K., Csanyi, G., Cubuk, E.
D., Deng, B., Drautz, R., Fu, X., Godwin, J., Honavar, V., Isayev, O., Johansson,
A., Kozinsky, B., Martiniani, S., Ong, S. P., Poltavsky, I., Schmidt, K. J., Takamoto,
S., Thompson, A. P., Westermayr, J. & Wood, B. M. (2025). Curr. Opin. Solid State Mater. Sci. 35, 101214. CrossRef Google Scholar
Johnson, C. K. (1965). ORTEP: a Fortran thermal-ellipsoid plot program for crystal structure illustrations. ORNL-3794. Oak Ridge National Laboratory, USA. Google Scholar
Kresse, G. & Furthmüller, J. (1996). Phys. Rev. B 54, 11169–11186. CrossRef CAS Google Scholar
Kresse, G. & Joubert, D. (1999). Phys. Rev. B 59, 1758–1775. CrossRef CAS Google Scholar
Lane, N. J., Vogel, S. C., Hug, G., Togo, A., Chaput, L., Hultman, L. & Barsoum, M.
W. (2012). Phys. Rev. B 86, 214301. CrossRef Google Scholar
Lawrence, J. L. (1973). Acta Cryst. A29, 94–95. CrossRef IUCr Journals Google Scholar
Madsen, A. Ø., Civalleri, B., Ferrabone, M., Pascale, F. & Erba, A. (2013). Acta Cryst. A69, 309–321. Web of Science CrossRef CAS IUCr Journals Google Scholar
Momma, K. & Izumi, F. (2011). J. Appl. Cryst. 44, 1272–1276. Web of Science CrossRef CAS IUCr Journals Google Scholar
Nosé, S. (1984). J. Chem. Phys. 81, 511–519. CrossRef CAS Web of Science Google Scholar
Perdew, J. P., Burke, K. & Ernzerhof, M. (1996). Phys. Rev. Lett. 77, 3865–3868. CrossRef PubMed CAS Web of Science Google Scholar
Reilly, A. M., Wann, D. A., Gutmann, M. J., Jura, M., Morrison, C. A. & Rankin, D.
W. H. (2013). J. Appl. Cryst. 46, 656–662. CrossRef CAS IUCr Journals Google Scholar
Reilly, A. M., Wann, D. A., Morrison, C. A. & Rankin, D. W. H. (2007). Chem. Phys. Lett. 448, 61–64. Web of Science CrossRef CAS Google Scholar
Sarkar, A., Porter, A. P., Viswanathan, G., Yox, P., Earnest, R. A., Wang, J., Rossini,
A. J. & Kovnir, K. (2024a). CCDC No. 2285845 experimental determination, https://doi.org/10.1039/d4ta01063a. Google Scholar
Sarkar, A., Porter, A. P., Viswanathan, G., Yox, P., Earnest, R. A., Wang, J., Rossini,
A. J. & Kovnir, K. (2024b). J. Mater. Chem. A 12, 10481–10493. CrossRef CAS Google Scholar
Sasaki, S., Fujino, K. & Takéuchi, Y. (1979). Proc. Jpn. Acad. Ser. B 55, 43–48. CrossRef CAS Google Scholar
Suekuni, K., Lee, C. H., Tanaka, H. I., Nishibori, E., Nakamura, A., Kasai, H., Mori,
H., Usui, H., Ochi, M., Hasegawa, T., Nakamura, M., Ohira–Kawamura, S., Kikuchi, T.,
Kaneko, K., Nishiate, H., Hashikuni, K., Kosaka, Y., Kuroki, K. & Takabatake, T. (2018).
Adv. Mater. 30, 1706230. CrossRef Google Scholar
Takahashi, S., Kasai, H., Liu, C., Miao, L. & Nishibori, E. (2024). Cryst. Growth Des. 24, 6267–6274. CrossRef CAS Google Scholar
Takamoto, S., Shinagawa, C., Motoki, D., Nakago, K., Li, W., Kurata, I., Watanabe,
T., Yayama, Y., Iriguchi, H., Asano, Y., Onodera, T., Ishii, T., Kudo, T., Ono, H.,
Sawada, R., Ishitani, R., Ong, M., Yamaguchi, T., Kataoka, T., Hayashi, A., Charoenphakdee,
N. & Ibuka, T. (2022). Nat. Commun. 13, 2991. CrossRef PubMed Google Scholar
Trueblood, K. N., Bürgi, H.-B., Burzlaff, H., Dunitz, J. D., Gramaccioli, C. M., Schulz,
H. H., Shmueli, U. & Abrahams, S. C. (1996). Acta Cryst. A52, 770–781. CrossRef CAS Web of Science IUCr Journals Google Scholar
Tsirelson, V. G., Avilov, A. S., Abramov, Yu. A., Belokoneva, E. L., Kitaneh, R. &
Feil, D. (1998). Acta Cryst. B54, 8–17. Web of Science CrossRef ICSD CAS IUCr Journals Google Scholar
White, G. K. & Anderson, O. L. (1966). J. Appl. Phys. 37, 430–432. CrossRef CAS Google Scholar
Whitten, A. E. & Spackman, M. A. (2006). Acta Cryst. B62, 875–888. Web of Science CrossRef CAS IUCr Journals Google Scholar
Willis, B. T. M. & Pryor, A. W. (1975). Thermal vibrations in crystallography. London: Cambridge University Press. Google Scholar
Yamada, T., Yoshiya, M., Kanno, M., Takatsu, H., Ikeda, T., Nagai, H., Yamane, H.
& Kageyama, H. (2023). Adv. Mater. 35, 2207646. CrossRef Google Scholar
This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.