Justin
Mark‡
a,
Wenhao
Zhang‡
ab,
Kazuhiko
Maeda
cf,
Takafumi
Yamamoto
d,
Hiroshi
Kageyama
e and
Takao
Mori
*ab
aInternational Center for Materials Nanoarchitectonics (WPI-MANA), National Institute for Materials Science (NIMS), Tsukuba, Ibaraki 305-0044, Japan. E-mail: MORI.Takao@nims.go.jp
bGraduate School of Pure and Applied Science, University of Tsukuba, Tsukuba, Ibaraki 305-8671, Japan
cDepartment of Chemistry, School of Science, Tokyo Institute of Technology, Tokyo, Japan
dLaboratory for Materials and Structures, Institute of Innovative Research, Tokyo Institute of Technology, Yokohama, Japan
eGraduate School of Engineering, Kyoto University, Kyoto, Japan
fLiving Systems Materialogy (LiSM) Research Group, International Research Frontiers Initiative (IRFI), Tokyo Institute of Technology, 4259 Nagatsuta-cho, Midori-ku, Yokohama, Kanagawa 226-8502, Japan
First published on 28th April 2023
The mixed-anion compounds Sn2SbS2−xSexI3 (x = 0, 0.2, 0.5) and Sn2BiS2I3 were synthesized and characterized. Rietveld refinement, elemental analysis, and diffuse reflectance measurements indicate successful substitution of Se for S and potential for bandgap tunability. Thermal conductivity measurements reveal ultralow thermal conductivities as low as 0.22 W m−1 K−1 at 573 K within this family of compounds. Such exceptionally low thermal conductivities are anomalous when compared to the previously reported heavier Pb2BiS2I3 (0.7 W m−1 K−1). Computational analysis in the form of electron localization function and Grüneisen parameter reveal Sn to be the most influential factor that controls the lattice thermal conductivity, due to its localized antibonding electrons and large contribution to structural anharmonicity.
The diverse structures of mixed-anion materials remain relatively under explored compared to single anion compounds, however, they have shown to be promising for energy applications. The irregular coordination polyhedra of mixed-anion materials have shown to be reliable building blocks for novel noncentrosymmetric materials, a prerequisite for second harmonic generation, as seen in Ba2[GaF4(IO3)2](IO3), BaGeOSe2, and Sr3Ge2O4Se3.5–7 Other examples of intriguing mixed-anion applications include ionic conductors and thermoelectrics. Na2(CB9H10)(CB11H12) demonstrates significantly improved ionic conductivity compared to the single anion parent compounds, reaching conductivities 7× greater than the well-known superionic conductor Li10GeP2S12, while BiCuSeO is a promising thermoelectric material with ultralow thermal conductivity achieving zT values > 1.8–12
Despite the promising properties and diverse structural opportunities in mixed anion materials, most work is reported for oxygen containing materials (oxyhalides, oxychalcogenides, etc.).1,3 However, a recent computational work has shown non-oxide mixed anion materials can exhibit a much more diverse set of crystal structures. This is attributed to the high bond strength of metal-oxide bonds which limit variability compared to those of sulfides, tellurides, etc.13 Thus, we seek to expand on the structure–property relationships in non-oxide mixed anion materials for future materials design.
Given the complexity which can arise in mixed-anion crystal structures, they can serve as a promising platform for low thermal conductivity materials. Indeed, some of the studied mixed-anion materials such as Pb2BiS2I3, BiCuSeO, and Bi4O4SeCl2 have exhibited exceptionally low thermal conductivities of 0.7 W m−1 K−1, 0.5 W m−1 K−1, and 0.1 W m−1 K−1 respectively.9,14,15 Thermal conductivity can be divided into two components, electronic (κe) and lattice (κl) thermal conductivities. The reduction of thermal conductivity by extrinsic methods has been extensively studied as it is easier to manipulate κl by extrinsic methods, for example, by various nano–microstructuring16–18 and/or introducing various defects.19–22 However, intrinsic thermal conductivity is closely related to the crystal structure and bonding, making it much more difficult to pinpoint and modify. Commonly, heavy elements in a complex crystal structure can lead to complex phonon branching, small phonon velocities and high rates of phonon–phonon scattering, all of them beneficial for achieving lower lattice thermal conductivity.23 Other structural factors, such as a large interatomic space, weak bonding, partial occupancy, and bonding heterogeneity can provide sources of lattice anharmonicity, further reducing the thermal conductivity.24–27
With these factors in mind, we consider the large family of non-oxide mixed-anion Tt2PnCh2I3 (Tt = Sn, Pb; Pn = Sb, Bi; Ch = S, Se) compounds for which thermal conductivity is only reported for Pb2BiS2I3.14,28–30 In general they can be described as van der Waals layered compounds with I atoms loosely bonded as bridging atoms within the inter-layer space. Site symmetry is mostly low, indicating potentially large anharmonicity. Furthermore, the large polarizability of iodine can be expected to introduce additional bonding dynamics and reduce lattice thermal conductivity.31 These interesting structural features and the relative lack of experimental results give us the motivation to investigate into these compounds.
For the phonon calculation, we relaxed the structure until the force on each atom is smaller than 1.5 × 10−5 Ry bohr−1. Interatomic force for the phonon calculation was extracted using the frozen phonon method as implemented in the phonopy packages.38 A displacement of 0.015 Å was used to generate the displacement pattern in a 4 × 1 × 3 supercell with 384 atoms. Reciprocal space was sampled using Γ point only when calculating the force.
The difficulty in experimental crystal structure determination mainly arises from the fact that X-ray scattering factors of Sn and Sb (or Pb and Bi) are nearly identical, making their atomic positions ambiguous and difficult to be truly distinguished by X-ray diffraction. However, the crystal structure is of crucial importance for our later theoretical analysis. Therefore, we attempted to determine the stability of the different possible crystal structures for Sn2SbS2I3 by comparing the energy of different relaxed structural models from first principles calculation. The structural models were chosen so that they do not include mixed occupancies (Pb2SbS2I3 LT) or split sites (reducing the split sites to fixed position), which are difficult to treat in DFT calculations. The key point here is the determination of the positions of Sn and Sb. We considered three models: (1) the original Cmcm structure with the split Sb reduced from a half occupied 8f to a fully occupied 4c site, (2) a P21/c (lower) symmetry structure with staggered occupancy of the 8f Sb site and (3) the LT-P21/c configuration of Pb2SbS2I3, reported by Doussier et al. with a Sn2SbS2I3 composition.28 We denote them as Cmcm-4c, P21/c-(Cmcm), and LT-P21/c, respectively. The Cmcm-4c and P21/c-(Cmcm) are derived from the reported Cmcm structure of Sn2SbS2I3 but treat the split of the Sb 8f sites in different ways, while LT-P21/c represents the configuration in which half the Sn positions are swapped with the Sb position compared to the reported structure. Atomic positions and the lattice parameters for each structure model were optimized using variable-cell optimization routine implemented in Quantum Espresso while keeping the symmetry unchanged. The three crystal structures considered, structure parameters and their respective phonon dispersion are shown in Fig. S7, S8 and Table S2.† We found the Cmcm-4c structure is dynamically unstable with the appearance of imaginary phonon modes occurring at most parts of the Brillouin zone, while the two structures in P21/c are dynamically stable. DFT calculations of the ground state energies shows that LT-P21/c has the lowest energy, with P21/c-(Cmcm) and Cmcm-4c being 0.38 eV and 0.64 eV higher per formula unit, respectively. The calculation results agree with the bond valence analysis given by the previous report.28 With these results, we propose that Sn2SbS2I3 shares the same P21/c structure as Pb2SbS2I3 in its ground state (Fig. 1d) and use it for our later theoretical calculations.
Powder X-ray diffraction indicates the three sample compositions are nearly identical and match well with the theoretical pattern of Sn2SbS2I3 (Fig. 2a). Upon closer comparison of the three experimental patterns, small peak shifts are observed for certain peaks in the direction of lower 2θ with increasing Se content. This indicates an expansion of the unit cells, consistent with the substitution of the larger Se in place of S. Furthermore, the magnitude of the peak shift is strongly correlated to the l index as observed in Fig. 2b, indicating the unit cell expansion is predominantly along the crystallographic c-axis. Rietveld refinement and determined unit cell parameters are shown in Fig. S1 and Table S1.† Compared to the reported single crystal unit cells of Sn2SbS2I3 and Sn2SbSe2I3, the determined unit cells for Sn2SbS2I3 and Se substituted samples are in good agreement.29 The Se substituted samples retain consistent a- and b-parameters, while a significant increase in c is observed, which is expected when comparing the pure S and Se analogs.29
![]() | ||
Fig. 2 Powder X-ray diffraction patterns of (a) the as synthesized Sn2SbS2−xSexI3 (x = 0, 0.2, 0.5) samples and (b) zoomed in view of the blue highlighted section in (a) emphasizing observed peak shifts between samples. The calculated pattern for Sn2SbS2I3 is shown in red for reference.29 |
To further ensure phase composition, energy dispersive X-ray spectroscopy (EDS) was performed. EDS reveals that although the bulk samples match the nominal stoichiometries quite well, there are secondary phases present (Fig. S2†). This may explain the abnormally high wR values of the Rietveld refinements in which some peak intensities are significantly different, despite factoring in possible preferred orientation (Fig. S1†). However, no possible overlapping binary or ternary phases (SnS, Sn2S3, Sn2SI2, etc.) could explain the difference in intensities, even based on measured EDS compositions. It is possible that the overlapping phase(s) could be related unidentified ternary or quaternary structures in this complex system.
This trend is contradictory to that observed when substituting Sn for Pb (increase in bandgap), but consistent with the substitution of Sb for Bi (decrease in bandgap) observed for related Tt2PnS2I3 (Tt = Sn, Pb; Pn = Sb, Bi) compounds.14 From an electronegativity standpoint, substitution of S for Se should shift the valence band energy higher and decrease the bandgap due to the lower electronegativity of Se, secondly, comparing the DFT relaxed structure with Sn2SbS2I3 and Sn2SbSe2I3, it is found that the difference between the two in structure mainly come from the components S(Se) and its nearest neighbours. Apart from the longer interatomic distance between S(Se) and Sn or Sb, their relative position is also distorted slightly. The change in bandgap is therefore due to both a change in electronegativity, as well as local structure, and both affect the covalent part of bonding and thus the bandgap. Such an ability to fine tune the bandgap in this family of materials may be beneficial as Pb2SbS2I3 was recently shown to be a robust material for solar cell applications.42
Thermal diffusivity measurements reveal exceptionally low thermal conductivity as low as 0.28 W m−1 K−1 for Sn2SbS2I3 at 573 K (Fig. 4a). A further reduction of thermal conductivity is observed with the substitution of Se, achieving 0.22 W m−1 K−1 at 573 K for Sn2SbS1.5Se0.5I3 (Fig. 4a). This can be expected from the increase of structural disorder and phonon scattering from substitution and defects. The measured thermal conductivity shows a slight decrease with temperature for the parent sample but a very weak temperature dependence for the Se substituted samples. This suggests that phonon–phonon Umklapp scattering is so strong that the phonon mean free path reaches the amorphous limit, and the weakly temperature dependent coherent part of thermal conductivity plays an important role.44
![]() | ||
Fig. 4 (a) High temperature thermal conductivity of Sn2SbS2−xSexI3 (x = 0, 0.2, 0.5) and Sn2BiS2I3 with respective pellet densities and (b) low temperature thermal conductivity for Sn2SbS2I3. |
To confirm such unexpectedly low thermal conductivity measurements, we conducted two further experiments. First, we synthesized another member of the Tt2PnCh2I3 (Tt = Sn, Pb; Pn = Sb, Bi; Ch = S, Se) family, namely Sn2BiS2I3 for direct experimental comparison. Using an identical synthesis procedure used for Sn2SbS2I3 similar purity product was confirmed through Rietveld refinement (Fig. S3†). Phase composition was confirmed through EDS and the optical bandgap matches well with previous reports (Fig. S4 and S5†).14 Measured thermal conductivity for Sn2BiS2I3 is shown in Fig. 4a and is of similar scale to Sn2SbS2I3, suggesting this low thermal conductivity is not specific to the Sn2SbS2−xSexI3 compositions, but can be observed in other members of the Sn containing family.
We also used a PPMS as a second method to measure thermal conductivity as a comparison. The thermal conductivity of the Sn2SbS2I3 sample at low temperature is shown in Fig. 4b. Initially, an amorphous glass-like behaviour can be observed at low temperatures, however a T4 dependence becomes dominant at higher temperatures. This suggests significant radiation loss during heating of the sample, which significantly skews thermal conductivity data above 150 K.45 By applying a correction which accounts for additional radiative losses, the data collected above 150 K significantly changes and we observe a maximum value of thermal conductivity of about 0.54 W m−1 K−1 at around 100 K, which slowly decreases with further increase in temperature (see ESI† for detail of the postprocess correction method). After applying the correction, the measured thermal conductivity is approximately 0.49 W m−1 K−1, at 300 K. While still exceptionally low this value is considerably larger than that observed through the LFA method. This inconsistency may come from the still existing underestimation of radiation loss, which is difficult to estimate correctly above 150 K and they are further enlarged by the non-optimal sample geometry (see Discussion in the ESI† for estimation of error in thermal conductivity measurements). Despite the possible errors in this temperature range, the important information revealed from the low temperature data is the absence of a well-defined Umklapp peak and that the thermal conductivity at low temperature cannot be described by the commonly used Debye model of lattice thermal conductivity including boundary scattering and phonon scattering, indicating that the assumptions used in the Debye model, such as the linear phonon dispersion and dominant acoustic phonon scattering are not sufficient to describe the actual case.46,47 Nonetheless, the rounded thermal conductivity maximum and the rather weak temperature dependence suggests the early onset of strong phonon scattering from multiple sources.
Electrical property measurements of the samples proved unsuccessful using ZEM 2 and simple measurements with a multimeter were unable to produce readings, suggesting high resistivity. This is consistent with measurements done by Islam et al. on Sn2BiS2I3 and Pb2SbS2I3, which had resistivities above 1 MΩ.14 The high resistivity indicates that the electronic contribution to the thermal conductivity is negligible and that the total thermal conductivity and lattice thermal conductivity are nearly identical.
Although defects, mixed and partial occupations that could be expected in experiment samples should play an important role in the ultralow thermal conductivity of current series of compounds, we do not attempt to cover their quantitative effects. Rather, we focus on the chemical reasoning through the anharmonicity of the structure, using Grüneisen parameter as an indicator. Grüneisen parameter is defined as the change in phonon frequency corresponding to cell expansion and is closely related to phonon anharmonicity.48,49 The Grüneisen parameter of Sn2SbS2I3 in the low frequency region is plotted in Fig. 5d. It is surprising to find the mode average Grüneisen parameter of all low frequency phonon (<1 THz) to be 2.44, with the highest value reaching over 5 for acoustic modes. This is on the same level with the calculated Grüneisen parameter of the strongly anharmonic SnSe (4.1, 2.1, 2.3 for a-, b-, and c-axis, respectively) and much larger in comparison with conventional thermoelectric compounds like PbTe (1.45).50
A microscopic picture of the anharmonicity is revealed by comparing the projected Grüneisen parameter of each atom, obtained by multiplying the mode Grüneisen parameter with the atomic component of phonon eigen-vectors, in the inset of Fig. 5d. From the projected Grüneisen parameter, it is found that Sn-involving phonon modes show high Grüneisen parameters. Fig. 5b also shows that Sn contributes to the largest phonon density of states in the lower frequency optical phonon region (0.5–1 THz). The phonon in these branches should scatter acoustic phonons effectively.47,51,52 Therefore, we can conclude that Sn atoms are the most important source of the high anharmonicity we found in the structure and lead to a low thermal conductivity. This result also explains the low thermal conductivity of the Sn containing Sn2BiS2I3 measured in the current work.
We attempt to establish the chemical reason for the high anharmonicity of the Sn atom from electronic structure calculation. Using the Bader's analysis method, we obtained the Bader's charge of each atoms Sn+1, Sb+1.85, I−0.5 and S−1.25.53–55 Incomplete charge transfer for S and I, compared with their formal charge of −2 and −1 indicates covalent character, i.e., the orbital components from Sn and Sb atoms are also important in the valence region. Indeed, the projected electronic density of state of Sn2SbS2I3 in Fig. 5a and b shows the hybridization of p orbitals in the energy window from −5 to 0 eV. We would like to point out that some density of state from s orbital of Sn can be seen just below the valence band maximum. The s component can be identified to be anti-bonding since the main bonding region of Sn s orbital is below −5 eV in energy as shown in projected density of states. This antibonding state would involve p orbital of Sn and S along their bonding direction due to the square pyramid like coordination of the Sn and stabilized due to the large interatomic space in the opposite side of S atom.56
The above anti-bonding Sn s component is exhibited in the electron localization function plotted in Fig. 6 with surface of ELF = 0.6.53,57 The crescent moon shaped region near the Sn atom with high electron localization shows the antibonding states containing Sn s orbital in real space. Using integrated local density of state (ILDOS) from −2 to 0 eV, it can be shown that this electron localized region corresponds to the anti-bonding s orbital component discussed above (Fig. S9†). It is known that such “lone pair like” localized antibonding electronic states are important in suppressing lattice thermal conductivity as they introduce restoring force strongly deviated from harmonic approximation, thus leading to the large anharmonicity observed for Sn and the entire crystal structure of Sn2SbS2I3.58–61 In comparison, Pb2BiS2I3 does not show strong localized antibonding states both in real space, as shown in Fig. 6 and in electronic density of states (Fig. S9†), highlighting the important role of the Sn antibonding s states in suppressing the lattice thermal conductivity.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3ta00609c |
‡ J. M. and W. Z. contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2023 |